You are on page 1of 335

Stanford Anesthesiology Regional Anesthesia Syllabus

The nuts and bolts of better living through sodium channel blockade

V1.6 Last Updated 10-20-2002

Stanford Anesthesiology Regional Anesthesia Syllabus


The nuts and bolts of better living through sodium channel blockade

V1.6 Last Updated 10-20-2002

Stanford Department of Anesthesiology Regional Anesthesia Syllabus Introduction


Sean Mackey, M.D.,Ph.D. Eric Amador, M.D. Welcome to the Stanford Anesthesiology Regional Anesthesia Rotation! We hope you will find this to be both an educational and entertaining month. This syllabus is meant to serve as a guide to the service, which will primarily focus on adult regional anesthesia with the majority of the emphasis on the outpatient setting; I refer you to the excellent guides produced by the OB and Pediatric groups for regional anesthesia targeting those patient groups. This regional anesthesia rotation is relatively new at Stanford, and as such, is expected to continuously undergo revision to better serve your, the patients, and surgeons needs. The goal of the regional rotation is to help you to develop a fundamental understanding and appreciation of regional anesthesia as well as the technical skills to act at the consultant level. We dont expect you to become masters of each block - that will only come with time and after performing a large number of them. We do, however, expect that by the end of the month, you will have developed a fund of knowledge that will allow you to discuss: The indications of each block including some of the published literature in support of the indication Appropriate choice of local anesthetic and adjuvants The functional anatomy and physiology associated with each block How to perform the regional block, which different approaches as appropriate Possible complications as well as their treatment

Following this rotation we encourage you to continue refining your skills as well as share your knowledge with others. Please give us feedback as to whether we are helping you to meet the goals set above. We hope you enjoy this month and develop a similar degree of enthusiasm for regional anesthesia as we have.

Educational Objectives Current Faculty, Pagers, Home numbers, things they like to talk about during those awkward moments when you are searching for something to say and you are all "taught out", etc. Faculty Sean Mackey, M.D.,Ph.D. Pager Home Number 13940 650-941-2959

V1.6 Last Updated 10-20-2002

Topics for irrelevant banter: His son Ian, functional neuroimaging, anything engineering and technology related (but not your latest computer problem unless you have exhausted all other resources), running, biking, swimming, pain, resident education issues

Eric Amador Tim Angelotti Brendan Carvalho Lindsey Vokach-Brodsky

13441 23160 650-330-1861 13980 650-468-5654 13919 408-253-3125

Topics for irrelevant banter: Rollerblading, San Diego Chargers

Topics for irrelevant banter: To be determined

Topics for irrelevant banter: To be determined

Topics for irrelevant banter: To be determined

Resident Responsibilities I believe it is a US Marines quote that goes "Prior Proper Planning Prevents Piss Poor Performance" (The 7 P's). This is certainly true for regional anesthesia. Regional anesthesia, much like any of the other anesthesia subspecialties, has its own preparations - some of them unique, and some overlapping with other specialties. Without these preparations, you can be assured that poor performance and unhappy patients, nurses, and faculty will result. I have divided up the residents responsibilities and suggestions for preparation into preoperative, intraoperative and postoperative sections to allow easier organization. Preoperative Discuss the case the night before with your attending. I know, particularly for the more senior residents, there is a tendency to skip this step for the uncomplicated ASA 1 or 2 patient in the ASC. We are expecting that call and looking forward to discussing the plan with you. Review the available literature the night before. There will be few things that annoy your attending more than showing up have no clue about the indications, anatomy, performance and complications of the block. There is no need to read the whole book, just the few pages in a regional anesthesia atlas to start. We will expect you to prepare more as the month progresses. Early morning preparation is a key to getting these cases started on time. Unfortunately, there is significantly more up front work involved with regional anesthesia than performing most GA/LMA techniques. When you first arrive, stop by the admitting desk and ask the staff to triage your patient back first and ask them to be put in the block area. They are used to hearing this and should not

V1.6 Last Updated 10-20-2002

make an issue. On Monday, this means you will miss a good portion of the AM lecture. As with the Cardiac Anesthesia Rotation, this is unfortunately unavoidable. Draw up your local anesthetics for at least the first case. Make sure you have the appropriate regional equipment available. If not, it can be obtained in the ASC and Main OR workrooms. If the block cart is not stocked, ask the techs to please stock it. Have your usual setup ready in the OR to perform GA if needed. Have enough sedation available for the patient to perform the block. A suggestion is Fentanyl 250ug, Versed 4mg. Bring your drawn up resuscitation drugs with you to the block area - just in case... Make sure oxygen , ambu bag, as well as face mask or nasal cannula are available Continually check with the nurses to help get the patient into the block area. Ideally, we would like to be placing the block 45-60 minutes before cut time. That leaves plenty of time for the local to provide surgical anesthesia (15-40 minutes) as well as a more relaxed environment for teaching so that you are not feeling rushed (remember how you felt your first few months as a CA-1?). It also leaves time to supplement the block in case it is not working adequately. Greet the patient. Talk about the anesthetic options. (Please see the Regional Anesthesia Rap section for suggestions). If you get a sense that the patient is not appropriate for a regional technique, obviously follow your clinical judgment and offer a GA. Position patient, place monitors, apply O2 Sedate patient. I believe you will find that most if not all of the regional faculty tend to heavily sedate the patient for placement of the block. We want the patient to be as comfortable as possible for the block and with little memory of the event. Perform block. With regard to performance of the block. If adequate preparation was made and the patient arrived on time, there should be plenty of time for you to place the block with teaching. However, if we are rushed for time, you may be only given 5 or 10 minutes maximum to place the block before the attending takes over. One of the primary requirements for a successful regional program is to avoid the perception by the surgeon that we are delaying the case to put in a block. And yes, we recognize that this is a teaching institution, however, unless the surgeon is an active or at least passive proponent, we will quickly find them advising the patients not to undergo regional blockade. Assess block. To do this, you must have an adequate functional knowledge of neuroanatomy. There is no substitute. Fortunately, we are not asking you to know every nerve and muscle function in the body. You will have to learn/relearn a subset including: Brachial plexus, lumbar plexus,

V1.6 Last Updated 10-20-2002

lumbosacral plexus, and dermatomes. This will include origin and termination of above as well as muscles innervated and sensory distributions. Always have Plan B, C, D ready in case the first block does not work. This may mean repeating the block at the same or different location, supplementing individual portions of the plexus blocked, having the surgeon provide local infiltration, placing a spinal or epidural, or performing a GA.

INTRAOP Before the surgeons start, you should be continuously assessing the adequacy of the block until you are convinced it will provide surgical anesthesia. This should be established with pin-prick test at least once. There are 27G dental needles available which are well suited for this. The short end of the needle is blunt and short beveled. You can use this to test without much fear of leaving a track of bloody pinpricks up and down the patient. Do not stick the patient with the other end, which has a long sharp tip. Dr. Leong uses an alternate choice of the pointy corner of an unopened alcohol swap because he is not a dentist and has no access to the dental needles. Communicate with the surgeon and the rest of the operating room team the status of the patient. You will often find that when the patient is placed on the OR table, the team expects you to perform a GA and will not start prepping the patient until they see a tub sticking out of the mouth. Let them know that they can start the prep immediately - if the block is established. If you are not absolutely convinced the block is working, but believe that it is something that can be handled with the surgeon placing some local anesthetic superficially or deep, then communicate this to the surgeon so that there are no surprises - they hate surprises. Do this in a manner that demonstrates your knowledge. For instance, for a knee arthroscopy, it is much more preferable to state "I performed a lumbar plexus and sciatic nerve block but the obturator component is not quite set up. You might have to inject a small amount of local on the medial aspect of the knee before placing your trochar", rather than, "I am not sure if this block is working, lets see what happens..." The latter does not instill much confidence. It is crucial to keep you eyes on the surgeon and the patient when they make their first cut. It is one thing to have an intubated, paralyzed patient who is somewhat light and another entirely to have a wide-awake patient who just underwent a surgical incision over an area that is not blocked. You have to be prepared to take very rapid action. While the surgery is underway, start thinking about the next case. We usually do this anyway; however, with regional anesthesia, it is particularly important to think WAY ahead. Ideally, you should be placing the block for the next patient 45-60min before the anticipated start of the surgery. This will require you to get things going upwards of 1.5 hours ahead of time, which means you

V1.6 Last Updated 10-20-2002

might be asking the facilitator resident to relieve you right after the surgeon starts. Part of this thinking ahead is communicating with the front desk to make sure the patients are brought back in a timely manner. Usually, the preop nurses have their own guidelines as to the appropriate amount of time needed to get the patient interviewed and paperwork finished. You need to communicate with them that you need the patient back and ready for you AT LEAST one hour before their anticipated surgery. Now a question that comes up is "What if I place the block early and the surgeon is delayed with the previous case?" No problem, with 0.5% bupivicaine/ropivicaine/levobupivicaine you will have 6 to 8 hours of surgical anesthesia - plenty of time unless the case is scheduled to last an unusually long time. In that case we would probably put in a continuous peripheral nerve catheter anyway. Use the facilitator wisely and efficiently. They are there specifically to relieve you to place the next block. Keep in contact with him/her and let them know when you will need them. If you call the facilitator and they are not available, communicate that with your attending rather than let time slip away. Talk with the surgeon about the next case and any special needs they have. It does you no good to plan a popliteal/saphenous block if the surgeon requires a thigh tourniquet for the Achilles tendon repair. This regional anesthetic will require a femoral/sciatic block. Let your attending know what regional topic you would like to discuss. We will have our own agenda and list of topics, but it does help if we can meet your particular interests and needs. Titrate the sedation to time for the end of the case. If you are running a propofol infusion, the patient should be wide-awake or perhaps only slightly sedated at the end of the procedure. Remember, one of the advantages of regional anesthesia is that we can fast-track the patients through the PACU. If the patient had a lower extremity peripheral nerve block, when you fill out the anesthesia postop orders, dont forget to put down in the comments section. "Pt to be non-weight bearing on operated limb for 24hrs". It's not good to ask the patient to walk out unassisted with a blocked leg. POSTOP FOLLOWUP Each patient should receive a phone call the next day to assess their experience with the block, satisfaction; time block wore off, as well as any problems or complications. This information is outlined in the Regional Anesthesia Case log, which is downloadable from one of the Stanford Anesthesia web sites at http://anesthesia.stanford.edu. Click on Resources, then Regional Anesthesia Resources. Preferably, this should be done early in the morning, before the patient follows up in the surgery clinic. It demonstrates to the patient and surgeon your dedication and

V1.6 Last Updated 10-20-2002

commitment to quality care and "ownership" of the patient. Additionally, it improves patient satisfaction and is the only way for you to get a feel for the duration of your blocks, the impact different local anesthetics have on this duration, and how the block worked for the patient in the postoperative period. Finally, it will allow you to assess the incidence, significance and follow-up of side effects and complications. Sometimes the patient is admitted to the ATU for a 23hr admission. You can hunt the patient down during one of your breaks. Just let us know the patient is in-house and we will give you extra time. If the patient had any questions or problems, please inform the attending you worked with.

Research Opportunities There are several research opportunities available for you to be involved with the goal of generating an abstract for a free trip to present it as well as shared authorship on a paper. Please discuss these opportunities with the regional faculty. Some, but not all, of the faculty interests are listed below: Sean Mackey, M.D.,Ph.D. Functional magnetic resonance imaging (fMRI) related to the understanding of pain (ask him to show you his pretty pictures), Patient outcomes related to regional anesthesia, advanced imaging techniques for regional anesthesia. Regional Anesthesia Attending Responsibilities (and you thought we would make it only apply to you) We will strive to help you meet the educational goals as set elsewhere in the document. We will make every attempt to get you out in a timely manner for the morning, lunch and afternoon breaks. Occasionally, when we are double covering two regional anesthesia rooms, we get extremely busy and may forget your break. Please gently remind us with a simple "Hey, would it be possible to get out for a bite or check my preops" and we will get you out. We will give you every chance possible to get the block in before "stealing it" as long as it does not delay the case or result in the patient turned into a pincushion. We will strive to protect you for regionally oriented cases for this month.

Regional Anesthesia Rap When you first meet your patient, they are usually anxious and unless their surgeon or the preop clinic suggested otherwise, they are expecting a general anesthetic technique You might consider with something like:

V1.6 Last Updated 10-20-2002

"Mr/Mrs/Ms Patient, I would like to talk with you about your anesthesia options for this surgery. I have my opinion about which is better, but I would like to present both of these options to you to help you make an informed decision" "The first option is general anesthesia. With this, we will place an IV in your hand or arm, give you some medication to relax you and take you down to the operating room. Once there, we will have you move over to the operating room table, place some monitors on you to measure vital signs and have you breathe some oxygen through a face-mask that will smell like a cheap shower curtain. We will then inject some medications into your IV to render you unconscious and place a flexible breathing tube, an endotracheal tube, in your trachea - your windpipe. You will be maintained unconscious by delivering anesthetic gasses through that tube. The surgery will then begin. When it is over, we will turn off the gasses, return you to consciousness, and bring you to the recovery room. We will provide you with narcotic medications during the surgery as well as after so that you are as comfortable as possible. The most frequent side effects associate with this technique is a sore throat that lasts a day or two. Additionally, some nausea and/or vomiting are not uncommon, which we will try to prevent by giving you some medications during the surgery, but can not always be prevented. Lip and dental damage are rare occurrences. " "The other option is a regional anesthetic technique. This involves places local anesthetic, or a numbing medication, directly on the nerves that cover the area to be operated on." You might ask if the patient has ever been to the Dentist. If so, you can relate the numbing feeling they had from the local to the same sensation they would have with a regional technique. I find this helpful as it gives patients a frame of reference. "After placing the block, we would bring you down to the operating room, place the monitors and then provide you with some medication to sedate you. During the surgery you can be as awake or as asleep as you want to be." I usually reassure patients that if they dont want to hear anything we can heavily sedate them so that they will not. I also point out that there is a difference between being ASLEEP and UNCONSCIOUS - asleep is what they did in their bed last night and they will be breathing on their own - unconscious will necessitate having a breathing tube placed. "The benefits of the regional technique are that since your body is not experiencing pain, I dont have to give you as much narcotic medication and anesthetic gasses to tolerate the procedure. Therefore you will have significantly less chance of nausea or vomiting, you will be clearer headed, and can probably leave the recovery room sooner. Additionally, by using a long acting local anesthetic, I can give you anywhere from 12 to 24 hours of excellent pain relief without narcotic medications - the time when pain after surgery is usually at its worst. The risks of regional anesthesia incl ude a very rare incidence of bleeding, infection,...(insert complications and relative risks for the block suggested here)." I usually tell people up front that: "Patients sometimes find it disconcerting after surgery to have a numb (insert body part) for almost a day.

V1.6 Last Updated 10-20-2002

This is an entirely natural feeling - but I have also found that, for most patients, given the choice between a numb (insert body part) with little or no pain, or having normal feeling back with pain - most would rather have a numb (insert body part)." With that last statement, I almost always see the patient nod their head in agreement. One of patients biggest fears is the thought of uncontrolled postoperative pain. You can help put that fear to rest. If they ask what will they do when the block wears off, let them know that the surgeons will be providing them with narcotic pain medication when they leave - it will be THEIR choice to take it IF they need it. Depending on the surgery, I have found they often do not need it. Obviously, don't try to memorize the above monologue. As with everything else in anesthesia, develop your own style you are comfortable with. The biggest point I want to emphasize is that many people think of general anesthesia as "going to sleep". A GA is not sleep and when you are truthful about what it is as well as reassuring them that you can make them relaxed, comfortable, and yes - even asleep, I have found the vast majority of people choose a regional anesthetic technique.

Educational Objectives At the end of the rotation you should have a working fund of knowledge of the below regional anesthesia topics. This will require some familiarity with the major scientific studies related to regional anesthesia. Additionally, we want you to be able to discuss it and defend your statements. Remember, there is an oral part of the boards and you are expected to act as a consultant when you leave the residency. Its never too early to practice your rap. Additionally you should be able to: o Effectively and independently interact with the patients, surgeons and nurses with regard to regional anesthesia topics. o o Effectively teach medical students and other residents. Communicate and execute a concise plan for regional anesthesia for your patient as well as how to rescue a failed block

V1.6 Last Updated 10-20-2002

Educational Topics Items for Daily Teaching Each of your attendings will have varying expertise and interests in these areas. Please ask each for at least one topic for the day and try to cover all the topics during the month. Many of these
topics were taken out of the ABA Content Outline. So, if you know this information, you will ace this portion of the ABA exam

A. Anatomy - Regional Anesthesia, Nerve Blocks


Understand the functional neuroanatomy and physiology of the following nerves and plexus 1. Autonomic a) stellate b) celiac c) lumbar sympathetic 2. Head and neck a) cervical plexus b) superior laryngeal c) transtracheal d) glossopharyngeal 3. Extremities a) brachial plexus b) ulnar c) radial d) median e) sciatic f) femoral g) lateral femoral h) cutaneous i) obturator j) ilioinguinal k) lumbar plexus 4. Trunk a) intercostals b) paravertebral somatic 5. Spine cervical, thoracic, lumbar, caudal a) epidural b) caudal c) intrathecal

B. Anesthetics - Local
1. Biotransformation and excretion 2. Comparison of drugs and chemical groups a) amide vs. ester b) pKa, protein binding, lipid solubility c) onset and duration of each local anesthetic d) motor/sensory differentiation of each local anesthetic 3. Ways to shorten onset or prolong duration of action a) Specifically the use of adjuvants such as: epinephrine, clonidine, and sodium bicarbonate 4. Toxicity of each local anesthetic

V1.6 Last Updated 10-20-2002

CNS cardiac allergy preservatives e) fetal 5. Neurophysiology of action potential generation, nerve conduction, and mode of action of local anesthetics

a) b) c) d)

C. Physics and Hardware


a) Principles of electrolocation of nerves b) Indications and characteristics for the commonly used needles and catheters

D. Specific Regional Anesthesia Techniques Indications, Contraindications, Anatomy, Techniques, Complications and Their Treatment
1. Brachial Plexus blocks including: o o o o o Interscalene Supraclavicular Infraclavicular Axillary Blocks at elbow, wrist, and hand

2. Superficial cervical plexus block 3. Blocks of the lumbosacral and lumbar plexus including: o o o o o o o o o o Lumbar plexus Femoral 3-in-1 Lateral femoral cutaneous Obturator Sciatic Ilioinguinal/iliohypogastric Saphenous Popliteal Ankle

4. Continuous or single shot lumbar and thoracic epidurals and spinals, as well as combined techniques 5. Paravertebral somatic nerve block 6. Continuous peripheral nerve catheter techniques for prolonged rehabilitation

V1.6 Last Updated 10-20-2002

7. IV regional blocks or Bier blocks 8. Complications of regional anesthesia and their treatment for the specific techniques including (I have included this again as it is such an important topic) Local anesthetic neurotoxicity and intravenous toxicity Epidural hematoma Post-dural puncture headache Inadvertent sub-dural or intrathecal block Postop nerve injury Premedication Patient position Monitoring and resuscitation equipment

9. General concepts for preparation

E. Principles of Perioperative Pain Management


1. Understand the concepts of transduction, transmission, modulation and perception related to pain processing 2. Understanding the pain processing pathways 3. Understand the concepts of the neuro-endocrine stress response to surgery or pain 4. Understand the concept of central hypersensitization 5. Implications of above for intra- and post-operative management. 6. Specifically the concepts of balanced analgesia, multimodality anesthesia and analgesia, costs, impact on OR efficiency, impact on periop outcomes and patient satisfaction. At the end of the syllabus you will find actual questions from the ABA exams related to regional anesthesia and pain management. As many of these questions repeat from year to year, if you can answer these.Well, you get the point.

V1.6 Last Updated 10-20-2002

REVIEW ARTICLE

Outcomes Research in Regional Anesthesia and Analgesia


Christopher L. Wu,
MD,

and Lee A. Fleisher,

MD

Department of Anesthesiology and Critical Care Medicine, *Division of Pain Medicine, and Medicine (Cardiology) and Biomedical Information Sciences and Health Policy and Management, The Johns Hopkins Hospital, Baltimore, Maryland

utcomes research evaluates the effectiveness of health care interventions in many aspects of patient care (clinical outcomes, functional health status, patient satisfaction, and economic measurements) and reflects national trends in determining the appropriateness, value, and quality of health care in the United States (1). Outcomes research incorporates diverse types of data and data analysis. Although outcomes measurements include a wide variety of patient-related assessments, anesthesiologists have traditionally focused on clinically related patient outcomes. The benefit of regional anesthesia on patient-related outcomes is controversial. Many questions involving study design, data analysis and sample size contribute to the uncertainty of the benefits of regional anesthesiaanalgesia on patient outcomes. Like those from other subspecialties in anesthesiology, investigators evaluating the efficacy of regional anesthesia have emphasized traditional, clinically oriented, patient-related outcomes. Little has been done to determine the consequences of regional anesthesia and postoperative analgesia on nontraditional patient outcomes, such as healthrelated quality-of-life (HRQL) measurements, patient satisfaction, and economic assessments. This article provides an overview of outcomes research, the current status of outcomes research in regional anesthesia, and future directions for determining the benefits of regional anesthesia and postoperative analgesia.

percentage of GNP spent on health care, the United States lags behind many other industrial nations in major indexes of health, such as life expectancy and infant mortality (1). These factors, along with the presence of significant geographical variations in clinical practice without differences in clinical outcomes, have resulted in political and economic pressures to reevaluate the appropriateness, value, and quality of health care in the United States (1).

What Are Outcomes?


Although outcomes research has become a popular topic for investigators, true outcomes research incorporates a wide variety of patient-related measurements, including those other than the more familiar clinically related assessments. In general, outcomes research involves an assessment of the effectiveness of a health care intervention on various aspects of patient benefits and includes not only clinical outcomes but also functional health status, patient satisfaction, and economic measurements (see Table 1) (1). Functional health status of a patient may be assessed by using validated instruments to measure quality-of-life, physical, psychologic, and social variables. There are few validated instruments to measure patient satisfaction with various aspects of anesthetic care, and patient satisfaction surveys must be carefully constructed and validated to ensure that specific health care interests are addressed (2). Economic measurements have become more prevalent in the anesthesia literature; however, there are a variety of economic analyses, each which may result in a different conclusion depending on the type and perspective (societal, patient, payer, provider) of analysis (3).

Outcomes Research: An Overview


Substantial increases in health care costs have contributed to the development of outcomes research in the United States. Health care costs constitute a significant percentage of the gross national product (GNP) and have increased at a rate much greater than that of inflation or overall growth in GNP. Despite the high
Accepted for publication July 7, 2000. Address correspondence and reprint requests to Christopher L. Wu, MD, The Johns Hopkins Hospital, Division of Pain Medicine, 550 N. Broadway, Suite 301, Baltimore, MD 21205. Address e-mail to chwu@jhmi.edu.

Types of Data and Data Analysis


Many types of data and data analysis may be used in outcomes research. Although a detailed discussion of all types of data and data analysis is beyond the scope of this article, a brief description of those most relevant to the evaluation of the benefits of regional anesthesia on patient outcomes will be discussed to facilitate interpretation of results and conclusions derived from outcomes research studies (see Table 2).
2000 by the International Anesthesia Research Society 0003-2999/00

1232

Anesth Analg 2000;91:123242

ANESTH ANALG 2000;91:123242

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

1233

Table 1. Types of Patient-Related Outcomes in Regional Anesthesia A. Clinically oriented or traditional outcomes 1. Mortality 2. Major morbidity Cardiovascular, coagulation, cognitive, gastrointestinal, immune, pulmonary, stress response B. Nontraditional outcomes 1. Health-related quality-of-life measurements 2. Patient satisfaction 3. Economic outcomes a. Types of costs and benefits: direct and indirect, medical and nonmedical b. Type of analysis: cost-effectiveness, cost-benefit, cost-utility c. Perspective for analysis: patient, payer, provider, societal
Adapted from Reference 3.

Prospective Data: Randomized, Controlled Trial. The randomized controlled trial (RCT) is considered by many to be the gold standard in evaluating the effect of an intervention on patient outcomes. Randomization minimizes the possibility that confounding factors may interfere with analysis of any potential association between risk factors and outcomes (4). By standardizing inclusion-exclusion criteria and treatment protocols, prospectively defining outcomes and using a placebo or accepted alternative treatment, a RCT maximizes the likelihood that outcomes obtained are the result of the intervention applied. Despite the strength of the RCT, it has several disadvantages, some of which are relevant when determining the effect and efficacy of regional anesthesia and analgesia on patient outcomes. Significant drawbacks to RCTs include the cost, time, and need for extremely large sample sizes when evaluating rare outcomes. For example, a sample size of 24,000 patients would be needed to determine if regional anesthesia would have a beneficial effect in decreasing the incidence of overall mortality by 50% (power of 80%) when compared with that from general anesthesia (5). Execution of such a study at one center would be extremely time consuming and expensive. Although multicenter trials are possible, protocol deviation and institutional differences may affect the study results. Furthermore, increasing sample size through multicenter trials may not necessarily improve statistical power (6). Other disadvantages of RCTs include ethical concerns and less external validity (applicability of the findings to a more heterogenous population). In addition, RCTs examining regional versus general anesthesia are necessarily unblinded (with exception of the Perioperative Ischemic Randomized Anesthesia Trial or PIRAT 2), allowing for the introduction of bias. Meta-analysis. Despite the increasingly common use of meta-analysis to combine and evaluate data from

various sources, meta-analysis is relatively controversial. The effect of an intervention, such as regional anesthesia, on patient outcomes may be difficult to determine as a result of the need for extremely large sample sizes. Studies in these areas are, for the most part, underpowered and may yield conflicting results. Meta-analysis, which involves strict criteria for the inclusion of studies for analysis and statistical methods specific for this type of analysis, attempts to integrate and synthesize the results from several smaller trials. There are several disadvantages of meta-analysis, one of the most significant of which is the creation of results and conclusions based on nonoriginal data from studies that may vary in study design, subject population, and outcomes criteria. Although some may argue that the heterogeneity of data reflects random error, small differences in one of several factors may affect whether a trial is acceptable to include for analysis, thus potentially altering the final results and conclusions of the metaanalysis (7). Meta-analysis may also contain publication biases (exclusion of non-English languages trials and unpublished data) (8). Although meta-analysis may be a useful tool to synthesize data from a variety of sources, the analysis and conclusions from a meta-analyses must be carefully worded (and read) in an attempt to prevent oversimplifying a complex issue (7). Finally, conclusions from meta-analysis may not correlate with that from subsequent large-scale RCTs (7). Retrospective Data: Databases. With the availability of large administrative or insurance claims databases, there have been an increasing number of studies analyzing database information in outcomes research. Database analysis and data acquisition generally cost less and require less time when compared with equally large RCTs. Databases contain information from regional and national populations, which may facilitate assessment of small frequency outcomes. In addition, information from databases reflects typical clinical practice. However, databases are retrospective in nature, may contain missing data points, and may not accurately measure the outcomes of interest. In addition, the informational content reflects the purpose of the database (mostly billing and claims) which, despite the presence of demographic and diagnosis codes, may not necessarily be useful in outcomes research. Incomplete coding may also hinder data analysis and result from limitations in the number of available diagnostic or procedural coding slots (9). Thus, there may not be enough information available to answer the hypothesis posed. Finally, large databases may be massaged to obtain significant associations between risk factors and outcomes (10). Although databases may provide an alternative form for outcomes research, especially in situations in which appropriately large RCTs would unlikely be

1234

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

ANESTH ANALG 2000;91:123242

Table 2. Randomized, Controlled Trials Versus Databases for Research in Regional Anesthesia Advantages Randomized-Controlled Trials Considered by many to be the gold standard Randomization minimizes effect of confounding factors Prospective definitions of outcomes Results likely caused by the effect of the intervention applied Databases Lower costs and time with data acquisition Data reflects regional and national populations Reflects typical clinical practice Facilitates assessment of rare outcomes Disadvantages Large sample sizes needed with rare outcomes Increases in cost and time with data collection Limited applicability to the general population Ethical concerns with randomization Retrospective data Created for billing and claims purposes Databases may be manipulated Does not provide cause-and-effect relationships

performed, it is important to remember that database analysis can only propose associations and not causation. The current format of many databases limits the type of questions that may be answered.

Table 3. Methodological Issues with Available Studies in Regional Anesthesia Inadequate sample sizes Use of surrogate endpoints Lack of incorporation of appropriate postoperative analgesic regimens Inadequate assessments of pain Limited ability to apply results of randomized, controlled trial to a more generalized population

Current Status of Outcomes Research in Regional Anesthesia


Despite some convincing data on specific organ systems (coagulation, pulmonary, gastrointestinal), the overall benefits of regional anesthesia on patientrelated outcomes is still controversial. Available studies are frequently beset by methodological issues, including study design and sample size, which contribute to the uncertainty of the effectiveness and efficacy of regional anesthesia on patient outcomes. Studies, for the most part, have only measured traditional, clinically oriented, patient-related outcomes (see Table 3).

Overview of Current Studies: Methodological Concerns


There are many methodological concerns in evaluation of the efficacy and effectiveness of regional anesthesia and analgesia on (clinically oriented) patient outcomes. One major criticism of available studies is inadequate sample sizes reflecting the small incidence of outcome events, thus affecting the ability of investigators to detect clinically significant differences. For example, approximately 3000 patients would be needed to determine if an intervention (regional anesthesia) could decrease the incidence of a surrogate endpoint (myocardial ischemia) from 30% to 20% (11). Enrollment of smaller-risk subjects or those undergoing less invasive surgery would increase sample sizes even further because of the consequent smaller incidence of the outcome event studied (1214). One of the difficulties in measuring some anesthesia-related clinical outcomes is the rarity in which these events occur (e.g., death, myocardial infarction). Because of the infrequency of such an event, it may be difficult (and expensive as a result of the

large sample size) to properly evaluate the effectiveness of an intervention on outcomes. Use of a surrogate endpoint, a related but more frequently or easily measured event, may facilitate determination of outcomes with a smaller sample size and decreased cost. Despite some controversy, anesthesiologists have commonly used surrogate endpoints, such as myocardial ischemia, hospital charges, and postoperative nausea and vomiting (15,16). Conclusions from a trial using surrogate endpoints may be valid if there is a positive relationship between the surrogate (e.g., myocardial ischemia) and true outcomes (e.g., myocardial infarction); however, it is rare that these relationships are established and validated. Occasionally, the relationship between surrogate endpoints and true outcomes have been invalidated with subsequent analysis (17). Strict validation of the relationship between a surrogate endpoint and true outcome may require a trial using a large sample size similar to that using true outcomes (18,19). Thus, surrogate endpoints may not be a reliable predictor of patient-related outcomes. Although most studies focus on the efficacy of intraoperative regional versus general anesthesia on clinically oriented patient outcomes, few trials have properly incorporated postoperative analgesic regimens into the study design. To properly determine the effects of postoperative regional analgesia on patient outcomes, an appropriate postoperative regional analgesic regimen (most likely using local anesthetics) should be administered for an adequate duration to provide maximal physiologic benefits as the incidence of some outcomes peak in the postoperative period

ANESTH ANALG 2000;91:123242

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

1235

possibly after discontinuation of regional analgesia (5,14,20). An example of an appropriately designed trial to investigate the effect of postoperative regional analgesia on outcomes is the PIRAT 2 study in which patients undergoing aortic cross-clamp procedures were randomized to one of four groups: intraoperative general anesthesia with postoperative epidural local analgesia or IV opioids, or intraoperative regional-general anesthesia with postoperative epidural local analgesia or IV opioids (21). In general, there are few trials in this area; consequently, it is difficult to determine the effect of postoperative regional analgesia per se on outcomes. In addition, many studies do not measure pain both at rest and with activity. Although regional anesthesiaanalgesia provides superior postoperative analgesia, static (at rest) pain control alone cannot improve clinically oriented outcomes despite the physiologic benefits of regional analgesia with local anesthetics. Only dynamic (with activity) pain control will potentially allow patients to participate in postoperative physiotherapy (e.g., deep breathing, ambulation), which may facilitate recovery and improvement in outcomes (22). Finally, even if a properly conducted RCT could be conducted with adequate sample sizes, it may not be appropriate to generalize results from such a structured protocol to typical clinical care in the real world, and at best, application of any conclusions from such a trial to the clinical setting would need to be made with caution (23). In general, RCTs (e.g., PIRAT 1 and 2) comparing the efficacy of regional versus general anesthesia on patient outcomes skew delivery of normal clinical care by setting management guidelines and limits on physiologic variables, such as blood pressure and heart rate. Thus, the results of regional versus general anesthesia RCTs may be difficult to generalize, as such tight control is not likely to occur in the typical clinical setting (23). However, this does not imply that the results from such RCTs would be useless. For example, detailed examination of the PIRAT 1 reveals that patients randomized to receive general anesthesia required more interventions to maintain study variables (24). As a result, regional anesthesia-analgesia may provide benefits in a less structured setting, especially in the presence of postoperative pain management protocols that may include regional analgesic techniques (23). Designing, funding, and executing a prospective, randomized trial investigating the efficacy of regional anesthesia to that of general anesthesia on patient outcomes is a difficult proposition at best, especially in the current academic environment. Multicenter trials may be possible; however, inherent problems with data collection and analysis from multiple centers and the difficulty of enrolling patients in randomized trials

with widely differing treatments may prevent completion of these trials (6,25,26). Current studies investigating this issue have methodological concerns, including inconsistent definitions for, and thus incidence of, outcomes, which may affect interpretation of data and result in potentially erroneous conclusions.

Current Outcomes Data


Clinical-Orientated Outcomes. Overall Mortality. Although many small trials comparing the efficacy of regional to general anesthesia on a variety of outcomes have been conducted, the global benefits of regional anesthesia-analgesia are controversial. However, there have been some preliminary data that attempt to address this issue. A group of investigators have performed an overview of all randomized trials comparing intraoperative regional or general anesthesia, regardless of the original outcomes of interest. Studies before January 1, 1997, were eligible, and the original study authors were contacted to confirm published data and provide additional unpublished details (27). One hundred forty-two trials with 9553 subjects were identified and analyzed on an intention-to-treatbasis. When compared with general anesthesia, regional anesthesia reduced overall mortality by approximately 30% (27). The use of regional anesthesia was associated with one fewer death per 100 patients within 30 days of randomization with more than 75% of deaths caused by pulmonary embolism, cardiac events, stroke, or infection (27). Although there was limited power to analyze certain subgroups and limitations to meta-analysis in general (see Types of Data and Data Analysis), regional anesthesia decreased the odds of deep venous thrombosis (DVT) by 44%, pulmonary embolism (PE) by 55%, transfusion by 50%, pneumonia by 39%, respiratory depression by 59%, myocardial infarction by 33%, and renal failure by 43% (27). Thus, it appears that use of intraoperative regional anesthesia globally decreases postoperative complications in a wide variety of surgical patients. Although not discussed in detail here, there are many analgesic and physiologic benefits of regional anesthesia that may explain and corroborate these findings (12). Coagulation. A hypercoagulable state occurs after surgery under general anesthesia and may be attenuated with use of regional anesthesia. Although the etiology of this hypercoagulable state is uncertain, possible mechanisms include potentiation by the stress response, endothelial damage with tissue factor activation, and synergism with inflammation (21). Postoperative hypercoagulability may lead to vasoocclusive and thromboembolic events, such as DVT, PE, and vascular graft failure, and may contribute to more

1236

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

ANESTH ANALG 2000;91:123242

than 200,000 deaths annually in the United States (12,28,29). Compared with general anesthesia, use of regional anesthesia is associated with a significant decrease in hypercoagulable-related events, especially after orthopedic and vascular surgery. Regional anesthesia clearly decreases incidence of DVT after orthopedic surgery, as documented by several randomized trials (30 36). A meta-analysis of 13 randomized trials comparing regional versus general anesthesia for repair of femoral neck fractures also confirmed that patients receiving general anesthesia had a 33% increased incidence of DVT (37). Continuation of postoperative regional analgesia with local anesthetics may also contribute to a decreased incidence of DVT (32). Although many of these trials have not concurrently used systemic DVT prophylaxis or have failed to specify presence of prophylaxis, retrospective data suggest that regional anesthesia will still provide a benefit in diminishing the incidence of DVT regardless of presence or absence of prophylaxis (38). Use of regional anesthesia is associated with a significant decrease in graft thrombosis after vascular surgery. Randomized trials have shown that the use of epidural anesthesia alone or in combination with general anesthesia decreases the incidence of graft occlusion or failure through attenuation of perioperative hypercoagulability (13,39,40). In addition, a randomized trial has shown a significantly decreased incidence of DVT in patients undergoing open prostatectomy under regional anesthesia (41). Finally, in an analysis of 18 randomized trials (27), regional anesthesia decreased the odds of DVT by 44% and PE by 55%. Gastrointestinal. Transient postoperative ileus is common after abdominal surgery under general anesthesia and may be caused by several factors, including an increase sympathetic efferent outflow from pain or stress response, postoperative use of opioids for analgesia, and spinal reflex inhibition of gastrointestinal motility (42). Use of regional anesthesia-analgesia facilitates recovery of postoperative gastrointestinal function and is associated with an earlier fulfillment of discharge criteria (42). Several randomized trials have demonstrated that, when compared with systemic opioid analgesia after general anesthesia, the use of thoracic epidural analgesia with a local anesthetic-based regimen is associated with significantly earlier return of gastrointestinal function after abdominal surgery (42 48). Two randomized trials revealed no differences between regional analgesia and systemic analgesia in return of bowel function after abdominal surgery; however, the duration of postoperative regional local anesthesiabased analgesia may have been too brief (24 hours) to provide significant physiologic benefits (14,49).

Four randomized trials have shown that, when compared with epidural opioids for postoperative analgesia, the use of epidural local anesthetics is associated with earlier return of gastrointestinal motility after abdominal surgery (42,46,50,51). It is unclear whether the use of neuraxial opioids will provide earlier return of gastrointestinal function when compared with that with systemic opioids, as the randomized trials comparing the effects of neuraxial to that of systemic opioids provide conflicting results (14,42,52). Pulmonary. There is a significant decrease in respiratory function after upper abdominal and thoracic surgery under general anesthesia as a result of inadequate analgesia, increase in upper abdominal and intercostal muscle tone, and spinal reflex inhibition of diaphragmatic function (12). Many individual trials with different study designs and analgesic regimens have been conducted to determine the effect of the analgesic regimen on a variety of pulmonary outcomes. It is difficult to draw any definitive conclusions from individual trials, as many are underpowered, use surrogate endpoints, and yield conflicting results. A meta-analysis of 48 RCTs investigating the effect of seven postoperative analgesic therapies on postoperative pulmonary function was conducted (17). Compared with systemic opioids, postoperative epidural analgesia with local anesthetics significantly decreased incidence of pulmonary morbidity, despite there being no differences in surrogate measures of pulmonary function (forced expiratory volume in 1 s, forced vital capacity, peak expiratory flow rate) between groups. Compared with systemic opioids, epidural opioids decreased the incidence of atelectasis but did not significantly diminish the incidence of pulmonary complications. There were no significant differences in analgesia or pulmonary function between lumbar or thoracic administration of opioids. The use of intercostal blocks, wound infiltration, or intrapleural analgesia was not associated with any significant improvement in pulmonary function or complications (17). Other (Cardiovascular, Stress Response, and Immune and Cognitive Function). The benefits of regional anesthesia-analgesia on outcomes with regard to other organ systems, such as the stress response, immune function, cognitive function, and cardiovascular system, are not certain. The neuroendocrine stress response after surgery under general anesthesia has been well documented and may be attenuated or even completely inhibited by the use of regional anesthesia. The stress response results in a hypermetabolic, catabolic state and may affect cardiovascular, immune, and coagulation function. Despite the many potentially detrimental effects on various organ systems, the association between the neuroendocrine stress response per se and patient outcomes is inconclusive. Thus, without establishing this relationship,

ANESTH ANALG 2000;91:123242

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

1237

any direct or indirect effect of regional anesthesiaanalgesia on patient outcomes through attenuation of the stress response cannot be determined. Postoperative immune function is diminished after surgery under general anesthesia and may adversely affect patient outcomes by contributing to the development of infections, increasing cost of care, and enhancing the possibility tumor growth and metastases (53,54). Although the etiology is not clear, stress response potentiation and perioperative administration of medications (opioids, anesthetics) may contribute to perioperative immunodepression. Regional anesthesia may preserve perioperative immune function either through attenuating immunodepression (via the stress response) or decreasing intraoperative blood loss, with a subsequent decrease in the need for perioperative opioid administration or blood product transfusions, which has been associated with immunosuppression (5558). Although the effect of regional anesthesia-analgesia on patient outcomes through preservation of immune function is not clear, two randomized trials have noted a decrease in the incidence of postoperative infectious complications in subjects receiving epidural analgesia (40,59). Furthermore, analysis of data from 14 randomized trials revealed that regional anesthesia was associated with a decreased risk of developing pneumonia or other infections (27). The effect of regional versus general anesthesia on cognitive function has been examined in several randomized trials (60 67). Only one demonstrated any advantage of regional anesthesia on postoperative cognitive function as measured by investigator interviews (59). Although most trials have shown no advantage of intraoperative regional anesthesia in the preservation of postoperative cognitive function, the effect of postoperative regional analgesia on cognitive function has not been carefully examined. There are many independent predictors for the development of postoperative delirium, including higher levels of postoperative pain (63,68 71). Because regional analgesic techniques are associated with superior analgesia when compared with systemic analgesic techniques, regional analgesia may provide some advantages with regard to postoperative cognitive function. Cardiac events, such as myocardial ischemia and infarction, congestive heart failure, ventricular arrhythmias, and sudden death, occur primarily in the postoperative period (72). Although regional anesthesiaanalgesia may provide many cardiovascular benefits by diminishing the stress response, attenuating postoperative hypercoagulability, and providing a favorable redistribution of coronary blood flow, there are no definitive conclusions concerning the effect of regional anesthesiaanalgesia on outcomes (12). Several randomized trials have not demonstrated any advantages of regional

anesthesia-analgesia on patient outcomes (5,73,74); however, one study did show an increase in myocardial ischemia on discontinuation of postoperative epidural analgesia (74). Although, analysis of 30 randomized trials revealed that there were approximately one-third fewer myocardial infarctions in those receiving regional anesthesia (27). Nontraditional Outcome Measurements. Economic Outcomes and Analysis. The effect of regional anesthesia-analgesia on economic outcomes has not been adequately examined. In general, comprehensive economic evaluation of any intervention on outcomes is difficult at best (75). Proper economic evaluation of the effects of regional anesthesia requires consideration of many issues, including types of economic measurements (e.g., cost) and analysis and the economic impact of any beneficial clinical outcomes provided by regional anesthesia-analgesia. Inappropriate measurements or analysis of economic data may result in erroneous conclusions regarding the economic impact of regional anesthesia-analgesia. Examining economic data can be complicated, because there are several types of economic analysis, including type of analysis (cost-effectiveness, costbenefit, cost-utility), types of costs and benefits (direct and indirect, medical and nonmedical), and perspective for analysis (patient, provider, payer, societal) (3). RCTs, which may be appropriate for addressing clinical hypothesis, may not be suitable for answering economic questions, as many RCTs include contemporaneous economic evaluation without appropriate or proper sample size calculations, descriptive statistics (including confidence intervals), or formal methods of statistical inference (76,77). Often, cost data are highly skewed, and the use of standard statistical methods may result in misleading results (76). The use of nonparametric analysis, such as bootstrapping, requires no assumptions with regard to the shape of the sampling distribution and is especially useful when working with ratios (e.g., cost-effectiveness or cost-utility analysis) where the numerator and denominator may have different distributions (77,78). The cost of a specific service incorporates the total resources used to provide that service (3). Clinicians commonly equate cost with what are considered direct hospital costs, or typical operational expenditures (e.g., equipment and medications) associated with routine clinical care. Although direct hospital costs are an important component of total cost, this concept of cost is not complete, as the total cost of a service should also include direct nonhospital or medical costs and assessment of indirect costs, which may include lost income, lost opportunities, decreased productivity, and indirect morbidity and mortality (3). Because unsubstantiated and unreliable statements regarding costs and economic analysis of interventions occur frequently in the literature, comprehensive

1238

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

ANESTH ANALG 2000;91:123242

economic data and appropriate evaluation of an intervention (such as regional anesthesia) for subsequent studies is important as clinical guidelines or health policy may be based on these data (79 82). To date, there has been no comprehensive examination of costs (direct or indirect) associated with use of regional anesthesia and analgesia. Although direct costs associated with regional analgesia may intuitively seem greater than that of systemic analgesics, regional anesthesia-analgesia may provide economic benefits through indirect costs, such as decreasing indirect costs, patient morbidity and mortality, or length of stay. For example, preliminary data from a meta-analysis of 142 randomized trials revealed that mortality was decreased by approximately 30% in patients receiving regional anesthesia when compared with those receiving general anesthesia (27). Regional anesthesia also decreased patient morbidity by decreasing risk of development of DVT, PE, transfusion, pneumonia, respiratory depression, myocardial infarction, and renal failure (27). Other randomized trials and meta-analysis also demonstrated a decrease in pulmonary complications and vascular graft failure (13,17,39). Regional anesthesia-analgesia may also provide economic benefits by decreasing length of stay through control of postoperative pain or physiologic benefits of a local anesthetic-based regional analgesic technique. Although postoperative pain per se is not an independent predictor of inpatient length of stay, inadequate control of postoperative pain is one of the leading reasons for readmission after ambulatory surgery (83). Randomized trials reveal that the physiologic benefits of thoracic epidural analgesia with local anesthetics facilitate fulfillment of discharge criteria significantly sooner than those receiving systemic opioids in patients undergoing colectomies (42,84). Thus, regional anesthesia-analgesia may significantly improve clinically oriented outcomes and decrease length of stay which may, in turn, confer beneficial economic outcomes. HRQL Measurements. Although commonly used in other specialties, HRQL measurements have not been widely used to assess the effects of regional anesthesia-analgesia. HQRL generally consists of validated instruments that assess physical, psychologic, and social variables. Data obtained from these instruments may be statistically analyzed. Although there are many validated instruments available for measuring HQRL, some studies may not incorporate those which are appropriate or standardized, which may cast some doubt as to the validity of their results. There are several reasons why HRQL measurements may not have been prevalent in the anesthesiology literature. Health care providers, like anesthesiologists, who interact with patients on a more acute basis are generally more familiar with traditional

outcomes measurements (e.g., mortality and morbidity). The application of HRQL measurements to daily clinical situations may be difficult to conceptualize, as there has been little research on models describing the relationships between clinical variables and HRQL measurements and the effect of clinical interventions on HRQL (85). Finally, HRQL measurements have not been applied in an acute setting, such as in the postoperative period, and have been generally used over a longer time frame (more than one month). Thus, widespread use of HRQL measurements is hindered by lack of information with regard to their value and benefits in an acute clinical setting and perception of measurements as being soft or unscientific (86). The effects of regional anesthesia-analgesia on HRQL measurements have not be extensively examined. Preliminary data suggest that regional anesthesia-analgesia as part of a multimodal approach to postoperative rehabilitation will facilitate earlier recovery of HRQL measurements assessed at three and six weeks after colorectal surgery (87). It is unclear whether using previously validated HRQL instruments to assess patient recovery in the immediate postoperative period (within one week after surgery) is appropriate. The nonspecific nature of some HRQL measurements and the possibility that the expected postsurgical decrease in general functional status may overwhelm any observable HRQL differences between postoperative analgesic treatments and may discourage the use of current validated HRQL measurements in the immediate postoperative setting (88,89). However, regional analgesia generally provides superior postoperative analgesia, and higher levels of postoperative pain may affect HRQL measurements by interfering with activity and sleep (90). The quality of analgesia is an important outcome per se and may directly improve physical and functional domains of HRQL. Thus, several issues need to be addressed before HRQL measurements can be meaningfully applied in the acute setting, and until that time, no definitive conclusions can be formed on the effect of regional anesthesia-analgesia on HRQL measurements. Patient satisfaction is an important measure of outcome and has become more significant as health care organizations use it as a measure of quality or as a part of marketing services. Measuring patient satisfaction may seem intuitively simple; however, poor survey design and use of unstandardized ratings may lead to erroneous conclusions. Many surveys of satisfaction lack refinement and cannot distinguish between satisfaction with the item of interest or with their overall surgical or hospital care (2). Because patient satisfaction reflects complicated psychological factors, its accurate assessment requires a psychometric methodology that can produce valid and reliable multidimensional instruments capable of capturing its true dimensions (2).

ANESTH ANALG 2000;91:123242

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

1239

Most studies investigating patient satisfaction and regional anesthesia-analgesia have not used validated instruments for measuring satisfaction. Patient satisfaction of regional anesthesia has been examined (91 93), but the effect of regional anesthesia-analgesia on patient satisfaction has not been fully evaluated. Although there are many factors that may influence patient satisfaction, increased ratings of postoperative pain are associated with decreased levels of patient satisfaction (90,94). By providing superior postoperative analgesia, regional analgesia may potentially favorably affect patient satisfaction; however, it may be difficult to differentiate between satisfaction with pain control per se and other aspects of medical care (95). Outcomes Research in Regional Anesthesia and Analgesia: Future Directions. The future of outcomes research in regional anesthesia-analgesia lies in the evaluation of the effect of regional anesthesia-analgesia on economic and nontraditional (HRQL and patient satisfaction) outcomes. Establishment of any potential economic or other patient-related benefits of regional anesthesia-analgesia is especially important in light of decreasing reimbursement for postoperative pain services. Another potential area for research is the evaluation of the contribution of postoperative regional analgesia per se to patient outcomes. Finally, adoption of an evidence-based approach to decision-making may facilitate the use of regional anesthetic techniques in the perioperative period. A paradigm for future investigation of the effect of regional anesthesia-analgesia on patient outcomes is illustrated in Figure 1. Movement Toward Nontraditional Outcomes. Despite the overwhelming use of clinically oriented measurements, we are beginning to see increasing use of validated HRQL and patient satisfaction instruments to evaluate the efficacy of regional anesthesiaanalgesia on patient outcomes. Validation and clinical incorporation of HRQL measurements in the acute perioperative setting may reveal advantages of regional anesthesia-analgesia, especially in the postoperative setting where regional analgesia may improve HRQL and patient satisfaction. Although there are several methodological issues that may hinder widespread use of these measurements in the acute postoperative setting, adoption of HRQL and patient satisfaction assessments in clinical trials may add another dimension to the efficacy of regional anesthesiaanalgesia on patient outcomes. Economic Evaluation of Regional Anesthesia and Analgesia. Although anesthesia providers may influence up to 5% of total health care costs in the United States (96), there is currently a lack of information regarding the effect of regional anesthesia-analgesia on economic outcomes. The relationship between the use of regional anesthesia and postoperative regional analgesia and potential beneficial economic outcomes, in part through improvements in clinically oriented

Figure 1. Paradigm for the investigation of regional anesthesia and postoperative regional analgesia on patient outcomes.

outcomes, and possibly decreases in length of stay, should be established. It is important that appropriate economic data presentation and analysis be incorporated into future studies, because administrators, policy-makers, and managed-care insurers may base clinical or economic policy decisions on these data. Differentiating the Contribution of Postoperative Analgesia to Outcomes. The focus of current trials has been to elucidate potential benefits of intraoperative regional anesthesia on clinically oriented patient outcomes. Despite some data suggesting improvement in outcomes (97), the role of postoperative regional analgesia per se on traditional and nontraditional patient outcomes has not been extensively investigated, even though many clinically oriented outcomes, such as myocardial ischemia or infarction and PE, may peak in the postoperative period. Part of the difficulty in teasing out the contribution of postoperative regional analgesia on patient-related outcomes lies in the fact that many postoperative processes begin in the intraoperative period (e.g., DVT) and continue into the postoperative period. Further research is needed on the effect of postoperative regional analgesia per se on various outcomes including HRQL, patient satisfaction, length of stay and other clinically related outcomes. In addition, the role of regional analgesia as an important component of a multimodal approach to promote postoperative patient recovery by facilitating early ambulation and enteral nutrition and diminishing complications needs to be further elucidated (22). Evidence-Based Approach to Decision-Making. Evidence-based medicine (EBM) refers to the integration of the best available evidence from research and application to clinical policy and practice. The relatively recent trend in using EBM in an attempt to provide the best care for an individual patient has been facilitated by easier access to research evidence (e.g., MEDLINE via the Internet). Despite some drawbacks (98 100), EBM may soon become a reality in daily clinical practice. In addition, EBM may play a significant role in the development of clinical practice guidelines. It is imperative

1240

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

ANESTH ANALG 2000;91:123242

that investigators continue to properly investigate the efficacy of regional anesthesia-analgesia on various outcomes as practitioners and policy-makers critically examine at the benefits of regional anesthesia-analgesia in clinical practice.

Conclusions
Outcomes research in regional anesthesia-analgesia has traditionally focused on clinically oriented outcomes. There is significant data to demonstrate the benefits of regional anesthesia-analgesia on overall mortality and certain organ systems (coagulation, pulmonary, gastrointestinal). The efficacy of regional anesthesia-analgesia on nonclinical outcomes, such as economic, patient satisfaction, and HRQL assessments, is not clear at this time. A trend toward more global assessments of patient-related outcomes and differentiating the contribution of postoperative regional analgesia per se will provide new opportunities in regional anesthesia-analgesia outcomes research as we begin the next millennium.

References
1. Orkin FK. Application of outcomes research to clinical decision making in cardiovascular medicine. In: Tuman K, ed. Outcome measurements. Baltimore: Williams & Wilkins, 1999:39 66. 2. Fung D, Cohen MM. Measuring patient satisfaction with anesthesia care: a review of current methodology. Anesth Analg 1998;87:1089 98. 3. Fleisher LA, Mantha S, Roizen MF. Medical technology assessment: an overview. Anesth Analg 1998;87:1271 82. 4. Cummings SR, Ernster V, Hulley SB. Designing a new study. I. Cohort studies. In Hulley SB, Cummings SR, eds. Designing clinical research. Baltimore: Williams & Wilkins, 1988:6374. 5. Bode RH, Jr, Lewis KP, Zarich SW, et al. Cardiac outcome after peripheral vascular surgery: comparison of general and regional anesthesia. Anesthesiology 1996;84:313. 6. Fisher DM. How do we obtain outcome data: randomized clinical trials versus observational databases. In: Tuman K, ed. Outcome measurements. Baltimore: Williams & Wilkins, 1999: 2337. 7. LeLorier J, Gregoire G, Benhaddad A, et al. Discrepancies between meta-analyses and subsequent large randomized, controlled trials. N Engl J Med 1997;337:536 42. 8. Gregoire G, Derderian F, Le Lorier J. Selecting the language of the publications included in a meta-analysis: is there a Towel of Babel bias? J Clin Epidemiol 1995;48:159 63. 9. Iezzoni LI. Assessing quality using administrative data. Ann Intern Med 1997;127:666 74. 10. Mills JL. Data torturing. N Engl J Med 1993;329:1196 9. 11. de Leon-Casasola OA, Lema MJ. Intensive analgesia reduces postoperative myocardial ischemia [letter]? Anesthesiology 1992;77:404 5. 12. Liu S, Carpenter RL, Neal JM. Epidural anesthesia and analgesia: their role in postoperative outcome. Anesthesiology 1995;82:1474 506. 13. Christopherson R, Beattie C, Frank SM, et al. Perioperative morbidity in patients randomized to epidural or general anesthesia for lower extremity vascular surgery. Anesthesiology 1993;79:42234. 14. Hjortso NC, Neumann P, Frosig F, et al. A controlled study on the effect of epidural analgesia with local anaesthetics and morphine on morbidity after abdominal surgery. Acta Anaesthesiol Scand 1985;29:790 6.

15. Fleisher LA. . . . but, is suppression of postoperative ST segment depression an important outcome? Anesth Analg 1997; 84:709 11. 16. Fisher DM. Surrogate endpoints: meaningful not! Anesthesiology 1999;90:355 6. 17. Ballantyne JC, Carr DB, deFerranti S, et al. The comparative effects of postoperative analgesic therapies on pulmonary outcome: cumulative meta-analyses of randomized, controlled trials. Anesth Analg 1998;86:598 612. 18. Prentice RL. Surrogate endpoints in clinical trials: definition and operational criteria. Stat Med 1989;8:431 40. 19. Buyse M, Molenberghs G. Criteria for the validation of surrogate endpoints in randomized experiments. Biometrics 1998; 54:1014 29. 20. Bois S, Couture P, Boudreault D, et al. Epidural analgesia and intravenous patient-controlled analgesia result in similar rates of postoperative myocardial ischemia after aortic surgery. Anesth Analg 1997;85:12339. 21. Rosenfeld BA. Benefits of regional anesthesia on thromboembolic complications following surgery. Reg Anesth 1996;21: S9 12. 22. Kehlet H. Multimodal approach to control postoperative pathophysiology and rehabilitation. Br J Anaesth 1997;78: 606 17. 23. Beattie C, Roizen MF, Downing JW. Cardiac outcomes after regional or general anesthesia: do we know the question [letter]? Anesthesiology 1996;85:12079. 24. Christopherson R, Glaven NJ, Norris EJ, et al. Control of blood pressure and heart rate in patients randomized to epidural or general anesthesia for lower extremity vascular surgery. J Clin Anesth 1996;8:578 84. 25. Mingus ML, Levitan SA, Bradford CN, Eisenkraft JB. Surgical patients attitudes regarding participation in clinical anesthesia research. Anesth Analg 1996;82:3327. 26. van der Berg L, Lobatto RM, Zuurmond WW, et al. Patients refusal to participate in clinical research. Eur J Anaesthesiol 1997;14:2879. 27. Schug SA. Is regional anesthesia better than general anesthesia? In: Syllabus of the 24th annual meeting of the American Society of Regional Anesthesia, Philadelphia, May 6, 1999: 62 4. 28. Kakkar VV. Low molecular weight heparins: prophylaxis of venous thromboembolism in surgical patients. Semin Hematol 1997;34:9 19. 29. Clagett GP, Anderson FA, Heit J, et al. Prevention of venous thromboembolism. Chest 1995;108:312S34S. 30. Modig J, Borg T, Karlstrom G, et al. Thromboembolism after total hip replacement: role of epidural and general anesthesia. Anesth Analg 1983;62:174 80. 31. Modig J, Borg T, Bagge L, Saldeen T. Role of extradural and of general anaesthesia in fibrinolysis and coagulation after total hip replacement. Br J Anaesth 1983;55:6259. 32. Jorgensen LN, Rasmussen LS, Nielsen PT, et al. Antithrombotic efficacy of continuous extradural analgesia after knee replacement. Br J Anaesth 1991;66:8 12. 33. McKenzie PJ, Wishart HY, Smith G. Long-term outcome after repair of fractured neck of femur: comparison of subarachnoid and general anaesthesia. Br J Anaesth 1984;56:5815. 34. McKenzie PJ, Wishart HY, Gray I, Smith G. Effects of anaesthetic technique on deep venous thrombosis: a comparison of subarachnoid and general anaesthesia. Br J Anaesth 1985;57: 8537. 35. Davis FM, Quince M, Laurenson VG. Deep vein thrombosis and anaesthetic technique in emergency hip surgery. BMJ 1980; 281:1528 9. 36. Davis FM, Laurenson VG. Spinal anaesthesia or general anaesthesia for emergency hip surgery in elderly patients. Anaesth Intensive Care 1981;9:352 8. 37. Sorenson RM, Pace NL. Anesthetic techniques during surgical repair of femoral neck fractures: a meta-analysis. Anesthesiology 1992;77:1095104.

ANESTH ANALG 2000;91:123242

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

1241

38. Eriksson BI, Ekman S, Baur M, et al. Regional block anaesthesia versus general anaesthesia: are different antithrombotic drugs equally effective in patients undergoing hip replacement retrospective analysis of 2354 patients undergoing hip replacement receiving either recombinant hirudin, unfractionated heparin or enoxaparin [abstract]. Thromb Haemost 1997;77: PS1992. 39. Rosenfeld BA, Beattie C, Christopherson R, et al. The effects of different anesthetic regimens on fibrinolysis and the development of postoperative arterial thrombosis. Anesthesiology 1993;79:435 43. 40. Tuman KJ, McCarthy RJ, March RJ, et al. Effects of epidural anesthesia and analgesia on coagulation and outcome after major vascular surgery. Anesth Analg 1991;73:696 704. 41. Hendolin H, Mattila M, Poikolainen E. The effect of lumbar epidural analgesia on the development of deep vein thrombosis of the legs after open prostatectomy. Acta Chir Scand 1981;147:4259. 42. Liu SS, Carpenter RL, Mackey DC, et al. Effects of perioperative analgesic technique on rate of recovery after colon surgery. Anesthesiology 1995;83:757 65. 43. Seeling W, Bruckmooser KP, Hufner C, et al. No reduction in postoperative complications by use of catheterized epidural analgesia following major abdominal surgery. Anaesthesist 1990;39:33 40. 44. Jayr C, Thomas H, Rey A, et al. Postoperative pulmonary complications: epidural analgesia using bupivacaine and opioids versus parenteral opioids. Anesthesiology 1993;78:666 76. 45. Ahn H, Bronge A, Johansson K, et al. Effect of continuous postoperative epidural analgesia on intestinal motility. Br J Surg 1988;75:1176 8. 46. Scheinin B, Asantila R, Orko R. The effect of bupivacaine and morphine on pain and bowel function after colonic surgery. Acta Anaesthesiol Scand 1987;31:161 4. 47. Wattwil M, Thoren T, Hennerdal S, Garvill JE. Epidural analgesia with bupivacaine reduces postoperative paralytic ileus after hysterectomy. Anesth Analg 1989;68:353 8. 48. Bredtmann RD, Herden HN, Teichmann W, et al. Epidural analgesia in colonic surgery: results of a randomized prospective study. Br J Surg 1990;77:638 42. 49. Wallin G, Cassuto J, Hogstrom S, et al. Failure of epidural anesthesia to prevent postoperative paralytic ileus. Anesthesiology 1986;65:2927. 50. Thoren T, Wattwil M. Effects on gastric emptying of thoracic epidural analgesia with morphine or bupivacaine. Anesth Analg 1988;67:68794. 51. Thorn SE, Wickborn G, Philipson L, et al. Myoelectric activity in the stomach and duodenum after epidural administration of morphine or bupivacaine. Acta Anaesthesiol Scand 1996;40: 773 8. 52. Rawal N, Sjostrand U, Christoffersson E, et al. Comparison of intramuscular and epidural morphine for postoperative analgesia in the grossly obese: influence on postoperative ambulation and pulmonary function. Anesth Analg 1984;63:58392. 53. Davey PG, Nathwani D. What is the value of preventing postoperative infections? New Horizons 1998;6:S64 71. 54. Ben-Eliyahu S, Yirmiya R, Liebeskind JC, et al. Stress increases metastatic spread of a mammary tumor in rats: evidence for mediation by the immune system. Brain Behav Immunol 1991; 5:193205. 55. Tonnesen E, Brinklov MM, Christensen NJ, et al. Natural killer cell activity and lymphocyte function during and after coronary artery bypass grafting in relation to the endocrine stress response. Anesthesiology 1987;67:526 33. 56. Amato AC, Pescatori M. Effect of perioperative blood transfusions on recurrence of colorectal cancer: meta-analysis stratified on risk factors. Dis Colon Rectum 1998;41:570 85. 57. Hole A, Unsgaard G. The effect of epidural and general anaesthesia on lymphocyte function during and after major orthopedic surgery. Acta Anaesthesiol Scand 1983;27:135 41.

58. Shir Y, Raja SN, Frank SM, Brendler CB. Intraoperative blood loss during radical retropubic prostatectomy: epidural versus general anesthesia. Urology 1995;45:9939. 59. Cuschieri RJ, Morran CG, Howie JC, McArdle CS. Postoperative pain and pulmonary complications: comparison of three analgesic regimens. Br J Surg 1985;72:495 8. 60. Hole A, Terjesen T, Breivik H. Epidural versus general anaesthesia for total hip arthroplasty in elderly patients. Acta Anaesthesiol Scand 1980;24:279 87. 61. Riis J, Lomholt B, Haxholdt O, et al. Immediate and long-term mental recovery from general versus epidural anesthesia in elderly patients. Acta Anaesthesiol Scand 1983;27:44 9. 62. Chung F, Meier R, Lautenschlager E, et al. General or spinal anesthesia: which is better in the elderly? Anesthesiology 1987; 67:4227. 63. Berggren D, Gustafson Y, Eriksson B, et al. Postoperative confusion after anesthesia in elderly patients with femoral neck fractures. Anesth Analg 1987;66:497504. 64. Williams-Russo P, Sharrock NE, Mattis S, et al. Cognitive effects after epidural vs general anesthesia in older adults. JAMA 1995;274:44 50. 65. Asbjorn J, Jakobsen BW, Pilegaard HK, et al. Mental function in elderly men after surgery during epidural analgesia. Acta Anaesthesiol Scand 1989;33:369 73. 66. Ghoneim MM, Hinrichs JV, OHara MW, et al. Comparison of psychologic and cognitive functions after general or regional anesthesia. Anesthesiology 1988;69:50715. 67. Nielson WR, Gelb AW, Casey JE, et al. Long-term cognitive and social sequelae of general versus regional anesthesia during arthroplasty in the elderly. Anesthesiology 1990;73:11039. 68. Marcantonio ER, Goldman L, Mangione CM, et al. A clinical prediction rule for delirium after elective noncardiac surgery. JAMA 1994;271:134 9. 69. Marcantonio ER, Juarez G, Goldman L, et al. The relationship of postoperative delirium with psychoactive medications. JAMA 1994;272:1518 22. 70. Dyer CB, Ashton CM, Teasdale TA. Postoperative delirium: a review of 80 primary data-collection studies. Arch Intern Med 1995;155:4615. 71. Lynch EP, Lazor MA, Gellis JE, et al. The impact of postoperative pain on the development of postoperative delirium. Anesth Analg 1998;86:7815. 72. Mangano DT, Browner WS, Hollenberg M, et al. Association of perioperative myocardial ischemia with cardiac morbidity and mortality in men undergoing noncardiac surgery. N Engl J Med 1990;323:1781 8. 73. Baron JF, Bertrand M, Barre E, et al. Combined epidural and general anesthesia versus general anesthesia for abdominal aortic surgery. Anesthesiology 1991;75:611 8. 74. Garnett RL, MacIntyre A, Lindsay P, et al. Perioperative ischaemia in aortic surgery: combined epidural/general anaesthesia and epidural analgesia versus general anaesthesia and i.v. analgesia. Can J Anaesth 1996;43:769 77. 75. Drummond MF, Richardson WS, OBrien BJ, et al. Users guide to the medical literature. XIII. How to use an article on economic analysis of clinical practice: are the results of the study valid? JAMA 1997;277:15527. 76. Barber JA, Thompson SG. Analysis and interpretation of cost data in randomised controlled trials: review of published studies. BMJ 1998;317:1195200. 77. Campbell MK, Torgerson DJ. Bootstrapping: estimating confidence intervals for cost-effectiveness ratios. QJM 1999;92: 177 82. 78. Mennemeyer ST, Cyr LP. A bootstrap approach to medical decision analysis. Health Econ 1997;16:7417. 79. Balas AE, Kretschmer RA, Gnann W, et al. Interpreting cost analyses of clinical interventions. JAMA 1998;279:54 7. 80. Baltussen R, Ament A, Leidl R. Making cost assessments based on RCTs more useful to decision-makers. Health Policy 1996; 37:163 83.

1242

REVIEW ARTICLE WU AND FLEISHER REGIONAL ANESTHESIA AND OUTCOMES

ANESTH ANALG 2000;91:123242

81. Baltussen R, Leidl R, Ament A. Real world designs in economic evaluation: bridging the gap between clinical research and policy-making. Pharmacoeconomics 1999;16:449 58. 82. Hornberger J, Wrone E. When to base clinical policies on observational versus randomized trial data. Ann Intern Med 1997;127:S697703. 83. Gold BS, Kitz DS, Lecky JH, Neuhaus JM. Unanticipated admission to the hospital following ambulatory surgery. JAMA 1989;262:3008 10. 84. Steinbrook RA. Epidural anesthesia and gastrointestinal motility. Anesth Analg 1998;86:837 44. 85. Wilson IB, Cleary PD. Linking clinical variables with healthrelated quality of life: a conceptual model of patient outcomes. JAMA 1995;273:59 65. 86. Cleary PD, Greenfield S, McNeil BJ. Assessing quality of life after surgery. Control Clin Trials 1991;12:189S203S. 87. Carli F, Klubien K, De Angelis R, et al. An intensive versus graded perioperative management program for recovery after colorectal surgery: preliminary results on quality of life [abstract]. Reg Anesth 1998;23:S11. 88. Kantz ME, Harris WJ, Levitsky K, et al. Methods for assessing condition-specific and generic functional status outcomes after total knee replacement. Med Care 1992;30:MS240 52. 89. Pocock SJ. A perspective on the role of quality-of-life assessment in clinical trials. Control Clin Trial 1991;12:257 65S. 90. McNeill JA, Sherwood GD, Starck PL, Thompson CJ. Assessing clinical outcomes: patient satisfaction with pain management. J Pain Symptom Manag 1998;16:29 40.

91. Tetzlaff JE, Yoon HJ, Brems J. Patient acceptance of interscalene block for shoulder surgery. Reg Anesth 1993;18:30 3. 92. Waters JH, Leivers D, Maher D, et al. Patient and surgeon satisfaction with extremity blockade for surgery in remote locations. Anesth Analg 1997;84:773 6. 93. Papanikolaou MN, Voulgari A, Lykouras L, et al. Psychological factors influencing the surgical patients consent to regional anaesthesia. Acta Anaesthesiol Scand 1994;38:60711. 94. Jamison RN, Ross MJ, Hoopman P, et al. Assessment of postoperative pain management: patient satisfaction and perceived helpfulness. Clin J Pain 1997;13:229 36. 95. Hester NO, Miller KL, Foster RL, Vojir CP. Symptom management outcomes: do they reflect variations in care delivery systems? Med Care 1997;35:NS69 83. 96. Johnstone RE, Martinec CL. Costs of anesthesia. Anesth Analg 1993;76:840 8. 97. Capdevila X, Barthelet Y, Biboulet P, et al. Effects of perioperative analgesic technique on the surgical outcome and duration of rehabilitation after major knee surgery. Anesthesiology 1999;91:8 15. 98. Haynes B, Haines A. Barriers and bridges to evidence based clinical practice. BMJ 1998;317:273 6. 99. Kenny NP. Does good science make good medicine? Incorporating evidence into practice is complicated by the fact that clinical practice is as much art as science. Can Med Assoc J 1997;157:33 6. 100. Kerridge I, Lowe M, Henry D. Ethics and evidence based medicine. BMJ 1998;316:11513.

ECONOMICS

AND

HEALTH SYSTEMS RESEARCH

SECTION EDITOR RONALD D. MILLER

Which Clinical Anesthesia Outcomes Are Important to Avoid? The Perspective of Patients
Alex Macario,
MD, MBA*,

Matthew Weinger,

MD,

Stacie Carney,

BA,

and Ann Kim,

BA

*Departments of Anesthesia and Health Research and Policy, Stanford University Medical Center, Stanford; Department of Anesthesiology, University of California San Diego and the San Diego Veterans Affairs Healthcare System, San Diego; and Stanford University, Stanford, California

Healthcare quality can be improved by eliciting patient preferences and customizing care to meet the needs of the patient. The goal of this study was to quantify patients preferences for postoperative anesthesia outcomes. One hundred one patients in the preoperative clinic completed a written survey. Patients were asked to rank (order) 10 possible postoperative outcomes from their most undesirable to their least undesirable outcome. Each outcome was described in simple language. Patients were also asked to distribute $100 among the 10 outcomes, proportionally more money being allocated to the more undesirable outcomes. The dollar allocations were used to determine the relative value of each outcome. Rankings and relative

value scores correlated closely (r2 0.69). Patients rated from most undesirable to least undesirable (in order): vomiting, gagging on the tracheal tube, incisional pain, nausea, recall without pain, residual weakness, shivering, sore throat, and somnolence (F-test 0.01). Implications: Although there is variability in how patients rated postoperative outcomes, avoiding nausea/vomiting, incisional pain, and gagging on the endotracheal tube was a high priority for most patients. Whether clinicians can improve the quality of anesthesia by designing anesthesia regimens that most closely meet each individual patients preferences for clinical outcomes deserves further study. (Anesth Analg 1999;89:6528)

n most industries, the quality of the product is assessed by the customer (1). Patients are customers of anesthesia service. Therefore, a logical step in perioperative healthcare is to determine what patients value, then tailor the anesthetic to meet each patients requirements. Many anesthesiologists already seek such preferences by asking, for instance, whether the patient would rather be awake (i.e., regional anesthesia) or asleep (i.e., general anesthesia) for a surgical procedure. The quality of medical decisions, patient satisfaction, and clinical outcomes can be improved by eliciting such patient preferences (2 4). The highest quality anesthetic (and related postoperative outcomes) for any patient may depend on a subjective assessment of his or her level of well being in different health states (expressed as preferences for those clinical anesthesia outcomes). For example, the

Funded in part by a FAER/Hoechst Marion Roussel, Inc/Society for Ambulatory Anesthesia Clinical Research Starter Grant from the Foundation for Anesthesia Education and Research (to AM). MW participated in this study as part of the Stanford Fellowship in the Management of Perioperative Services. Accepted for publication April 21, 1999. Address correspondence to Alex Macario, MD, MBA, Department of Anesthesia (H3580), Stanford University Medical Center, Stanford, CA 94305-5115, Address e-mail to amaca@leland.stanford.edu.

choice of an opiate to relieve postoperative pain may actually reduce the quality of the recovery period of a postoperative patient who considers nausea more objectionable than pain. In this patient, a less emetogenic, nonopioid analgesic may provide the patients desired postoperative outcome. Knowing how patients prioritize clinical anesthesia outcomes will help anesthesiologists to customize care. How patients rank the relative importance of avoiding low-morbidity, yet common anesthesia outcomes, such as nausea or shivering, is unknown. For example, it is unknown whether patients perceive a sore throat after anesthesia as less desirable than being somnolent after anesthesia, or whether patients consider avoiding postoperative nausea to be more important than pain relief. Clinicians may use the term outcome to mean the results of patient care, such as an intermediate end point or adverse event. Donabedian (5) defined outcome more broadly as a change in a patients. . . health status that can be attributed to antecedent health care. This definition certainly applies to surgical outcomes that can affect the long-term health of a patient. However, in anesthesia for routine surgery, except in the case of an anesthetic disaster, anesthesiologists may seldom be able to influence more than patient comfort during the perioperative period.
1999 by the International Anesthesia Research Society 0003-2999/99

652

Anesth Analg 1999;89:6528

ANESTH ANALG 1999;89:6528

ECONOMICS AND HEALTH SYSTEMS RESEARCH MACARIO ET AL. PATIENT PREFERENCES FOR ANESTHESIA OUTCOMES

653

For purposes of this study, we used the phrase clinical anesthesia outcome to refer to adverse clinical events associated with anesthesia. Anesthesiologists are unable to predict which common, low-morbidity anesthesia outcomes are of highest importance to a particular surgical patient (6). Patient preferences for clinical outcomes are difficult to discern without informing patients about the expected outcomes of the procedure and asking them about their specific preferences in a structured manner (7). Davies and Ware (8) suggested that most patients have the knowledge base (more health information is being made available to patients) to make such judgments. The goals of this study were to survey patients to: 1) rank order their preferences, from most to least important, for avoiding specific clinical anesthesia outcomes; and 2) quantify any variability in how surgical patients perceive common anesthesia side effects. Because there is no gold standard for asking patients about their subjective judgments of the value of avoiding acute conditions (e.g., nausea) that characterize emergence from anesthesia, we used two separate techniques (used by health economists)priority ranking and relative value scalesto study patient preferences.

income, education, marital status, work history, inpatient or outpatient surgery) and previous experience with side effects of anesthesia; 2) a rankings section; and 3) a relative value section (explained below). The order of the assessments was the same for all patients. The questions and outcome descriptions were designed to flow from previous questions. Each question expressed one idea (i.e., no question contained and), and no question was phrased in a negative tense (i.e., not or neither). Patients were asked to rank (order) 10 possible postoperative outcomes from their most undesirable to their most desirable outcome. Patients were given the following written and verbal instructions:
We want to determine your preferences for each of the following possible outcomes of anesthesia care (i.e., which ones you think are better or worse than the others). Please carefully read each of the following descriptions of outcomes you could experience in the recovery room after your anesthesia and surgery. Assume that each situation described is equally likely. While it is impossible to know how long each condition will last, assume that each will last for an equal length of time. Rank each of these postoperative outcomes in relation to each other from 1 to 10 from the most undesirable (1) to the most desirable (10).

Methods
The study took place at Stanford University Medical Center, a university- and community-affiliated university hospital, and was approved by the Stanford Human Subjects Committee. A comprehensive list of clinical anesthesia outcomes was developed from a computerized literature search (MEDLINE) for 1986 1997 using the following term: anesthetic outcome, complications. This yielded 100 published studies (a sample of these studies includes References 9 16) that were read by AM to generate a complete list of clinical anesthesia outcomes. This survey study did not include all possible outcomes, as that would have required giving patients an excessively long questionnaire. Rather, the complete list was reviewed, and nine items were selected (to represent a range of severity) for study. We then developed simple descriptions (25 45 words) of the clinical outcomes. The descriptions were reviewed and edited by four senior board-certified anesthesiologists in the anesthesia department for perceived validity and accuracy (see Table 1 for the actual language used to describe each of the outcomes). The descriptions reflected a constellation of symptoms with a focus on a particular outcome. A normal outcome, or side effect-free recovery, was included as 1 of the 10 outcomes studied. The survey instrument was organized into three parts: 1) standard demographic items (age, sex, race,

To determine the value of each outcome relative to the other outcomes, respondents were asked to assign 100 hypothetical dollars across the outcomes: more dollars were to be assigned to the less desirable outcomes. Patients were given the following written and verbal instructions:
Distribute the $100 according to your preferences such that the more money you spend on a condition, the less likely that it will occur. Thus, you should spend more money on outcomes you most want to avoid. Important: You must spend all of your $100 (and no more than that).

The actual dollar allocations assigned to a particular outcome were used determine the relative value of each outcome. If the patients assigned more than a total of $100, the values for each outcome were standardized to 100. A random number generator was used to select which patients would be asked to participate in this study. We aimed to obtain 100 completed surveys. A research assistant trained in preference assessments research methodology was available to answer any questions a patient had while completing the survey instrument. After the formal anesthesia evaluation and patient education sessions in the preoperative evaluation anesthesia clinic, patients completed and returned the survey anonymously to a mailbox. The preanesthetic visit and patient education process is standardized by the preoperative clinic. This standardization was not specifically confirmed for each patient who participated in this study. Per our usual

654

ECONOMICS AND HEALTH SYSTEMS RESEARCH MACARIO ET AL. PATIENT PREFERENCES FOR ANESTHESIA OUTCOMES

ANESTH ANALG 1999;89:6528

Table 1. Description of Postoperative Clinical Anesthesia Outcomes Outcome Nausea Recall without pain Gag on endotracheal tube Shivering Vomiting Residual weakness Somnolence Sore throat Normal Pain Description You are lying on your side, awake and aware of your surroundings in the recovery room. You are extremely queasy, as if you were seasick on a boat in rough seas. The least movement makes the nausea worse. You become aware of your surroundings in the recovery room and realize that you were awake during the surgery. You remember lying on the operating room table, unable to move or talk while the surgical procedure was underway. You are lying on your back, alert and aware that you are in the recovery room. You have a breathing tube in your windpipe, which makes it more difficult to breathe and causes you to gag. It is impossible to speak. You are lying on your back, alert and aware that you are in the recovery room. Your entire body is shivering uncontrollably so that you are unable to hold a cup of water or speak clearly. You are lying on your side, awake and aware of your surroundings in the recovery room. You feel waves of nausea and are throwing up. Your abdominal and chest muscles ache from vomiting. You are lying on your back, alert and aware that you are in the recovery room. You are so weak that you can not move any of your muscles. You can blink your eyes, but speaking is almost impossible and you feel short of breath. You are in the recovery room and are drifting off to sleep even though you want to wake up and go home. You are unable, despite your best effort, to stay awake long enough to tell the nurse how you are feeling. You are lying on your back, alert and aware that you are in the recovery room. Your throat is sore and your voice is hoarse, as if you had laryngitis. You are lying on your back, alert and aware that you are in the recovery room. You have no pain or nausea, feel good, and are ready to go home. You are lying on your back, awake and aware of your surroundings in the recovery room. Your surgical incision really hurts, as if a knife was stabbing you. Movement makes the pain worse, and no position seems to make it better.

practice, all patients provided consent for general anesthesia, even if a regional anesthetic was likely, in case general anesthesia was required. Patients 18 yr gave their written informed consent before beginning to complete the survey. Patients were eligible for the study if they were scheduled to undergo surgery either in the outpatient surgery center or in the main tertiary hospital surgery suite. Patients unable to speak or read English or who had cognitive disabilities were excluded. To gain insight into the internal validity of the instrument, we analyzed the association (i.e., correlation) between the relative value data and the ranking data for each outcome. One would expect that the relative value assignment ($0 $100) for an outcome to correspond with the ranking of that outcome (17). In other words, the less desirable the outcome by rank, the more dollars ($0 $100) that should be assigned to avoid the outcome. One would also expect that the normal (or side effect-free) outcome should be ranked 10 (highest) and would have the lowest relative value (fewest dollars) assigned. Two-way analysis of variance of ranking and relative value data, followed by Newman-Keuls tests for multiple comparisons, was used to evaluate the statistical significance of the two outcomes (18). Correlation between the rank data and the importance scores were calculated by using Pearsons correlation coefficients.

Subgroup analyses were performed. For example, it was hypothesized that patients who have actually experienced a particular outcome would rate it differently than patients who have not. The MannWhitney U-test was used to determine whether the rank or relative value data were different for patients who had experienced a particular outcome compared with those who had not.

Results
One hundred ninety-five surveys were distributed. One hundred thirty patients returned the survey. Twenty-nine of the surveys were returned but were incompletely completed and so were excluded from the data analysis. Thus, 101 patients completed the questionnaire (see Table 2 for demographic characteristics of patients). Clinical characteristics of the survey participants are summarized in Table 3. Sixty-two of the patients reported that they had previously experienced at least one of the outcomes studied. In this patient population, vomiting was the least desirable outcome by both the ranking methodology and the relative value methodology (F-test 0.01) (Table 4). The relative value scores suggested, for instance, that relief of nausea was 56% (i.e., 11.82/7.60) more important that relief of shivering.

ANESTH ANALG 1999;89:6528

ECONOMICS AND HEALTH SYSTEMS RESEARCH MACARIO ET AL. PATIENT PREFERENCES FOR ANESTHESIA OUTCOMES

655

Table 2. Demographic Characteristics of the Survey Participants Age (yr) Sex (male/female) Marital status Single Married Widowed/divorced Ethnicity Caucasian African-American Hispanic Other Years of schooling completed after kindergarten 12 (did not finish high school) 12 14 16 16 Household income $50,000 $50,000 45 16 (1983) 40/61 26 60 14 81 4 3 11 2 8 31 28 32 32 69

Table 3. Clinical Characteristics of the Survey Participants Type of surgery planned Urologic Cardiac Neurosurgical Otolaryngological General Orthopedic Gynecological Other H/o previous surgery D/C home day of surgery expected Clinical outcomea Vomiting Gagging on endotracheal tube Pain Nausea Recall without pain Shivering Residual weakness Sore throat Somnolence
Values represent percentages of patients. n 101. H/o history of, D/C discharge. a Experienced by 62 patients.

15 7 12 5 12 37 7 5 77 38 32 5 76 60 2 46 34 48 31

Values are mean sd (range) or number of patients. n 101.

The results showed internal consistency. Ranking and relative value data were positively and significantly correlated (r2 0.69, P 0.0001). There was appreciable interindividual variability among patient preferences for different anesthesia outcomes (Table 5). Previous experience with a certain anesthesia outcome was not related to a patients ranking of outcomes. For example, patients who had experienced nausea ranked nausea similarly to those patients who had not had experienced nausea. Patients studied were asked to list other outcomes that they had experienced after surgery and anesthesia. No single clinical outcome (e.g., dizziness, fainting, infection, urinary retention) was suggested by more than one respondent. All 101 respondents ranked the normal outcome after anesthesia as most desirable and allocated $0 to it.

Table 4. Ranking and Relative Value of Anesthesia Outcomes Outcome Vomiting Gagging on endotracheal tube Pain Nausea Recall without pain Residual weakness Shivering Sore throat Somnolence Normal Rank 2.56 0.13 2.97 0.15 3.46 0.2 4.02 0.17 4.85 0.26 5.34 0.17 5.36 0.20 8.02 0.11 8.28 0.11 10.00 Relative valuea 18.05 1.09 17.86 1.43 16.96 1.59 11.82 0.87 13.82 1.58 7.99 0.8 7.60 0.6 3.04 0.26 2.69 0.25 0

Values are mean sem. a This means that, for example, patients assigned $18.05 of $100 to avoid vomiting.

Discussion
For clinicians, it is important to know how patients perceive clinical outcomes, then to design the anesthetic to minimize the incidence or severity of those anesthesia-related outcomes that a particular patient feels are most important to avoid. Clinicians may make anesthetic regimen decisions based partly on what they believe is important medically and partly on their perceptions of what an average, or typical, patient would want to have as an ideal outcome after

anesthesia. We used two separate preference assessment tools to determine how patients rank (from most severe to least severe) common, low-morbidity outcomes associated with anesthesia. Patients rated vomiting as most undesirable, followed (in order) by gagging on the tracheal tube, incisional pain, nausea, recall without pain, residual weakness, shivering, sore throat, and somnolence. Because serious adverse outcomes from anesthesia are rare, improvements in the quality of anesthesia care may come from addressing these more common side effects. Given the variability in how patients responded, it is difficult to know a priori which clinical anesthesia outcomes are of highest concern for any given patient. Thus, it may be useful to actively engage patients (as part of the preoperative evaluation

656

ECONOMICS AND HEALTH SYSTEMS RESEARCH MACARIO ET AL. PATIENT PREFERENCES FOR ANESTHESIA OUTCOMES

ANESTH ANALG 1999;89:6528

Table 5. Percentage of Patients Who Gave an Anesthesia Outcome a Particular Rank Rankings Outcome Vomiting Gagging on endotracheal tube Pain Nausea Recall without pain Shivering Residual weakness Sore throat Somnolence Normal
a

First 24 22 21 6 20 1 7 0 0 0
a

Second 31 18 17 19 6 6 5 0 0 0

Third 22 24 17 15 6 8 10 0 0 0

Fourth 17 20 12 17 10 16 8 1 1 0

Fifth 5 13 14 19 10 20 15 1 2 0

Sixth 1 3 11 21 18 23 19 3 3 0

Seventh 0 1 9 4 14 17 26 14 15 0

Eighth 1 0 0 0 17 10 11 81 79 100

Of the patients, 24% ranked vomiting as their least desirable outcome.

and informed consent process) to identify, for example, their three most important clinical outcomes, then tailor the anesthetic to address these preferences. Interestingly, we found no measurable differences in opinion about the relative severity of outcomes between patients who reported no personal experience with a particular outcome and those who had experienced the outcome during a previous anesthetic. This may support the validity of the descriptions used in the study. Further investigations are required to include other outcomes not evaluated in this study and to further understand whether patients who have had unpleasant outcomes after a previous anesthetic tend to rate that outcome as being most important to avoid during a subsequent anesthetic. Our results showing the importance to patients of avoiding nausea are consistent with an earlier study.1 In a study (20) of 800 patients focusing on patients knowledge and attitudes about anesthesia, patients reported their highest level of concern for (in order) being able to wake up after surgery, postoperative pain, becoming paralyzed, having pain medications available, waking up in the middle of surgery, and postoperative nausea. We also showed that failure to wake up from an anesthetic (brain injury or dying during surgery) is a primary concern of patents. Although the rate of this adverse outcome is very low and further improvements in this end point may be difficult to obtain or measure, anesthesiologists should also address patient concerns surrounding rare but catastrophic events. Patients who experience an adverse clinical anesthesia outcome may perceive different effects on their state of well-being. In other words, although two patients may both experience nausea, their perception of the impact of nausea on their quality of life (as measured by how patients rank outcomes relative to one another, as done in this study) may be quite different.
1 Orkin F. What do patients want [abstract]? Anesth Analg 1992; 74:S225.

For example, Nease et al. (2) found that patients suffering from angina with similar functional limitations varied considerably in their tolerance of their symptoms. These authors recommended that medical management of angina should be based mainly on the preferences of the patient. Similarly, in a study of terminally ill patients, Danis et al. (4) recommended that the use of life-sustaining medical therapy should be guided primarily by patient preferences. Some of the observed variability in how patients rank any particular outcome may be due to measurement error. However, the high correlation (r2 0.69) between the two ranking techniques may support the validity of the rank order of clinical outcomes we obtained. The current study was not powered to study whether demographic variables (e.g., age or sex) or timing (preoperatively or postoperatively) of the survey affected responses. We have also undertaken a larger study to measure whether the presence of preoperative symptoms (e.g., would a person experiencing preoperative pain as a result of the surgical diagnosis have a different priority about the postoperative outcome?) or the type of surgery (e.g., if one patient was to undergo a major cancer operation and another a minor diagnostic procedure) is correlated with importance of outcomes. Monitoring the incidence over time of key clinical outcomes, such as those rated highly by patients in this study, may be a more useful measure of clinical quality than other quality measurement instruments, such as patient satisfaction scores. Patient satisfaction scales may not be fine enough to detect changes in the quality of clinical care by an anesthesia group. Patient satisfaction relies on a standard or expectation against which care is compared (21). Because this expectation of what the anesthesia experience will be can differ among patients, satisfaction may not be a reliable or valid way of detecting changes in care. In the setting of perceived risk (anesthesia), satisfaction ratings are dominated by a sense of relief (22).

ANESTH ANALG 1999;89:6528

ECONOMICS AND HEALTH SYSTEMS RESEARCH MACARIO ET AL. PATIENT PREFERENCES FOR ANESTHESIA OUTCOMES

657

This study focused on clinical anesthesia outcomes, rather than other aspects of caresuch as the affect of care (how nice providers are to patients), the environment of care (how attractive the facility is), or the timeliness of care (whether the surgery started on time). In fact, these other aspects of care may be more noticeable and important to patients than the clinical outcomes about which physicians may be concerned. For example, one study suggested that friendliness of the operating room staff is the primary determinant of patient satisfaction with outpatient surgery (23). However, prioritizing the numerous nonclinical outcomes associated with anesthesia was beyond the scope of the present study. Patient valuation of different outcomes is necessary for economic studies in anesthesia. Because anesthesia drugs and interventions almost always have side effects, clinicians and administrators must make tradeoffs among options with regard to desirable and undesirable properties. To optimize patient care, it is necessary to quantify how patients value these various outcomes. The relative value data (fraction of 100) suggest, for instance, that vomiting is almost 6 times (18.05/3.04) more undesirable than a sore throat or that relief of nausea is 56% (11.82/7.6) more important than relief of shivering. These data may help to complete economic analyses of anesthetic interventions that make tradeoffs among anesthesia outcomes. As in most studies in healthcare, including clinical trials, the current patient sample depended on patients willingness to participate. Respondents may have differed from the general population in an unpredictable number of attributes that could bias the data. The potential for selection bias was minimized by using a sampling strategy intended to represent a wide range of age, income, and surgical procedures. However, most patients who completed the survey were well educated. Some socioeconomic groups may not be able to complete accurately the ranking or relative value questions. We were unsuccessful in completing a follow-up study of the nonresponders to either improve the response rate or evaluate whether the responders are drawn from the same population as the nonresponders. This may have biased our results. The expectations of patients also tend to have a cultural component. This study was performed in the United States, and all patients had medical insurance to pay for healthcare costs, which may have affected how the patients responded. In countries in which medical care is not available, tolerance for lowmorbidity outcomes such as we studied may be assessed differently by patients fortunate enough to be treated. It is unlikely that any one patient will have experienced (and be able to rank based on actual experience) all outcomes under study. In addition, the outcome

descriptions we used were chosen by investigators in consultation with other anesthesia providers. Wording from patients may yield more accurate data (24,25). Expressed patient preferences may be influenced by the way questions are phrased, and further studies are required to refine this methodology. Patients undergoing surgery are fearful of experiencing adverse side effects from anesthesia. Asking patients explicitly to define their preferences can be part of the informed consent process. This is also consistent with patient autonomy, allowing patients to influence treatment decisions once the alternatives have been explained. On initiating this study, there was some concern that, by virtue of making postoperative adverse outcomes more explicit, patients would become more fearful or worried about their upcoming surgery. In fact, this happened in only a few patients and was managed by further conversation with the nurse educator or the physician. However, some patients did decline to participate in the study because of their concerns about making adverse outcomes more explicit. We have learned that the benefits of a better educated patient, along with knowledge about each patients preferences for different outcomes, may outweigh the risks. An important component of improving the quality of healthcare is that relevant patient information, including patient preferences and expectations, be incorporated into clinical care decisions. However, a review of the understanding of patients attitudes toward anesthesia suggests that there is substantial variation in the quantity and nature of information given to patients preoperatively about their anesthetic care (26). In this study, we provided some indication of patients relative preferences for anesthesia outcomes. Although there is substantial variability in patient preferences for postoperative outcomes, avoiding postoperative nausea/vomiting seems to be a high priority for most patients. Data obtained from physician and patient interaction on patient preferences may guide anesthesiologists to choose the anesthesia regimen that results in the highest value to each patient by best meeting his or her preferences. Whether clinicians can customize care based on elicited preferences, such as was done in this study, and improve the quality of anesthesia care deserves further study.

References
1. Laffel G, Blumenthal D. The case for using industrial quality management science in health care organizations. JAMA 1989; 262:2869 73. 2. Nease R, Kneeland T, OConnor G, et al. Variation in patient utilities for outcomes of the management of chronic stable angina: implications for clinical practice guidelines. JAMA 1995; 273:118590.

658

ECONOMICS AND HEALTH SYSTEMS RESEARCH MACARIO ET AL. PATIENT PREFERENCES FOR ANESTHESIA OUTCOMES

ANESTH ANALG 1999;89:6528

3. Schover L, Yetman R, Tuason L, et al. Partial mastectomy and breast reconstruction: a comparison of their effects on psychosocial adjustment, body image, and sexuality. Cancer 1995;75: 54 64. 4. Danis M, Mutran E, Garret J, et al. A prospective study of the impact of patient preferences on life sustaining treatment and hospital cost. Crit Care Med 1996;24:18117. 5. Donabedian A. Evaluating the quality of medical care. Millbank Mem Fund Q 1966;44(Part 2):166 206. 6. Shafer A, Fish P, Gregg K, et al. Preoperative anxiety and fear: a comparison of assessments by patients and anesthesia and surgery residents. Anesth Analg 1996;83:128591. 7. Hornberger J, Habraken H, Bloch D. Minimum data needed on patient preferences for accurate, efficient medical decision making. Med Care 1995;33:297310. 8. Davies A, Ware J. Involving consumers in quality of care assessment. Health Affairs 1988;15:339. 9. Chye EPY, Young IG, Osborne GA, et al. Outcomes after sameday oral surgery. J Oral Maxillofac Surg 1993;51:846 9. 10. King B. Patient satisfaction survey: day surgery unit. Aust Clin Rev 1989;9:1279. 11. Burrow B. The patients view of anaesthesia in an Australian teaching hospital. Anaesth Intensive Care 1982;10:20 4. 12. Philip B. Patients assessment of ambulatory anesthesia and surgery. J Clin Anesth 1992;4:355 8. 13. Keep P, Jenkins J. From the other end of the needle: the patients experience of routine anaesthesia. Anaesthesia 1978;33:830 2. 14. Osborne GA, Rudkin GE. Outcome after day-care surgery in a major teaching hospital. Anaesth Intensive Care 1993;21:8227. 15. Moerman N, van Dam F, Oostino J. Recollections of general anaesthesia: a survey of anesthesiological practice. Acta Anaesthesiol Scand 1992;36:76771.

16. Dodds CP, Harding MI, More D. Anaesthesia in an Australian private hospital: the consumers view. Anaesth Intensive Care 1985;13:3259. 17. Boyd N, Sutherland H, Heasman K, et al. Whose utilities for decision analysis? Med Decis Making 1990;10:1058 67. 18. Brown W, Hollander M. Statistics: a biomedical introduction. New York: J Wiley & Sons, 1977. 19. Deleted in proof. 20. Shevde K, Panagopoulos G. A survey of 800 patients knowledge, attitudes, and concerns regarding anesthesia. Anesth Analg 1991;73:190 8. 21. Donabedian A. The definition of quality and approaches to its measurement. Ann Arbor, MI: Health Adminsitration Press, 1980. 22. Fung D, Cohen M. Measuring patient satisfaction with anesthesia care: a review of current methodology. Anesth Analg 1998; 87:1089 98. 23. Tarazi E, Philip B. Friendliness of OR staff is top determinant of patient satisfaction with outpatient surgery. Am J Anesthesiol 1998;4:154 7. 24. Llewellyn-Thomas H, Sutherland H, Tibshirani A, et al. Describing health states: methodologic issues in obtaining values for health states. Med Care 1984;22:54352. 25. McNeil B, Pauker S, Sox H, Tversky A. On the elicitation of preferences for alternative therapies. N Engl J Med 1982;306: 1259 62. 26. Klafta J, Roizen M. Current understanding of patients attitudes toward and preparation for anesthesia: a review. Anesth Analg 1996;83:1314 21.

World J. Surg. 24, 690 695, 2000 DOI: 10.1007/s002689910111

WORLD
Journal of

SURGERY
2000 by the Socie te Internationale de Chirurgie

Manipulation of the Metabolic Response in Clinical Practice


Henrik Kehlet, M.D., Ph.D.
Department of Surgical Gastroenterology 435, Hvidovre University Hospital, DK-2650 Hvidovre, Denmark Abstract. Surgical injury is followed by profound changes in endocrine metabolic function and various host defense mechanisms leading to catabolism, immunosuppression, ileus, impaired pulmonary function, and hypoxemia. These physiologic changes are supposed to be involved in the pathogenesis of postoperative morbidity. Effective afferent neural blockade with continuous epidural local anesthetic techniques inhibits a major part of the endocrine metabolic response, leading to improved protein economy but without important effects on inammatory or immunologic responses. In contrast, pain treatment with other modalities such as nonsteroidal antiinammatory drugs (NSAIDs) and opioids has only a small inhibitory effect on endocrine metabolic responses. Preoperative high-dose glucocorticoid therapy provides additional pain relief and improves pulmonary function, but it reduces the inammatory response (acute-phase proteins, cytokines, hyperthermia) and immune function. Minimally invasive surgery leaves the endocrine metabolic responses largely unaltered but reduces the inammatory response and immune suppression. Thus several techniques are available to modify the stress responses in elective surgery patients. The effect of these techniques to alter endocrine metabolic and inammatory responses during severe surgical illness has not been established. Neural blockade and minimally invasive surgery have improved outcome following elective surgery, especially when integrated into a multimodal postoperative rehabilitation program. Application of this knowledge from pathophysiologic responses to uncomplicated surgical injury should be explored in patients with severe surgical illness.

on the effect of pain-relieving techniques, the use of high-dose glucocorticoid, and the use of minimally invasive surgery. Effect of Pain Relief on Surgical Metabolism Pain is induced by activation of the peripheral and central nervous systems, which also are involved as one of the major release mechanisms of the metabolic response to surgical injury [1, 3]. Pain relief is a necessary but not sufcient technique to improve surgical outcome [1], and pronounced differences exist between the metabolic effects of the various pain-relieving techniques. Nonsteroidal Antiinammatory Drugs (NSAIDs) NSAIDs are used routinely for acute pain treatment. Although they have been demonstrated to attenuate the endocrine metabolic response to endotoxin administration in human volunteers, most surgical studies have shown NSAIDs to have no or only a slight inhibitory effect on classic catabolic stress hormones, acutephase protein responses, and protein economy [3]. However, a few studies have shown that ibuprofen reduces adrenocorticotropic hormone (ACTH) and cortisol release as well as interleukin-6 (IL-6) after laparotomy [4] and that it decreases fever, tachycardia, and oxygen consumption in sepsis patients [5]. It has no effect on morbidity or survival. In summary, the well established analgesic effects of NSAIDs may have only a slight or inconsistent inhibitory effect on metabolic responses in surgical patients and may therefore serve only as one of the components in a multimodal effort to reduce catabolism. Opioids Opioids administered systemically in low dosages in patient-controlled analgesia or in an intermittent conventional regimen have only slight or no stress-reducing effects [1, 3, 6]. In contrast, high-dose opioid anesthesia may reduce intraoperative, but not postoperative, endocrine metabolic changes [3]. Epidural analgesia with opioids has also only a slight inhibitory effect on catecholamine and cortisol responses to surgery, although the results are not consistent in the literature [3]. In general, the effects are relatively small, especially during major operations. Accordingly, the effect of opioid administration on

Major surgery is still associated with undesirable sequelae such as pain, cardiopulmonary, infectious and thromboembolic complications, cerebral dysfunction, nausea, gastrointestinal paralysis, fatigue, and prolonged convalescence. It has been hypothesized that the surgical stress response and resulting increase in demands on organ functions may be responsible for these sequelae, and that a multimodal approach to controlling postoperative pathophysiology, metabolism, and rehabilitation subsequently may improve surgical outcome [1]. Furthermore, inhibition of the initial response to the surgical stimulus may theoretically be advantageous should a complication occur, based on the second-hit theory of multiple organ failure [2]. This article reviews established techniques that manipulate the metabolic response in elective surgical patients, and it is suggested that such changes when combined with those of other techniques may also be useful in severe surgical illness. The chapter focuses
Correspondence to: H. Kehlet, M.D., Ph.D., e-mail: henrik.kehlet@ hh.hosp.dk

Kehlet: Manipulation of Metabolism Table 1. Effect of epidural and spinal anesthesia and analgesia on postoperative nitrogen economy. Anesthesia Lumbar epidural local anesthetic Surgery Hysterectomy Colonic Hip surgery Hip surgery Thoracic epidural local anesthetic Colonic Gastric Comment 24-Hour block with inhibition of corticol and glucose response and improvement in nitrogen balance 44-Hour block. Postoperative nitrogen balance improved and 3methylhistidine excretion reduced No effect of single-dose epidural bupivacaine on urinary nitrogen and 3-methylhistidine excretion 24-Hour block with reduction of cortisol and glucose response as well as usual 72-hour postoperative shifts in amino acid composition in skeletal muscle 24-Hour block with reduction of urinary excretion of catecholamines but not cortisol; urinary nitrogen excretion and whole body protein turnover (leucine oxidation) reduced 48-Hour block with reduced plasma cortisol and glucagon response and decreased urinary catecholamine excretion; urinary nitrogen excretion reduced compared to pain relief with systemic opioids and epidural opioids 24-Hour block with reduced plasma cortisol and urinary catecholamine excretion; no effect on urinary nitrogen and cortisol excretion (n 2 5) Single-dose postoperative block (6 hours duration) with slightly reduced plasma catecholamines, glucagon, and cortisol levels; isotope study with decrease in glucose and urea turnover rates 24-Hour block with no inuence on plasma glucose, free fatty acids, and lactate except at the end of operation; and no effects on postoperative nitrogen urea or 3-methylhistidine excretion Single-dose epidural analgesia; no effect on plasma cortisol, glucose prolactin, or nitrogen balance 24-Hour block; no effects on plasma cortisol, glucose, or prolactin; improved nitrogen balance 48-Hour continuous epidural block reduced postoperative protein breakdown more effectively than a 24-hour block 48-Hour continuous epidural block improved postoperative protein synthesis rate compared with general anesthesia and systemic opioids 48-Hour treatment (epidural meperidine); no effect on nitrogen and 3-methylhistidine excretion Intraoperative epidural local anesthetic 72-hour postoperative epidural morphine; reduced plasma cortisol and glucagon and urinary excretion of catecholamines and nitrogen 24-Hour intermittent local anesthesia and 72-hour epidural morphine with insignicant reduction in urinary catecholamines and cortisol; unchanged 4-day nitrogen excretion 24-Hour block with improved glucose homeostasis and insignicant reduction in urinary nitrogen excretion Author Brandt [3]a

691

Vedrinne [3]a Carli [3]a Christensen [3]a Carli [3]a Tsuji [3]a

Aortic Abdominal Abdominal and thoracic Abdominal Abdominal Colonic Colonic Lumbar epidural opioid Colonic Gastric Abdominal Hysterectomy
a

Smeets [3]a Shaw [3]a Seeling [3]a De Lalande [3]a De Lalande [3]a Carli [7] Carli [8] Vedrinne [3]a Tsuji [3]a Hjorts [3]a Licker [3]a

These studies were referred to in a recent review [3].

protein economy is minor and inferior to that obtained by epidural local anesthetics [3] (Table 1). Neural Blockade with Local Anesthetics Because activation of the peripheral and central nervous systems plays a key role in initiating the hormonal and metabolic responses to surgical injury [3], an afferent neural blockade with local anesthetics may profoundly alter these responses. For lower body procedures, where epidural local anesthetics provide an effective afferent blockade [3], these techniques result in inhibition of the classic endocrine metabolic response parameters, such as catecholamines, cortisol, and glucose (Table 2). Subsequently, continuous epidural analgesia with local anesthetics improves postoperative nitrogen economy [3] (Table 1). Similarly, fat metabolism with increased lipolysis is reduced as are the usual increases in lactate and ketones [3]. Of major clinical importance is

the fact that a single-dose block has no important prolonged effects on metabolism or protein economy, and a 24-hour block improves nitrogen economy, with further improvement by a 48hour block (Table 1) [3]. Although neural blockade with local anesthetics modies several of the classic hormonal responses that may inuence renal function (cortisol, catecholamines, aldosterone, renin, antidiuretic hormone), no clinically important effects on postoperative uid or electrolyte balance have been demonstrated, except for reduced potassium excretion, which parallels the reduced catabolism [3]. Oxygen consumption is reduced in accordance with the sympathetic block and reduced catabolism. No studies are available on the effect of short- or long-term epidural analgesia on the endocrine metabolic response during critical surgical illness, where the characteristics and physiologic effects of hormonal changes are different from the initial acute response [9]. In contrast to the pronounced inhibition of endocrine metabolic

692

World J. Surg. Vol. 24, No. 6, June 2000

Table 2. Effect of neural blockade with local anesthetics on endocrine-metabolic and inammatory responses to elective surgery. Type of response Pituitary Inhibition or improvement No important effect T3 and T4 Calcitonin gene-related peptide Coagulation and brinolysis Acute-phase protein and interleukin-6 Water and sodium balance Granulocytosis and neutrophil function Liver enzymes and antipyrin clearance Hyperthermia No data Gastrointestinal peptides Testosterone Estradiol Somatomedin Ca2, Mg2, Zn, phosphate balance Macrophage-derived peptides (interleukins, tumor necrosis factor)

Adrenal, renal, and nervous systems

Metabolic

Immunologic

-Lipotrophin Adrenocorticotropin -Endorphin Growth hormone Arginine vasopressin Thyroid-stimulating hormone Luteinizing hormone and folliclestimulating hormone Prolactin Cortisol Aldosterone Renin Epinephrine Norepinephrine Hyperglycemia and glucose tolerance Insulin resistance Lipolysis Muscle amino acids Nitrogen balance Hepatic urea production Oxygen consumption Urinary potassium excretion Complement activation (C3a, C5a) Lymphopenia Natural killer cell suppression

Modied from Kehlet [3]. T3: triiodothyronine; T4: thyroxine.

responses by neural blockade during lower body procedures, a smaller reduction of catabolism and improvement in protein economy is achieved during upper abdominal/thoracic procedures [3] (Table 1). The explanation for this discrepancy between lower and upper body procedures is a less effective afferent blockade with thoracic epidural local anesthetic [3] and unblocked phrenic afferents during sub- or supradiaphragmatic surgery [10]. Another factor may be the unblocked vagal afferents during thoracic epidural analgesia, as experimental studies have demonstrated that vagotomy blocks the hyperthermic response to intraperitoneal IL-1 stimulation [11]. The limited clinical data in humans with intraoperative vagal blockade do not suggest that this nerve is important for the cortisol and hyperglycemic response to abdominal surgery [3]. Nevertheless, continuous thoracic epidural local anesthetic administration improves postoperative catabolism and protein economy, and an advantageous effect is achieved in regard to postoperative ileus [1, 3]. Thus epidural local anesthetics may further improve the anabolic catabolic balance by allowing early oral nutrition [1]. Epidural local anesthetics are equally effective for reducing the endocrine-metabolic response during normothermic or hypothermic recovery [12], despite the fact that hypothermia usully amplies the immediate postrecovery stress response. In contrast to the effects of a neural block on hormonal and metabolic responses, no important effects have been demonstrated on inammatory responses, that is, postoperative changes in acute-phase protein and cytokines (IL-6) [3]. These responses are thus independent of the neural activation and pain per se. Accordingly, the inammatory response is mostly independent on endocrine activation, although some studies have shown that etomidate (which selectively inhibits adrenocorticol responses) administration results in higher postoperative IL-6 values [13]; this nding suggests that endogenous glucocorticoids may modulate the IL-6 response to surgery.

General Anesthesia and Sedation The type of general anesthesia has not been demonstrated to have any long-term clinical relevant metabolic effects into the postoperative period, and only quantitatively minor differences exist among anesthetic agents in terms of the intra- and early postoperative endocrine metabolic responses [3]. The main effort to improve outcome for most types of surgery includes a multimodal approach with early rehabilitation [1]; therefore the effects of sedation techniques on metabolism are not relevant except in special circumstances. Currently, fast-track early recovery programs are under development for cardiac and other procedures but so far without investigating the endocrine-metabolic response. Continuous sedation with proprofol may reduce urinary and plasma catecholamine and cortisol responses, but only for a 12hour period of sedation [14]. In another study with more aggressive sedation using high-dose opioids for 24 hours during neonatal cardiac surgery, endocrine metabolic stress responses were reduced, and the outcome was improved [15]. Such a regimen requires extended ventilatory support, however, and has not been evaluated in other procedures. In summary, long-term sedation and its metabolic effects have been investigated only sporadically in elective surgical patients. Its role in modulating endocrine responses and improving catabolism and outcome is questionable. Glucocorticoids Evidence has accumulated that an overaggressive, protracted inammatory and hormonal host defense (stress) response may be an important factor in impairing the outcome after severe surgical illness. These data, together with the second-hit hypothesis [2], support the theory that modulating the initial endocrine metabolic

Kehlet: Manipulation of Metabolism

693 Table 3. Effect of laparoscopic versus open surgery on endocrinemetabolic and inammatory-immunologic responses. Endocrine-metabolic responses Cortisol 3 ACTH 3 Catecholamines 3 GH, prolactin 3 Glucagon 3 Insulin sensitivity 1 Nitrogen balance 31 Inammatory and immunologic responses CRP 2 IL-6 2 Leukocytosis 2 Neutrophil elastase 23 PHA response 1 Monocyte HLA-DR expression 23 Pulmonary function PaO2 or SpO2 1 Pulmonary function (FVC, FEV1) 1 Data are from published reviews [2729]. 1: increased response/function in laparoscopy group; 3: no difference between laparoscopy and open surgery; 2: decreased response/ function in laparoscopy group; ACTH: adrenocorticotropic hormone; GH: growth hormone; CRP: C-reactive protein; IL-6: interleukin-6; PHA: phytohemagglutinin; HLA-DR: class II human leukocyte antigen; FVC: forced vital capacity; FEV1: forced expiratory volume in 1 second; PaO2: partial oxygen pressure in arterial blood; SpO2: pulse oximetry.

and inammatory responses to the elective surgical procedure may reduce the risk following a second injury or complication. Although our understanding of the antiinammatory effects of glucocorticoids has improved [16], the clinical use of glucocorticoids for elective or severe surgical illness remains to be established. The use of glucocorticoids has been controversial, especially in severe surgical illness but with a recent refocusing on the potential positive effects of glucocorticoid therapy [17, 18]. The effects of preinjury glucocorticoid administration on various inammatory responses are relatively well established. Thus preendotoxin glucocorticoid may attenuate symptoms as well as IL-6, C-reactive protein (CRP), tumor necrosis factor (TNF) and IL-8 responses, but not the IL-1 receptor antagonist response [19, 20]. Similarly, high-dose preoperative methylprednisolone (30 mg/ kg) may attenuate catecholamine and arginine vasopression responses; the IL-6, CRP, and prostaglandin E2 (PGE2) responses; and activation of the plasma cascade system [2124]. The hyperthermic response is also inhibited [21, 22], and pain and fatigue are reduced [21]. In contrast, IL-10 responses are increased following methylprednisolone [23], and pulmonary function is improved following glucocorticoids in colonic surgery patients [21, 22]. The postoperative delayed hypersensitivity immunologic response is further reduced, leading to anergy [22]; but collagen accumulation in wounds is not altered [22]. The clinical consequences of a single preoperative dose of methylprednisolone has not been established in small-scale randomized studies, although one of the studies suggests less morbidity and a shorter intensive care unit (ICU) stay [24]. In relatively small-scale procedures the anti-inammatory effects of a preoperative glucocorticoid administration may be benecial because it reduces pain and swelling after dental procedures [25]. Also in patients undergoing arthroscopic meniscectomy, combined intraarticular glucocorticoid, bupivacaine, and morphine have a pronounced advantageous effect on postoperative pain, inammatory response (joint swelling, acute-phase protein), and convalescence (mobilization and sick leave) [26]. In summary, further data are needed for specic procedures on the short- and long-term potential (or risk) of modulation of the inammatory responses by preinjury glucocorticoid administration. So far, patients with preoperative severely impaired pulmonary function represent a group for whom glucocorticoids improve organ (pulmonary) function. Minimally Invasive Surgery Classic open operations are, as mentioned above, followed by profound changes in endocrine metabolic function and host defense mechanisms, thereby increasing the risk of subsequent organ dysfunction. Following the introduction of various minimally invasive surgery techniques to reduce wound size, a large body of data has shown that laparoscopic surgery leads to a reduced inammatory response (predominantly cytokine and acute-phase proteins), reduced immunomodulatory response, improved pulmonary function, and less hypoxemia and pain [2729]. In contrast, relatively little effect has been found in the classic endocrine metabolic responses for laparoscopic versus open operations [27 29]. Most studies have been performed for cholecystectomy or herniorrhaphy, however, and the few studies on major colonic operations suggest a smaller catabolic hormonal response with subsequent improvement in protein economy [27, 29]. The overall

effects on endocrine metabolic and inammatory responses during minimally invasive surgery are listed in Table 3. The clinical implications of the modied inammatory response during minimally invasive surgery have been considered obvious because of reduced pain, organ dysfunction, need for hospital stay, and convalescence. These ndings have also questioned the role of the endocrine metabolic responses in the determination of outcome; studies therefore have focused on the inammatory response [2729]. Although these data clearly suggest that altering the initial injury response may be benecial for minimally invasive surgery, the amount of tissue injury is reduced, not only the pathophysiologic response. The nal therapeutic advantage of the more costly minimally invasive surgical technique compared to open operation remains to be established, as treatment bias may have been introduced in the laparoscopic studies with insufcient blinding, and the open groups in most studies have not been appropriately treated with optimal pain relief, early oral feeding, and mobilization [1, 29]. However, preliminary observations for colonic surgery support that combined use of neural blockade and minimally invasive surgery may hasten recovery and reduce morbidity [1]. In conclusion, data on the effect of minimally invasive surgery on the endocrine metabolic and inammatory response suggest that reduction of the wound size (and thereby the inammatory response) may be benecial on outcome and the risk of developing a second severe surgical illness. These data also suggest that a short-lasting endocrine-metabolic response may not be as important for determining outcome. The minimally invasive surgery data should stimulate the development of other interventional techniques to modify the inammatory responses and to apply them in surgical patients in whom minimally invasive surgery cannot be used.

694

World J. Surg. Vol. 24, No. 6, June 2000

Conclusions and Future Directions The catabolic endocrine metabolic and inammatory response may have important consequences for outcome in major elective surgical procedures due to the profound changes in body composition with loss of weight and muscle mass and resistance to infection. Other sequelae of the surgical stress response are pain, sleep disturbance, ileus, and pulmonary and cardiac dysfunction, all of which may contribute to convalescence and fatigue [1]. Recent advances in perioperative care have included optimization of pain relief, nutrition, uid management, anesthesia and surgical techniques, and antimicrobial prophylaxis. Despite the fact that many unimodal intervention studies have reported a positive effect on outcome, the pathogenesis of common postoperative sequelae is multifactorial. Subsequently, a rational strategy to demonstrate substantial reduction of perioperative morbidity and shortened convalescence most be multimodal to modify all pathogenic mechanisms leading to the conventional postoperative cascade to dependence [1]. Such efforts include intensied preoperative information, stress reduction with neural blockade or humoral mediator modication, and sufcient pain relief, allowing early mobilization and facilitating recovery of gastrointestinal function to allow early restoration of oral intake. Based on such multimodal intervention studies [1], it is hypothesized that modication of catabolism and sympathetic responses to surgery may be advantageous [1]. Inhibition of the endocrine metabolic response may reduce catabolic consequences on muscle mass and function. Furthermore, modication of these responses by a neural blockade has obvious secondary advantageous effects in reducing nausea, vomiting, and ileus, thereby facilitating oral intake, which otherwise has been shown to reduce postoperative morbidity and risk of infectious complications. The inammatory responses are not inuenced by optimal pain relief or reduction of catabolism by afferent neural blockade but are reduced by glucocorticoids and minimally invasive surgery. The clinical advantageous effects of minimally invasive surgical techniques are well established and easy to understand because they reduce the amount of wound trauma. In contrast, modication of inammatory responses by pharmacologic intervention with glucocorticoids or other antiinammatory/anticytokine agents is controversial, as a certain magnitude of response obviously is necessary for sufcient immune function and resistance to infection and wound healing, but an exaggerated inammatory response is undesirable and contributes to the development of multiple organ failure. Based on the data derived from elective surgical procedures, it is concluded that surgically induced neural reex responses and endocrine metabolic catabolic responses should be alleviated by effective pain-relieving techniques and neural blockade. Modication of other (inammatory/immunologic) responses remains to be evaluated and explored regarding the clinical signicance on outcome. Such knowledge should serve as a scientic basis for potential applications in patients with severe surgical illness, but so far this has not been sufciently evaluated to allow any recommendations for clinical practice. Re sume Lagression chirurgicale est suivie de changements profonds de la fonction endocrine me tabolique et des me canismes de de fense de lho te, responsables dun etat catabolique, dune

immunosuppression, dun ile us, dune fonction pulmonaire diminue e et dune hypoxe mie. Ces changements physiologiques ont probablement une part dans la pathogene `se de la morbidite postope ratoire. Lorsque fait de fac on effectif, le blocage neuronal affe rent par une technique dane sthe sie epidurale continue, inhibe en partie la re sponse me tabolique endocrine, ame liorant ainsi le pargne prote inique, mais sans effet notable sur les re sponses inammatoire ou immunitaires. En contraste, le traitement de la douleur par les AINS ou les opiace s a un effet inhibiteur minime sur les re sponses me taboliques endocrines. Une glycocorticothe rapie pre ope ratoire ` a haute dose soulage la douleur et ame liore la fonction pulmonaire, mais re duit la re ponse inammatoire (prote ines de la phase aigue , cytokines, hyperthermie) et la fonction immune. La chirurgie mini-invasive produit des re ponses me taboliques endocrines non modie es, mais re duit la re ponse inammatoire et limmunosuppression. Ainsi plusieurs techniques sont disponibles pour modier la re ponse au stress des patients en chirurgie elective. Leffet de ces techniques sur la modication des re ponses me taboliques et inammatoires endocrines dans lagression chirurgicale se ve `re nest pas etabli. Le blocage neuronal et la chirurgie mini-invasive ont certes ame liore le volution apre `s la chirurgie elective, surtout lorsquils sont inte gre s dans un programme de re habilitation postope ratoire multimodale. Lapplication de ces connaissances, allant depuis les re ponses physiopathologiques ` a la le sion chirurgicale non complique e, devrait etre explore e chez le patient ayant une maladie chirurgicale, se ve `re. Resumen El trauma quiru rgico produce profundos cambios en la funcio n metabo lica y en una variedad de mecanismos de defensa que resultan en catabolismo, inmunosupresio n, leo, alteracio n de la funcio n pulmonar e hipoxemia. Tales cambios f siolo gicos supuestamente esta n involucrados en la morbilidad postoperatoria. Un control efectivo del bloqueo neural aferente mediante anestesia epidural inhibe una parte mayor de la respuesta metabo lica endocrina con mejoramiento de la econom a proteica, aunque sin efecto importante sobre las respuestas inamatoria o inmunitaria. Por el contrario, el manejo del dolor con otros me todos tales como AINES u opia ceos apenas ejerce un efecto inhibitorio m nimo sobre las respuestas metabo licas endocrinas. La terapia preoperatoria con altas dosis de glucocorticoides provee control adicional del dolor y mejora la funcio n pulmonar, pero reduce la respuesta inamatoria (prote nas de fase aguda, citocinas, hipertermia) y la funcio n inmunitaria. La cirug a m nimamente invasora produce m nima alteracio n en la respuesta metabo lica endocrina, pero reduce la reaccio n inamatoria y el grado de inmunosupresio n. Se puede ver que existen diversos me todos para modicar las respuestas de estre s en los pacientes sometidos a cirug a electiva. El efecto de tales me todos, en cuanto a la alteracio n de las respuestas metabo licas endocrinas inamatorias en la enfermedad quiru rgica severa, au n no ha sido bien denido. El bloqueo neural y la cirug a m nimamente invasora han mejorado los resultados de la cirug a electiva, especialmente cuando se integran a un programa multinodal de rehabilitacio n postoperatoria. La aplicacio n de todo este conocimiento sobre las respuestas siopatolo gicas al trauma quiru rgico no complicado debe ser investigada en los pacientes con enfermedad quiru rgica grave.

Kehlet: Manipulation of Metabolism

695 G., Howse, J., Herskowitz, A., and Mangano, D.T.: Urine and plasma catecholamine and cortisol concentrations after myocardial revascularisation: modulation by continuous sedation. Anesthesiology 86:785, 1997 Anand, K.J.S., and Hickey, P.R.: Halothane-morphine compared with high-dose sufentanil for anesthesia and postoperative analgesia in neonatal cardiac surgery. N. Engl. J. Med. 326:1, 1992 Barnes, P.J.: Antiinammatory action of glucocorticoids: molecular mechanisms. Clin. Sci. 94:557, 1998 Meduri, G.U., and Kanangat, S.: Glucocorticoid treatment of sepsis and acute respiratory distress syndrome: time for a critical re-appraisal. Crit. Care Med. 26:630, 1998 Matot, I., and Sprung, C.L.: Corticosteroids in septic shock: resurrection of the last rites? Crit. Care Med. 26:627, 1998 Santos, A.A., Scheltinga, M.R., Lynch, E., Brown, E.F., Lawton, P., Chambers, E., Browing, J., Dinarello, C.A., Wolff, S.M., and Wilmore, D.W.: Elaboration of interleukin-1-receptor antagonist is not attenuated by glucocorticoids after endotoxemia. Arch. Surg. 128:138, 1993 Rock, C.S., Coyle, S.M., Keogh, C.V., Lazarus, D.D., Hawes, A.S., Leskiw, M., Moldawer, L.L., Stein, T.P., and Lowry, S.F.: Inuence of hypercortisolemia on the acute phase protein response to endotoxin in humans. Surgery 112:467, 1992 Schulze, S., Sommer, P., Bigler, D., Honnens, M., Shenkin, A., Cruickshank, A.N., Bukhave, K., and Kehlet, H.: Effect of combined prednisolone, epidural analgesia and indomethacin on the systemic response after colonic surgery. Arch. Surg. 127:325, 1992 Schulze, S., Andersen, J., Overgaard, H., Nrgaard, P., Nielsen, H.J., Aasen, A., Gottrup, F., and Kehlet, H.: Effect of prednisolone on the systemic response and wound healing after colonic surgery. Arch. Surg. 132:129, 1997 Tapardel, Y., Duchatteau, J., Schmartz, D., Marecaux, G., Shala, M., Barvais, L., le Clerc, J-L., and Vincent, J-L.: Corticosteroids increase blood interleukin-10 levels during cardiopulmonary bypass in men. Surgery 119:76, 1996 Taketa, S., Ogawa, W.R., Nakanishi, K., Kim, C., Miyashita, M., Sasajima, K., Unda, M., and Takano, T.: The effect of preoperative high-dose methylprednisolone in attenuating the metabolic response after oesophageal resection. Eur. J. Surg. 163:511, 1997 Skjelbred, P., and Lkken, P.: Reduction of pain and swelling by corticosteriod injected 3 hours after surgery. Eur. J. Clin. Pharmacol. 23:141, 1982 Rasmussen, S., Larsen, A.S., Thomsen, S.T., and Kehlet, H.: Intraarticular glucocorticoid, bupivacaine and morphine reduces pain, inammatory response and convalescence after arthroscopic meniscectomy. Pain 78:131, 1998 Kehlet, H., and Nielsen, H.J.: Impact of laparoscopic surgery on stress responses, immunofunction and risk of infectious complications. New Horiz. 6:s80, 1998 Vittimberga, F.J., Foley, D.P., Meyers, W.C., and Callery, M.P.: Laparoscopic surgery and the systemic immune response. Ann. Surg. 227: 326, 1998 Kehlet, H.: The surgical stress response: does endoscopic surgery confer an advantage? World J. Surg. 23:801, 1999

Acknowledgment This study was supported by a grant from the Danish Medical Research Council (28809). References
1. Kehlet, H.: Multimodal approach to control postoperative pathophysiology and rehabilitation. Br. J. Anaesth. 78:606, 1997 2. Garison, R.N., Spain, D.A., Wilson, M.A., Keelen, P.A., and Harris, P.D.: Microvascular changes explain the two-hit theory of multiple organ failure. Ann. Surg. 227:851, 1998 3. Kehlet, H.: Modication of responses to surgery by neural blockade: clinical implications. In Neural Blockade in Clinical Anesthesia and Management of Pain, Cousins, M.J., and Bridenbaugh, P.O., editors, Philadelphia, Lippincott-Raven, 1998, pp. 129 175 4. Chambrier, C., Shassard, D., Bienvenu, J., Saudin, F., Paturel, B., Garrique, C., Barbier, Y., and Bulletreau, P.: Cytokine and hormonal changes after cholecystectomy: effect of ibuprofen pretreatment. Ann. Surg. 224:178, 1996 5. Bernard, G.R., Wheeler, A.P., Russell, J.A., Shein, R., Summer, V.R., Steinberg, K.B., Fulkoson, W.J., Wright, P.E., Christman, B.W., Dupont, W.D., Heaggins, S.B., and Swindell, B.: The effects of ibuprofen on the physiology and survival of patients with sepsis. N. Engl. J. Med. 336:912, 1997 6. Mangano, D.T., Ciliciano, D., Hollenberg, M., Leum, J.M., Browner, W.S., Goehner, P., Merrick, S., and Verrier, E.: Postoperative myocardial ischemia; therapeutic trials using intensive analgesia following surgery. Anesthesiology 76:342, 1992 7. Carli, F., and Halliday, D.: Modulation of protein metabolism in the surgical patient: effects of 48-hours continuous epidural block with local anesthetics on leucine kinetics. Reg. Anesth. 21:430, 1996 8. Carli, F., Phil, M., and Halliday, D.: Continuous epidural blockade arrests the postoperative decrease in muscle protein fractional synthetic rate in surgical patients. Anesthesiology 86:1033, 1997 9. Van den Berghe, G., Dezegher, F., and Boullion, R.: Acute and prolonged critical illness as different neuroendocrine paradigms. J. Clin. Endocrinol. Metab. 83:1827, 1998 10. Segawa, H., Mori, K., Kasai, K., Fukata, J., and Nakao, K.: The role of the phrenic nerves in stress response in upper abdominal surgery. Anesth. Analg. 82:1215, 1996 11. Watkins, L.R., Goehler, L.E., Relton, J.K., Tartaglin, N., Silbert, L., Martin, D., and Maier, S.F.: Blockade of interleukin-1 induced hyperthermia by subdiaphragmatic vagotomy: evidence for vagal mediation of immune-brain communication. Neurosci. Lett. 183:27, 1995 12. Motamed, S., Klubien, K., Edwardes, M., Mazza, L., and Carli, F.: Metabolic changes during recovery in the normothermic versus hypothermic patients undergoing surgery and receiving general anesthesia and epidural local anesthetic agents. Anesthesiology 88:1211, 1998 13. Jameson, P., Desborough, J.P., Bryant, A.E., and Hall, G.M.: The effect of cortisol suppression on interleukin-6 and white blood cells responses to surgery. Acta Anaesthesiol. Scand. 41:304, 1997 14. Plunkett, J.J., Reeves, J.D., Ngo, L., Bellows, W., Shafer, S.L., Roach,

15. 16. 17. 18. 19.

20.

21.

22.

23.

24.

25. 26.

27. 28. 29.

Easing the Pain of JCAHO Accreditation Standards: Focus on Post-op Pain Gregory Holmquist, Oncology & Pain Management Pharmacist Specialist Palliative Care Strategies rxrelief@aol.com
( 1999 all rights reserved Palliative Care Strategies)

Pain is one of the most common reasons people consult a physician, yet it frequently is inadequately treated, leading to enormous social cost in the form of lost productivity, needless suffering, and excessive healthcare expenditures. Consensus statement from the American Academy of Pain Medicine and the American Pain Society, 1997. Successful assessment and control of pain depends, in part, on establishing a positive relationship between health care professionals and patients. Patients should be informed that pain relief is an important part of their health care, that information about options to control pain is available to them, and that they are welcome to discuss their concerns and preferences with the health care team. AHCPR Clinical Practice Guidelines - Acute Pain Management in Adults Unrelieved pain has negative physical and psychological consequences. Aggressive pain prevention and control that occurs before, during, and after surgery can yield both short- and long-term benefits. AHCPR Clinical Practice Guidelines - Acute Pain Management in Adults The reasons for inadequate treatment (of pain) are many. These include deficiencies in knowledge and skills on the parts of health care providers, patients and those responsible for the management of health care systems. International Association for the Study of Pain

Patients have a right to appropriate assessment and management of pain


JCAHO Pain Management Standards, 1999 "Doctors have the means at hand to relieve the suffering of millions of Americans. Why arent they doing it? Front Cover - U.S. News & World Report, March 17, 1997 PAIN is the enemy, not the patient. Greg Holmquist, Pain / Oncology Specialist

I. INTRODUCTION: Application of principles learned in the management of cancer pain to that of acute pain and chronic pain syndromes. ! Pain, regardless of the type and source, is a complex syndrome that encompasses both physical
manifestations and emotional responses. ! Proper assessment and follow-up of patients with pain are essential steps in providing the foundation for successful management. ! Many times myths and fears on the part of health care providers and patients can affect the appropriate prescribing of medications by practitioners, the administration of pain medications by nurses and the compliance with therapies by patients. !" Knowledge deficits regarding the pharmacology of opioids and pain adjuvant therapies can hinder the ability to obtain effective pain control in patients with both acute and chronic pain.

II.

A COMPARISON OF ACUTE AND CHRONIC PAIN

Physical Aspects / Impact of Pain on Life Location of pain Description of pain Duration of pain Therapy model Mood of patient Rate of depression Pain tolerance Impact of pain on family and relationships Impact of pain on job Impact of pain on patients function Response from physicians to pain syndrome Goal of work-up Role of analgesic therapy

Acute Pain Defined Easy Short Curative model Anxiety, fear Same as general population Usually OK Helpful and supportive Usually OK Usually improves as tissue heals Comforting, fix the problem

Chronic Non-malignant Pain Diffuse, spread out Difficult Ongoing, usually very long term Restorative model Depression, guilt, frustration, irritability, anger, hopelessness Three to four times the general population. Decreased, wears person out Tired, may deteriorate Questionable, jeopardy Dysfunction often becomes steady-state over time. Many times, blame, the pain is in your mind, add pills, less actual follow-up Move beyond endless and fruitless diagnostic work-up Improved function and life quality

Cancer Pain Can be both defined and diffuse Difficult Intermediate to long term Palliative model Anxiety, depression, fear, worry, hopelessness. Higher than general population Decreased, wears person out Supportive, but fearful. Depends on situation Can lead to significant dysfunction Comforting, helplessness, need for consultation with the experts Assist in decision making Pain relief, high quality of life, death with dignity

Clarity of somatic diagnosis Pain control while healing occurs

III. JCAHO Standards for Pain Management in Hospital Settings (excerpted from the official JCAHO Web site http://www.jcaho.org) A. Rights and Ethics Chapter (RI) Standard RI.1.2: Patients are involved in all aspects of their care.

Hospitals promote patient and family involvement in all aspects of their care through implementation of policies and procedures that are compatible with the hospital's mission and resources, have diverse input, and guarantee communication across the organization. Examples of implementation RI.1.2: (Per JCAHO) #" A hospital includes a commitment to pain management in its mission statement, patient and family bill of rights, or service standards (for example, Patients have the right to expect a quick response to reports of pain"). #" The following statement(s) on pain management are posted in all patient care areas:
Patients Rights: As a patient at this hospital, you can expect information about pain and pain relief measures, a concerned staff committed to pain prevention and management, health professionals who respond quickly to reports of pain, your reports of pain will be believed, state-of-the-art pain management, and dedicated pain relief specialists. Patient Responsibilities As a patient at this hospital, we expect that you will ask your doctor or nurse what to expect regarding to pain and pain management, discuss pain relief options with your doctors and nurses, work with your doctor and nurse to develop a pain management plan, ask for pain relief when pain first begins, help your doctor and nurse assess your pain, tell your doctor or nurse if your pain is not relieved, and tell your doctor or nurse about any worries you have about taking pain medication.

Standard RI.1.2.8: Patients have the right to appropriate assessment and management of pain. Pain is a common part of the experience; unrelieved pain has adverse physical and patient psychological effects. The patients right to pain management is respected and supported. The health care organization plans, supports, and coordinates activities and resources to assure the pain of all patients is recognized and addressed appropriately. #" Initial assessment. #" Regular re-assessment. #" Education of relevant providers in pain assessment and management. #" Education of patients, and families when appropriate, regarding their roles in managing pain as well as the potential limitations and side effects of pain treatments. #" Communication to patients and families that pain management is an important part of care.

III.

JCAHO Standards for Pain Management in Hospital Settings (continued) (excerpted from the official JCAHO Web site http://www.jcaho.org) A. Rights and Ethics Chapter (RI) (continued) Examples of implementation RI.1.2: #" Pain is considered the "fifth" vital sign. Pain intensity ratings are recorded along with temperature, pulse, respiration, and blood pressure. #" Every patient is asked a "screening question regarding pain on admission. Patients and families receive information verbally and in an electronic or printed format at the time of initial evaluation that effective pain relief is an important part of their treatment. #" Competency in pain assessment and treatment is determined during the orientation of all new clinical staff. #" The hospital demonstrates its commitment to pain management by holding twice-annual staff awareness events regarding pain assessment and treatment. #" The hospital supplies educational materials about pain to all patients. For outpatient surgery patients, information is mailed to patients prior to the day of surgery. #" All telephone follow-up (for example, outpatient surgery, short stay obstetrics, evaluation of discharge planning) includes asking the patient about their pain status. #" The following statement on pain management is posted in all patient care areas (patient rooms, clinic rooms, waiting rooms, and so forth):

All patients have a right to pain relief.

B. Assessment of Patients Chapter (PE) Standard PE.1.4: Pain is assessed in all patients. In the initial assessment, the organization identifies patients with pain (or at risk for pain). The assessment and a measure of pain intensity are recorded in a way that facilitates regular reassessment and follow-up according to criteria developed by the organization.

III. JCAHO Standards for Pain Management in Hospital Settings (continued) (excerpted from the official JCAHO Web site http://www.jcaho.org) B. Assessment of Patients Chapter (continued) Examples of implementation PE. 1.4: #" All patients at admission are asked the following screening or general questions about the presence of pain: Do you have pain now? Have you had pain in the last several weeks or months? If the patient responds "yes" to either question, additional assessment data are obtained: $" Pain intensity $" Location $" Quality, patterns of radiation, if any, character $" Onset, duration, variation, and patterns. $" Alleviating and aggravating factors. $" Present pain management regimen and effectiveness. $" Pain management history $" Effects of pain $" The patient's pain goal $" Physical exam/observation of the pain site. #" Patients often have more than one site of pain. An assessment system or tools with space to record data on each site is provided on the assessment sheet. #" A hospital may need to use more than one pain intensity measure. For example, a hospital serving both children and adults selects a scale to be used with each of those patient populations. Assessment of cognitively impaired patients may also require assessment of behavioral factors signaling pain or discomfort. #" Staff are educated about pain assessment and treatment including the barriers to reporting pain and using analgesics. Staff encourage the reporting of pain when a patient and/or family member demonstrates reluctance to discuss pain, denies pain when pain is likely to be present (for example, post-operative, trauma, burns, cardiac emergencies), or does not follow through with recommended treatments. #" Pain intensity scales are enlarged and displayed in all areas where assessments are conducted. For organizations using clinical pathways, pain assessment is incorporated in some way into every appropriate clinical pathway. #" An organization selects pain intensity measures to ensure consistency across departments, for example, the 0-10 scale, the Wong-Baker FACES pain rating scale (smile-frown), and the verbal descriptor scale. Adult patients are encouraged to use the 0-10 scale. If they cannot understand or are unwilling to use it, the smile-frown or the verbal scale is used. 5

III. JCAHO Standards for Pain Management in Hospital Settings (continued) (excerpted from the official JCAHO Web site http://www.jcaho.org) C. Care of Patients Chapter Standard TX.3.3 Policies and procedures support safe medication prescription or ordering. #" PCA, Spinal administration of medications, and other pain management technologies utilized in the care of patients with pain. #" Distribution and administration of controlled substances. #" PRN and scheduled prescriptions or orders and times of dose administration. #" Distribution of medications to patients at discharge. Example of implementation of TX 3.3: (Per JCAHO) #" Before initiating patient-controlled analgesia (PCA) for surgical patients, an interdisciplinary team of physicians, pharmacists, and nurses:
Reviewed the literature on PCA. Drafted policies, procedures, and standing orders. Obtained approval from the pharmacy and therapeutics committee and medical staff. Oriented all staff. Conducted a pilot test on the general surgery patient care unit.

Standard TX.5.4 The patient is monitored during the post-procedure period. The patient is monitored continuously during the post-procedure period. The following items are monitored: #" Physiological and mental status. #" Impairments and functional status. #" Pain intensity, duration, location, character, and responses to treatment. Example of implementation of TX 5.4: (Per JCAHO) #" In a day surgery setting: Discharge criteria are set, including pain, that determines if a patient is ready for discharge to home. The day surgery center also contacts the patient the following day. One of the assessment parameters asked of the patient, depending on the surgical procedure, includes changes in pain intensity, relief from prescribed medications, and ability to rest. Based on the results of the assessment, follow up is initiated per the day surgery protocols.

III. JCAHO Standards for Pain Management in Hospital Settings (continued) (excerpted from the official JCAHO Web site http://www.jcaho.org) D. Education Chapter (PF) Standard PF.1.7 Patients are taught that pain management is a part of treatment. Hospitals offer education to patients and families to give them the specific knowledge and skills they need to meet the patient's ongoing health care needs. Clearly, such instruction needs to be presented in ways that are understandable to those receiving them. The hospital uses guidelines in educating patients on the following topics: Safe and effective use of medication. Understanding pain and the importance of effective pain management. Rehabilitation. Educational resources in the community. Follow-up care. Examples of implementation of PF.1.7: (Per JCAHO) #" Recognizing the impact uncontrolled pain has on a patient's functional status, a hospital includes information about pain management to all of its patients at discharge. This information includes: General information about pain. Use of medications, if indicated or prescribed. The use of non-pharmacological interventions, including heat, cold, exercise, and physical therapy. Specific directions on when to call a health care professional for additional assistance. #" The pharmacy department reviews its computer-generated individual education information on pain medication and realizes the material on opioids does not reflect a balanced and accurate reflection of the incidence and severity of possible side effects and cautions for use. A committee is formed to revise these materials. The pharmacy then shares the revisions with the outpatient pharmacies in its system and forwards them to the software developer to include in the next revision of the individual education materials.

E. Improving Organization Performance Chapter Standard PI The organization collects data to monitor its performance.
Data that the organization considers for collection to monitor performance include the appropriateness and effectiveness of pain management.

IV. REASONS FOR NOT PROVIDING EFFECTIVE POST-OPERATIVE PAIN CONTROL Exaggerated fears and misunderstanding of addiction. Failure to appreciate the negative consequences of uncontrolled pain. Underestimation of the effective dosage range and overestimation of the duration of action of common pain medications. Lag time between pain perception and provision of pain relief. Lack of dose individualization. Opioid dose variation #" Differences in pain intensity #" Pharmacokinetic differences: absorption, first-pass effect, half-life #" Use or non-use of co-analgesics #" Differences in pain threshold #" Previous use or non-use of opioids #" Previous inadequate use of analgesics (fear-factor) #" Tolerance #" Patient anxiety Over-reliance on PRN dosing regimens of short-acting pain medications.

V. KEY STRATEGIES FOR CONTROLLING ACUTE PAIN Assess pain routinely Treat pain as early as possible (pre-emptive analgesia) Use drug and non-drug interventions Select treatments according to the clinical setting and promptly modify according to the patients response. Provide continuity of pain control after discharge.

VI. ASSESSMENT AND FOLLOW-UP


Key understandings: #"Wide variations from patient to patient in the amount of pain that is experienced in response to a particular insult. #"Wide variations in responses to particular therapeutic strategies.
Genetic differences Past experiences Levels of anxiety, fear, meaning of pain, ethnocultural background, sense or lack of control.

Include both patient (and family when possible) in the assessment process.
#"Assume poor factual knowledge base of patient and family and a preponderance of myths, partial truths, fears leading to anxiety, confusion. #"Ask about pain regularly believe the reports of pain.

Assess pain systematically:


#"Self-reports #"Where does it hurt #"Severity / intensity

primary source of assessment DO YOU HAVE PAIN NOW? identify ALL sites (HINT: usually more than one site) Use reliable instruments that will help gauge the success or failure of a particular treatment plan. Acute pain

Areas to measure:
Level of discomfort/comfort (e.g. 0 to 10 scales, verbal word scales, faces scales, etc.) Effect of pain on ambulation Effect of pain on sleep Effect of pain on diet Effect of pain on urine/bowel habits Effect of pain on function: %"Ability to participate in PT/rehab %"Ability to communicate/socialize %"Ability to exercise %"Ability to return to work

+++ ++++ ++++ ++ +++ ++++ ++ ++ ++++

++++ extremely important area to measure; +++ very important area to measure ++ may or may not be useful to measure for particular type of pain; + probably not necessary to measure

#"Description of pain (quality) Have patient describe pain in their own words. Dull, aching versus sharp / stabbing Burning versus shooting, electric shock Constant vs. intermittent #"Onset, duration, variation, and patterns. #"Aggravating/relieving factors Have patient state what makes the pain worse or better

#"Previous therapy experiences Have patient tell you what has been tried before (successfully / unsuccessfully) Psychosocial issues
#" Anxiety/depression history #" Patient concerns regarding controlled substances use #" Patient knowledge, preferences, expectations, the patient's pain goal "Health professionals should ask about pain, and the patient's self-report should be the primary source of assessment." AHCPR Clinical Practice Guidelines - Chronic Cancer Pain

VII. FEARS AND MISCONCEPTIONS WITH THE USE OF OPIOIDS Physical dependence versus Addiction Physical dependence is a physiological phenomenon defined by the development of an abstinence syndrome following abrupt discontinuation of therapy, substantial dose reduction, or administration of an antagonist drug. Addiction is compulsive use of a substance resulting in physical, psychological or social harm to the user and continued use despite that harm.
Addiction is characterized by a core of aberrant drug-related behaviors including: 1. Loss of control of drug-use. 2. Compulsive drug-use. 3. Continued use despite harm.

Tolerance - has not proven to be a prevalent limitation to long-term opioid use. -Rapid escalation of drug dose in cancer pain usually due to disease progression. Respiratory depression Withholding the appropriate dose of opioids from a patient who is experiencing pain on the basis of respiratory concerns is unwarranted. Patients most at risk: Opioid naive Very elderly Rapid infusion of high-doses parenteral Underlying respiratory diseases (pneumonia, COPD) Use of agents that can accumulate (methadone) Diversion - Attention to patterns of prescription requests and the prescribing of opioids as part of an ongoing relationship between a patient and a health care provider can decrease the risk of diversion. Utilize agents with a lower potential for diversion (e.g. long-acting). Review by medical boards - necessary to document appropriate use of opioids: Clear goals of therapy Functional improvement of the patient Level of comfort experienced by the patient Adverse reactions, side effect management Informed patient consent

The greatest enemy of the truth is very often not the lie - deliberate, contrived and dishonest - but the myth - persistent, persuasive and unrealistic. John F. Kennedy 10

VIII. OVERALL STRATEGIES FOR THE USE OF MEDICATIONS IN THE PREVENTION AND TREATMENT OF PAIN Acute Pain
Goals of analgesics #"Minimize discomfort #"Facilitate recovery #"Avoid treatment related side-effects #"Simple (acetaminophen, NSAIDs) #"Opioids (short and long-acting) #"Adjuvants play a lesser role #"Post-op pain PCA preferred (IV / oral) Initially use parenteral route, then oral route when able to take po. IM falling out of favor Spinal route for some patients #"Initially, regularly scheduled, then tailored to pain / situation. #"Individualize therapy (doses / medication selection) to patients needs. #"Post-op: Loading dose to achieve steady state level of opioid quickly Have patient self-titrate PCA to desired level of pain control. #"Pre-emptive analgesia may play an important role. #"Increased aggressiveness in managing post-op pain. (use of strong opioids, higher doses, longer treatment time periods). #"Use of oral long-acting opioid agents to provide PCA in hospital and after discharge.

Chronic Non-malignant Pain


#"Improve function and life quality #"Avoid treatment related side-effects #"Simple (acetaminophen, NSAIDs) #"Opioids (long-acting) #"Adjuvants play a large role #"Oral #"Spinal, rarely

Cancer Pain
#"Palliate present pain quickly and prevent future pain. #"High quality of life #"Avoid treatment related side-effects #"Simple (acetaminophen, NSAIDs) #"Opioids (long-acting) #"Adjuvants play a large role #"Oral as first-line #"Transdermal, rectal as second-line #"Parenteral (Continuous SQ, PCA) #"Spinal, rarely

.. Types of analgesics utilized .. Routes utilized for opioids

.. Dosing / scheduling

#"Regularly scheduled #"Individualize therapy (doses / medication selection) to patients needs. #"Very small quantity of break-through / rescue medications #"Titration period over weeks

#"Regularly scheduled #"Individualize therapy (doses / medication selection) to patients needs. #"Rapid titration when pain out of control. #"Always provide patient with supply of short-acting morphine or oxycodone for new pains or pains that breakthrough. #"Unacceptable to deny the use of strong opioids or wait until death is imminent before using strong opioids for the palliation of pain. #"Long-acting oxycodone preferred over long-acting morphine due to equal efficacy, lower incidence of side effects and greater patient acceptability.

.. Titration

.. NEWER STRATEGIES

#"Prevent acute pain from becoming chronic pain. #"Focus on functional improvements #"Interdisciplinary approach #"Use of opioids (specifically, longacting agents) for the ongoing management of pain. #"Use of treatment contracts, patient diaries

11

IX. POST-OPERATIVE STUDIES UTILIZING LONG-ACTING ORAL OPIOIDS A. Use of oral opioids for patient-controlled analgesia (PCA)
Ginsberg B et al Conversion from IV PCA morphine to oral controlled-release oxycodone tablets for post-operative pain management. IASP Meeting Presentation. Orlando, FL 1998

#"Open label, multi-center study, 189 patients #"Elective surgery abdominal (41%), orthopedic (32%), gynecologic (26%), urological (1%) #"Post-surgery patients on IV PCA (84% Morphine) for 12-24 hrs, able to tolerate oral, no evidence of ileus or respiratory depression. #"Conversion to long-acting controlled-release (CR) oxycodone Based on previous 24 hour PCA Morphine IV dose IV PCA discontinued at the start of oral long-acting controlled-release oxycodone Immediate-release oxycodone (5mg) available as rescue Stopped oral long-acting CR oxycodone when pain < 4 for 24 hours Patients followed for up to seven days as outpatients #"Assessed: Pain intensity, sleep, side effects and perceived acceptability of therapy #"Results: Efficacy: Oral CR oxycodone provided significantly better pain control than IV PCA Dosing: Mean dose of IV PCA MS ranged from 40 mg to 60 mg Mean dose of oral CR oxycodone ranged from 42 mg to 62 mg #"Conclusions: Safe and effective to place patients within the first 24 hours post-op on oral long-acting CR oxycodone for pain relief. Use a factor of 1.0 to 1.3 when converting from IV PCA MS to oral CR oxycodone Treat constipation

B. Use of oral opioids for post-operative pain relief following ACL surgery
Postoperative Analgesia with Controlled-Release Oxycodone for Outpatient ACL Surgery Reuben SS, Connelly NR, Maciolek H. Anesth Analg 1999;88:1286-91

#"Open label, randomization to one of three oxycodone arms Group 1: (PRN group): Oxycodone 10 mg (short-acting) Q 4 H prn Group 2: (fixed-dose short-acting group): Oxycodone 10 mg (short-acting) Q 4 H on a fixed schedule starting 1 hour before d/c Group 3: (Controlled-release group): Oxycodone 20 mg CR (OxyContin) Q 12 H on a fixed schedule starting 1 hour before d/c #"Post-operative patients recovering from ACL surgery #"Assessed: Pain, sedation, sleep, satisfaction, side effects, amount of medication used #"Results: Significantly lower pain scores in Group 3 (controlled-release long-acting oxycodone) Significantly less sedation, greater satisfaction, less side effects and less drug utilized in Group 3 (controlled-release long-acting oxycodone)

12

X. PHARMACOLOGICAL DECISIONS A. Short-acting opioids (e.g. percocet, tylox, vicodin, loratab demerol, dilaudid, etc.) versus long-acting oral opioids (e.g. OxyContin, MS Contin, Oramorph SR, Kadian)
!"Short-acting opioids have a greater incidence of diversion.- due to quicker onset, enhanced euphoriant effects, ability to dissolve in solution and administer parenterally. !"Short-acting opiates reinforce the cycle of discomfort and dysfunction due to rapid onset and rapid loss of action. Uncontrolled pain usually leads to more uncontrolled pain with resultant need for higher doses of medication to control the pain. !"Short-acting opioids have a much greater fluctuation in blood levels leading to euphoria and sedation when the blood levels peak, and to pain escalation when the blood levels bottom out. Short acting opioids require that patients must take self-administer the medication every two to four hours in order !" to obtain around-the-clock pain relief. This is not only cumbersome, but can result in poor compliance, interrupted sleep patterns for patients who must awaken during the night to take a scheduled routine dose, and contribute to a decreased quality of life for the chronic pain patient.

How the use of short-acting opioids (e.g. vicodin, loratabs, percocet, dilaudid, etc.) can lead to poor outcomes for patients with chronic pain.
Sedation, euphoria, dysphoria Sedation, euphoria, dysphoria

Opioid Blood Levels

Pain relief Pain returns

Pain relief Pain returns

Pain relief Pain returns

Pain relief Pain returns

4 hrs

8 hrs

12 hrs

16 hrs

How the use of long-acting oral opioids can maximize the potential of a positive outcome for patients with chronic pain.
Sedation, euphoria, and / or dysphoria are minimized

Opioid Blood Levels

Pain relief Return of pain occurs less frequently

Pain relief

4 hrs

8 hrs

12 hrs

16 hrs
13

X. PHARMACOLOGICAL DECISIONS (continued) B. Drug Selection ***DO NOT USE THESE AGENTS UNLESS THERE IS NO OTHER OPTION***

Propoxyphene (Darvon)- lack of proven activity (beyond placebo effect), physical


dependency, risk of adverse reactions.

Pentazocine (Talwin) low level of activity, increased incidence of hallucinations,

delirium, and agitation, ability to induce withdrawal reactions in patients stabilized on opioid therapy.

Meperidine (Demerol) - very short duration of action, metabolite causes delirium

and seizures, oral tablets minimally effective unless LARGE doses utilized.

MYTHS of MEPERIDINE: #"Only patients with renal dysfunction can get seizures from meperidine. #"Other than meperidine, there are no good alternatives to use for patients who cannot tolerate morphine. #"Meperidine has a decreased incidence of respiratory depression. #"For some patients, meperidine is the only analgesic that will work for their pain.

Be aware of the potential hazards of meperidine (Demerol) and mixed agonist-antagonists, particularly pentazocine (Talwin). .meperidine should not be used for more than 48 hours for acute pain in patients WITHOUT renal or CNS disease, or at doses greater than 600 mg / 24 hours, and should not be prescribed for chronic pain. American Pain Society, Principles of Analgesic Use in the Treatment of Acute Pain and Cancer Pain 1999.

When propoxyphene (Darvon), meperidine (Demerol) or pentazocine (Talwin) is being prescribed for a patient, the question should not be, Why not use it?, instead the question should be, Why use it?. In other words, why is propoxyphene (Darvon), meperidine (Demerol) or pentazocine (Talwin) the only option that we can use in this patient given that there are many other newer pain medications that will provide greater pain relief and have a much smaller risk of causing side effects. Greg Holmquist, Pain / Palliative Care Specialist

14

X. PHARMACOLOGICAL DECISIONS (continued) B. Drug Selection #" Agents to use with caution:

NSAIDs nonselective agents [e.g. Ibuprofen (Motrin), naproxen (Aleve)]


Few long-term controlled trials in elderly, risks of multiple adverse drug reactions including PUD, renal insufficiency, GI bleeding. ! COX-II selective NSAIDs [e.g. celecoxib (Celebrex, rofecoxib (Vioxx) Few long-term controlled trials in elderly, appear to have less long-term risk of GI bleeds. Unknown long-term effects on renal function, liver damage.

forms

Codeine (most commonly prescribed with acetaminophen as Tylenol #3, for milder

of pain, side effects such as constipation, nausea/vomiting limit its use.

Acetaminophen (Tylenol) - acute and chronic toxicities affecting liver and kidneys
Question - What is the maximum dose of acetaminophen we should administer to patients on a daily basis? Answer - For the average patient, no more than 2-4 grams of acetaminophen per day. REMEMBER: Add up the amount of acetaminophen in all medications the patient is receiving. EXAMPLES:
Vicodin - has 500 mg of acetaminophen per tablet Vicodin ES - has 750 mg of acetaminophen per tablet Lorcet- has 650 mg of acetaminophen per tablet Percocet has 325 mg of acetaminophen per tablet Tylox has 500 mg of acetaminophen per capsule

Tramadol (Ultram)- low ceiling of activity, constipation, expensive Butorphanol (Stadol)- nasal spray low ceiling of activity, misuse by patients Methadone - long half-life leads to accumulation of drug which increases risk of
over-sedation and respiratory depression

Fentanyl transdermal (Duragesic) ISSUES: not effective for pain out of control,
difficulty with dose titration, fluctuations in blood levels, variability in duration of action, economics.

When is it appropriate to use the transdermal fentanyl patch??? Checklist for appropriateness Patient should have BOTH of these criteria: % Patient has moderate to severe pain (hopefully patient is not opioid nave) AND % Patients pain is relatively stable (not fluctuating) In addition, patient should have ONE of these criteria: % Patient cannot take oral therapies; or, % Patient has documented history of ALLERGIC reactions to the morphine class of drugs.

15

X. PHARMACOLOGICAL DECISIONS (continued) B. Drug Selection (continued)


#"

MORPHINE Thought of by most practitioners as the strongest of pain medications Has many positive features in terms of safety, ability to control pain, variety of dosing routes. Has many negative features in terms of side effects, societal stimatization, myths and fears. Long-acting morphine (MS Contin, Purdue Pharma, Oramorph SR, Roxane, Kadian)
Available strengths - 15 mg, 30 mg, 60 mg, 100 mg, (200 mg - MS Contin only) Oral bioavailability: 20 % to 45% Onset of analgesia is within 2 hours Duration of action: MS Contin: most patients Q12 H dosing, about 20% need Q 8 H dosing Oramorph SR: most pts Q 8 H dosing, about 20% can utilize Q 12 H Kadian: some patients Q 24 H, some patients Q 12 H, some pts Q 18 H Nausea, vomiting, confusion, oversedation a concern for many patients Is there a difference between MS Contin and Oramorph SR and Kadian? FDA has declared that Oramorph,Kadian and MS Continare NOT bioequiv. Metabolite issue Is morphine still the drug of choice for severe pain?

#" SEMI-SYNTHETIC FORMS OF MORPHINE Oxycodone Available for many years as a short-acting formulation in combination with acetaminophen (e.g. percocet, tylox, roxicet), and as a short-acting formulation without acetaminophen as tablets, capsules and liquid formulations. Long-acting oxycodone (OxyContin, Purdue Pharma)
Available strengths - 10 mg, 20 mg, 40 mg, 80 mg. High oral bioavailability (60 % to 87%). Onset of analgesia is within one hour. Single-entity agent no need for concern with acetaminophen toxicity. No ceiling effect to analgesia. Q 12 H dosing.

Potential advantages of semi-synthetic forms of morphine (e.g. oxycodone, hydromorphone) Same efficacy as morphine no ceiling to the analgesic effect Improved delivery system of new long-acting formulations of oxycodone and hydromorphone
Quicker onset, easier titration. Biphasic absorption, leading to less fluctuations in blood levels. Metabolites of oxycodone and hydromorphone do not have the side-effect profile as morphines metabolites

Semi-synthetic morphine drugs represent a cleaner, less side effect alternative Less societal stigmatization than that which occurs with morphine. 16

X. PHARMACOLOGICAL DECISIONS (continued) #" Opioid selection decisions Base agent selection on the following factors: %" Efficacy Will the patients pain condition likely respond to opioid therapy? How
strong of an opioid does the patients pain condition warrant? %" Short acting versus long acting agents? Is the patients pain condition sporadic or continuous? Is compliance with therapy a potential issue? What are your goals of therapy? Is pain affecting sleep? Is there potential for diversion? %" Adverse effects Is the patient at risk for specific adverse effects? Does the patient have impaired renal function?

%" Route Can the patient take oral therapy?


%" Economics Are there economic issues that need to be addressed?

C. Side effect management:


Constipation - need to soften stool and increase smooth muscle motility Nausea/vomiting - determine underlying cause (GI irritation vs effect on CTZ vs vestibular) Respiratory depression - sedation precedes resp. depression. Hypersensitivity reactions:

D. Route selection
Oral
simple, least costly, available as long-acting preferred by most patients - greater patient self-control recommended by WHO, AHCPR guidelines, American Pain Society etc wide variety of dosage strengths, formulations. easy to use alternative when unable to use oral route. sometimes socially unacceptable. minimal dosage strengths available. bitter taste may be a problem for some patients. unpredictable absorption, especially in patients close to death. quick onset on action. limited dosage strengths. difficult to titrate, slow onset, expensive

Rectal

Sublingual

Transdermal long-acting alternative when unable to use oral route. Parenteral


invasive, expensive, more rapid onset, more predictable bioavailability Intramuscular - painful, slower onset, can be irritating to tissue, variable absorption PCA safe, improved pain relief, greater patient satisfaction Note: When using PCA for patients with cancer pain, the bolus dose should be set at 25% of the basal rate with a lock-out of 15 minutes. Example: A patient receiving a continuous infusion of morphine at 20 mg per hour should have a PCA bolus dose available at 5 mg with a lock-out of 15 minutes. Continuous infusions - can be continuous IV or SQ Epidural / intrathecal - most expensive route, less than 2% of chronic pain patients need this form of therapy.

17

References: Acute Pain Management General Overview: Reuben SS, Connelly NR, Maciolek H. Postoperative analgesia with controlled-release oxycodone for outpatient anterior cruciate ligament surgery. Anest Analg 1999;88:1286-91. Smythe M. Patient-controlled analgesia: A review. Pharmacotherapy 1992;12:132-143. Sidebotham D, Dijkhuizen MRJ, Schug SA. The safety and utilization of patient-controlled analgesia. J Pain Symptom Manage 1997;14:202-209. Acute Pain Management Guidelines Panel. Acute Pain Management: Operative or medical procedures and trauma. Clinical practice guideline. AHCPR Pub. No. 92-0032. Rockville, MD: Agency for Health Care Policy and Research, Public health Service, US Department of Health and Human Services; 1992. Ginsberg B, Sinatra R, Crews J et al. Conversion from IV PCA morphine to oral controlled-release oxycodone tablets for postoperative pain management. IASP Meeting Presentation. Orlando, FL 1998. Hagmeyer KO, Mauro LS, Mauro VF. Meperidine-related seizures associated with patient-controlled analgesia pumps. Annual of Pharmacotherapy 1993;27:29-31. Jacox A et al. Managing acute pain. Am J Nurs 1992;92:49-55. Justins DM. Postoperative pain: A continuing challenge. Ann R Coll Surg Engl 1992;74:78-79. Chumbley G, Hall GM, Salmon P. Patient-controlled analgesia: an assessment by 200 patients. Anaesthesia 1998;53:216-221. Agents: Narcessian E, Cleville A, Chen A. Oxycontin following unilateral total knee arthroplasty: A double-blind randomized controlled trial. Amer Pain Society 17th annual meeting abstract booklet. 1998;17:127,Abstract # 719. Babul N, Provencher L, Laberge F, et al. Comparative efficacy and safety of controlled-release morphine suppositories and tablets in cancer pain. J Clin Pharmacol 1998;38:74-81. Maddocks I, Somogyi A, Abbott F, et al. Attenuation of morphine-induced delirium in palliative care by substitution with infusion of oxycodone. J Pain Symptom Manage 1998;12:182-189. Bruera E, Belzile M, Pituskin E et al. Randomized, double-blind, cross-over trial comparing safety and efficacy of oral controlled-release oxycodone with controlled-release morphine in patients with cancer pain. JCO 1998;16:3222-3229. Janicki PK. Pharmacology of morphine metabolites. Current Review of Pain. 1997;1:264-270. Olarte Nunez JM. Opioid-induced myoclonus. Eur J Palliative Care 1996;2:146-150. Ashby M, Fleming B, Wood M, et al. Plasma morphine and glucuronide (M3G and M6G) concentrations in hospice patients. J Pain Symptom Manage 1997;14:157-167. Kalso E and Vainio A. Morphine and oxycodone hydrochloride in the management of cancer pain. Clin Pharmacol Ther 1990;47:639-646. Sawe J. High-dose morphine and methadone in cancer patients: Clinical pharmacokinetic considerations of oral treatment. Clin Pharmacokinetics 1986;11:87-106. Brescia FJ, Walsh M et al. A study of controlled-release oral morphine (MS Contin) in an advanced cancer hospital. J Pain Symptom Manage 1987;2:193-198. Khojasteh A, Evans W et al. Controlled-release morphine sulfate in the treatment of cancer pain with pharmacokinetic correlation. J Clin Onc 1987;5:956-961. Mignault GG, Latreille J et al. Control of cancer-related pain with MS Contin: A comparison between 12-hourly and 8-hourly administration. J Pain Symptom Manage 1995;10:416-422. Hunt TL and Kaiko RF. Comparison of the pharmacokinetic profiles of two oral controlled-release formulations in healthy young adults. Clin Ther 1991;13:482-488. Yee LY and Lopez JR. Transdermal fentanyl. Ann Pharmacotherapy 1992;26:1393-1399. Kaiko R, Benzinger DP et al. Pharmacokinetic/pharmacodynamic relationships of controlled-release oxycodone. Clin Pharmacol Therapeutics 1996;59:52-61.

JCAHO and Acute/Post-operative Pain Management

18

Recovery Profiles and Costs of Anesthesia for Outpatient Unilateral Inguinal Herniorrhaphy
Dajun Song, MD, PhD*, Nancy B. Greilich, MD*, Paul F. White, Mehernoor F. Watcha, MD, and W. Kendall Tongier, MD*
PhD, MD*, Departments of Anesthesiology and Pain Management, *University of Texas Southwestern Medical Center at Dallas, Dallas, Texas, and Childrens Hospital of Philadelphia, Philadelphia, Pennsylvania

The use of an ilioinguinal-hypogastric nerve block (IHNB) as part of a monitored anesthesia care (MAC) technique has been associated with a rapid recovery profile for outpatients undergoing inguinal herniorrhaphy procedures. This study was designed to compare the cost-effectiveness of an IHNB-MAC technique with standardized general and spinal anesthetics techniques for inguinal herniorrhaphy in the ambulatory setting. We randomly assigned 81 consenting outpatients to receive IHNB-MAC, general anesthesia, or spinal anesthesia. We evaluated recovery times, 24-h postoperative side effects and associated incremental costs. Compared with general and spinal anesthesia, patients

receiving IHNB-MAC had the shortest time-to-home readiness (133 68 min vs 171 40 and 280 83 min), lowest pain score at discharge (15 14 mm vs 39 28 and 34 32 mm), and highest satisfaction at 24-h follow-up (75% vs 36% and 64%). The total anesthetic costs were also the least in the IHNB-MAC group ($132.73 33.80 vs $172.67 29.82 and $164.97 31.03). We concluded that IHNB-MAC is the most costeffective anesthetic technique for outpatients undergoing unilateral inguinal herniorrhaphy with respect to speed of recovery, patient comfort, and associated incremental costs. (Anesth Analg 2000;91:876 81)

ocal anesthesia with IV sedation (so-called monitored anesthesia care [MAC]), spinal (subarachnoid) anesthesia, and general anesthesia are all commonly used anesthetic techniques for outpatients undergoing inguinal herniorrhaphy procedures (1 8). In the current cost-conscious environment, it is important to examine the impact of anesthetic techniques on the recovery process after ambulatory surgery because prolonged recovery times and perioperative complications increase the cost of patient care. In addition, patient satisfaction is improved when the anesthetic technique chosen for the procedure is associated with a small incidence of postoperative side effects. The ilioinguinal-hypogastric nerve block (IHNB) also decreases postoperative pain after MAC in outpatients undergoing inguinal hernia repair procedures (1,2).

We hypothesized that the technique of MAC with an IHNB and propofol sedation would be superior to both general and spinal anesthetic techniques with respect to its recovery and side effects profile. Therefore, this study was designed to evaluate the recovery times, side effects, patient satisfaction, and associated anesthetic-related institutional costs with three standardized anesthetic techniques in outpatients undergoing unilateral inguinal herniorrhaphy procedures.

Methods
After obtaining institutional review board approval, 81 consenting ASA physical status I and II outpatients, ages 18 65 yrs, scheduled for a unilateral inguinal hernia repair procedure were enrolled in this clinical study. Patients with known cardiovascular, respiratory, renal/hepatic, or metabolic disease, active gastrointestinal reflux, as well as those with mental dysfunction, morbid obesity, or history of substance abuse, were excluded from the study. Patients were randomly assigned, according to a computergenerated, random-number table, to receive one of the following three anesthetic techniques: Group 1, a MAC technique consisting of an IHNB and propofol sedation; Group 2, general anesthesia with laryngeal
2000 by the International Anesthesia Research Society 0003-2999/00

Supported, in part, by The Ambulatory Anesthesia Research Foundation in Los Angeles, Ca, and the White Mountain Institute in Los Altos, CA (of which PFW is President). Presented, in part, at the annual meeting of International Anesthesia Research Society, Los Angeles, CA, 1999. Accepted for publication June 9, 2000. Address correspondence to Paul F. White, MD, Department of Anesthesiology and Pain Management, University of Texas Southwestern Medical Center, 5161 Harry Hines Blvd., F 2.208, Dallas, TX 75235-9068. Address e-mail to paul.white@utsouthwestern.edu.

876

Anesth Analg 2000;91:87681

ANESTH ANALG 2000;91:876 81

AMBULATORY ANESTHESIA SONG ET AL. RECOVERY PROFILES, COSTS, AND INGUINAL HERNIA REPAIR

877

mask airway; or Group 3, spinal anesthesia with a hyperbaric bupivacaine-fentanyl solution. All patients were premedicated with 2 mg of IV midazolam and 25 g IV fentanyl. In Group 1, patients received an IHNB with a 30 mL of mixture containing 0.25% bupivacaine and 1% lidocaine injected through the oblique muscles approximately 1.5 cm medial to the anterior superior iliac spine. A 75 g kg1 min1 IV propofol infusion, was started after the IHNB and subsequently varied between 25 and 150 g kg1 min1 to maintain a level of sedation at which the patient readily responded to verbal or light tactile stimulation. Surgery was initiated approximately 8 10 min after the IHNB was completed. In Group 2, anesthesia was induced with 2.5 mg/kg IV propofol, and a laryngeal mask airway was placed for airway management. Anesthesia was initially maintained with 1% inspired sevoflurane in combination with 65% nitrous oxide in oxygen, and the inspired sevoflurane was varied between 0.5% and 2% with the patient breathing spontaneously. In Group 3, patients were administered spinal anesthesia using the midline approach with a 25-gauge pencil-point needle at the L2-3 or L3-4 intervertebral space with the patient in the sitting position. The subarachnoid injection contained a mixture of 1.21.5 mL of 0.75% bupivacaine and 25 g of fentanyl. Prior to skin incision and during surgery, the operative site (and genital-femoral nerve) was infiltrated with 10 mL of the solution containing 0.25% bupivacaine and 1% lidocaine in all three groups. The protocol also allowed the anesthesia provider to administer 2550 g IV boluses of fentanyl and 10 20 mg IV boluses of propofol to treat pain and purposeful movements, respectively, during the operation in all three groups. Patients in Groups 1 and 3 who failed to achieve adequate surgical or anesthetic conditions were converted to general anesthesia with propofol and sevoflurane/nitrous oxide. Recovery times were recorded from the end of surgery to awakening (opening eyes on verbal command), orientation (correctly stating the date, place, and person) and home readiness (meeting the criteria for discharge home from the day surgery unit). Before leaving the operating room (OR), all patients were evaluated for fast-track eligibility (score 12) by the attending anesthesiologist (9). Those who achieved fast-track eligibility prior to leaving the OR were taken directly to the Phase 2 recovery area, bypassing the postanesthesia care unit (PACU). A 100-mm visual analog scale (VAS), with 0 none to 100 most severe, was used to assess pain and nausea prior to anesthesia administration (baseline), on arrival at the recovery area, and subsequently at 30 min intervals until discharge. Home readiness was assessed at 15 min intervals in the Phase 2 recovery unit by a blinded observer. At 24-h postoperatively, adverse events were assessed by a blinded investigator (DS)

Table 1. Basic Cost Assumptions for the Economic Analysis Cost (USD) Anesthetic equipment costs Anesthetic circuit, suction Laryngeal mask airway (cleaning, sterilizing, 50 uses) Infusion pump tubing and disposable Spinal tray, gloves, needle, lidocaine Salter cannulae Anesthetic drug costs 200 mg propofol 100 g fentanyl Sevoflurane 250 mL Lidocaine 2% (20 mL) Bupivacaine 0.5% (30 mL) Recovery room drug costs Droperidol 4 mg ondansetron Hydrocodone/acetaminophen 25 mg meperidine Recovery room resources costs Emesis management (per episode) Oxygen delivery equipment Nursing labor costs (hourly) Operating room costs (hourlytwo nurses, one aide)
USD United States dollars.

7.58 7.75 1.68 10.54 2.55 15.00 1.89 189.00 0.60 2.92 0.19 16.35 0.50 0.66 2.50 0.66 22.00 55.00

using a standardized postoperative telephone interview. Patient satisfaction with the anesthetic technique was evaluated using a three-point scoring system of 1 poor, 2 good, or 3 excellent. An a priori power analysis based on previously published data, suggested that a minimum of 25 patients in each group would be required to detect a 30% reduction in total institutional costs, with a power of 90% at the 0.05 level of significance. This group size would also be adequate to detect a 30% difference in VAS scores for pain and nausea with a power of 0.8 ( 0.05). Data analysis was on an intent-to-treat basis, where data from patients who required general anesthesia when the local/sedation or spinal anesthetic technique failed were included in the original assignment group. Continuous data were analyzed using one-way analysis of variance and if significant differences were noted, a Student-Neuman-Kuels test was used for intergroup comparisons. Categorical data were analyzed using the 2 test with Yates continuity correction or Fishers exact test, where appropriate, with P 0.05 considered statistically significant. The perspective used in the cost analysis was that of the chief financial officer of the ambulatory surgical center. The marginal costs of drugs and resources (Table 1) were calculated based on the actual acquisition costs to the center and not based on patient charges. These included the costs of anesthetic drugs administered in the OR and analgesic and antiemetic drugs administered in the recovery area. Drugs and resources common to all

878

AMBULATORY ANESTHESIA SONG ET AL. RECOVERY PROFILES, COSTS, AND INGUINAL HERNIA REPAIR

ANESTH ANALG 2000;91:876 81

Table 2. Patient Demographic Characteristics, Anesthesia, Surgery, and Recovery Times for Ilioinguinal Hypogastric Nerve Block-Monitored Anesthesia Care (IHNB-MAC), General Anesthesia, or Spinal Anesthesia for Inguinal Herniorrhaphy Proceduresa IHNB-MAC (Group 1) Number (n) Age (yr) Weight (kg) Height (cm) Sex (M/F) ASA physical status (I/II) (n) Time of anesthesia (min) Time of surgery (min) Intraoperative MAP (mm Hg) Intraoperative HR (bpm) Intraoperative RR (bpm) ETco2 (mm Hg) Propofol (mg) Fentanyl (g) IV Fluids (mL) Recovery times (min)b Awakening Orientation Phase 1 PACU (min) Phase 2 unit (min) Home-readiness (min) Actual discharge (min)
a b

General anesthesia (Group 2) 28 36 16 75 10 171 14 24/4 20/8 119 29 93 31 72 11 67 9 13 3 44 9 166 41* 125 76 1189 422 5 2* 11 5* 40 13* 168 58 171 40* 208 56*

Spinal anesthesia (Group 3) 25 39 14 73 14 169 8 20/5 11/14 116 24 91 22 74 8 62 8 14 3 37 4 84 44* 66 49b 1112 286 0 1* 1 2* 35 22* 276 86* 280 80* 309 83*

28 42 18 73 9 177 8 26/2 16/12 109 23 86 21 79 9 67 8 15 3 39 5 312 192 94 44 1230 354 32 54 5 14 153 67 133 68 158 71

Values are mean sd and numbers. Following discontinuation of the anesthetic. HR heart rate, MAP mean arterial pressure, RR respiratory rate, PACU postanesthesia care unit. * P 0.05 versus IHNB-MAC group. P 0.05 versus general anesthesia group.

three groups (electrocardiogram leads, pulse oximeter probes, IV catheters, and administration sets, etc) were not included, but the cost of wasted drugs was included. The cost of sevoflurane was calculated using the formula (10): cost (delivered concentration fresh gas flow time molecular weight cost of 1 mL)/(2412 density of sevoflurane). The cost of resources used in the recovery areas for managing and treating postoperative pain and nausea was included in the total costs. Nursing labor costs were based on the actual time spent by the nurse with a patient and prorated for the number of patients cared for at that time. For patients in the PACU, the nurse/patient ratio was 1:2 and in the Phase 2 recovery area it was 1:5, in keeping with the recommendations of the American Association of PACU Nurses. The total costs of each anesthetic technique were calculated by summing the costs of drugs, nursing labor, and resources used.

Results
There were no statistically significant differences among the three anesthetic treatment groups with respect to demographic characteristics, duration of anesthesia and surgery, and intraoperative mean arterial

pressure, heart rate, respiratory rate, end-expiratory carbon dioxide values, as well as the amount of IV fluid administered during the operation (Table 2). Operating conditions and analgesia were unsatisfactory in two patients in Group 1 and one patient in Group 3. These three patients required general anesthesia for completion of the operation. The total dosage of propofol used during surgery was largest in Group 1 and significantly different from the other two groups (Table 2). Intraoperative fentanyl requirements were significantly larger in Group 2 compared with Group 3 (Table 2). Patients in Groups 1 and 3 had faster awakening and orientation times than patients in Group 2. With the exception of the two patients who required rescue general anesthesia, all patients in Group 1 were transferred directly from the OR to the Phase 2 recovery area. In Group 3, 16% of the patients were judged to be fast-track eligible and were taken directly to the Phase 2 recovery unit. The time-to-home readiness (Table 2) and the maximum postoperative pain score (Table 3) were significantly decreased in Group 1 compared with the other two groups. However, the percentages of patients taking oral pain medication after discharge home were similar with all three techniques

ANESTH ANALG 2000;91:876 81

AMBULATORY ANESTHESIA SONG ET AL. RECOVERY PROFILES, COSTS, AND INGUINAL HERNIA REPAIR

879

Table 3. Anesthetic-Related Side Effects and Patient Satisfaction in the Ilioinguinal Hypogastric Nerve BlockMonitored Anesthesia Care (IHNB-MAC), General Anesthesia, or Spinal Anesthesia for Inguinal Herniorrhaphy Proceduresa General Spinal IHNB-MAC anesthesia anesthesia (Group 1) (Group 2) (Group 3) Postoperative side effects (n[%]) Backache Drowsiness Headache Knee weakness Muscle aches Nausea and/or vomiting Pruritus Sore throat Urine retention Maximum nausea VAS (mm) Maximum pain VAS (mm) Oral analgesia after discharge (n[%]) Satisfaction with anesthetic technique Poor Good Excellent

groups during the intraoperative period, but was significantly lower in Group 1 during the postoperative period. The combined cost of drugs and supplies used in the postoperative period was significantly higher in the general anesthesia group compared with the other two groups. The total perioperative cost was significantly lower in the IHNB-MAC group compared with the other two groups, but did not differ between the general and spinal anesthesia groups.

0 4 (14) 2 (7) 3 (11) 0 2 (7) 0 0 0 15 15 14 16 (57)

0 15 (54)* 4 (14) 1 (4) 2 (7) 17 (61)* 0 6 (22)* 0 27 27* 39 28* 18 (64)

6 (24)* 3 (12) 3 (12) 3 (12) 0 3 (12) 6 (24)* 2 (8) 5 (20)* 4 1 34 32* 17 (68)

Discussion
This study demonstrates that the use of an IHNB with propofol sedation for inguinal herniorrhaphy provides significant advantages over both general and spinal anesthesia. Patients receiving an IHNB-MAC technique had a shorter time-to-home readiness and lower pain scores at discharge. This technique was also associated with the lowest overall cost and highest patient satisfaction scores. In other studies, the use of a MAC technique for inguinal hernia repair has been found to have the additional advantages of early postoperative mobilization (4,5,7,11,12) and decreased incidences of urinary retention (4,6,13,14), nausea, vomiting, and sore throat (2,5,6). Furthermore, the ability to test the integrity of the repair during the operation (3 6,13) with a MAC anesthetic is another advantage compared with spinal or general anesthesia. The combination of high patient satisfaction, low cost, and early discharge suggests that the highest quality (cost/outcome) anesthetic was achieved with the IHNB-MAC technique. Cost estimates of various anesthetic regimens are available, but many of these pharmacoeconomic studies have limited cost considerations to only the acquisition cost of the drugs; and not the total expenses associated with the technique used. The total cost should include both the acquisition cost of drugs and the labor required for managing side effects (PONV, pain, drowsiness, bladder dysfunction). Since nursing personnel costs constitute a major proportion of expenses in the OR and recovery areas, anesthetic techniques associated with a greater need for nursing services will be more expensive (15). This study included nursing labor costs in the total cost of an anesthetic regimen, using the cost accountants standard concept of opportunity cost. This assumes that the time a nurse spends with one patient is time away from other activities that will then have to be performed by another salaried individual. However, it may be inappropriate to assume there is a linear relationship between labor cost and the time spent providing a clinical service (15). There is a much clearer relationship between lower cost and bypassing of the Phase 1 recovery unit. The major labor cost in the PACU is related to the peak number of patients admitted to the unit at any given

0 7 (25) 21 (75)

0 18 (64)* 10 (36)*

0 9 (36) 16 (64)

a Values are mean sd, numbers (n), and percentages (%). * P 0.05 versus IHNB-MAC group. P 0.05 versus general anesthesia group.

(Table 3). All general anesthesia patients initially recovered in the PACU. The incidence of side effects, namely sore throat, drowsiness, and postoperative nausea and vomiting (PONV), as well as the maximum VAS nausea scores, were significantly higher in the general anesthesia group (Table 3). Patients receiving spinal anesthesia had the highest incidence of postoperative pruritus, urinary retention, lumbar backache (Table 3), and the longest time to achieve home discharge criteria (Table 2). Finally, patient satisfaction with anesthesia is summarized in Table 3. None of the study patients reported a score of poor. However, compared with the general anesthetic technique, the use of IHNB-MAC was associated with significantly higher patient satisfaction scores. The cost of drugs used during the intraoperative period differed significantly in the three groups, with the lowest cost in Group 3 and highest in Group 2 (Table 4). The cost of anesthetic supplies was lowest in Group 1. Labor cost did not differ among the three

880

AMBULATORY ANESTHESIA SONG ET AL. RECOVERY PROFILES, COSTS, AND INGUINAL HERNIA REPAIR

ANESTH ANALG 2000;91:876 81

Table 4. Incremental Costs in the Operating Room (OR) and the Postanesthesia Care Units Associated With Ilioinguinal Hypogastric Nerve Block-Monitored Anesthesia Care (IHNB-MAC), General Anesthesia, or Spinal Anesthesia for Inguinal Herniorrhaphy Proceduresa IHNB-MAC (Group 1) Intraoperative costs Drugs Supplies OR non-labor OR labor TOTAL COSTS Recovery costs Drugs Supplies Nursing labor Phase 1 Phase 2 Total TOTAL COSTS Perioperative costs Total drug cost Total supplies Total resources used Total labor costs TOTAL COSTS
a Values are mean sd (in US dollars). * P 0.05 versus IHNB-MAC group. P 0.05 versus general anesthesia group.

General anesthesia (Group 2) 42.62 9.88* 13.83 1.12* 56.45 9.88* 109.87 9.88 166.32 26.91* 8.82 8.51* 0.86 0.68* 7.11 2.77* 13.10 6.65* 20.21 7.79* 29.88 9.68* 51.44 14.80* 14.69 0.68* 66.12 15.09* 130.08 27.91* 172.67 29.82*

Spinal anesthesia (Group 3) 17.13 10.42* 13.84 2.77* 30.97 12.98* 107.32 22.17 138.30 28.01* 1.03 3.23* 0.73 0.37* 6.34 4.10* 19.04 10.12* 25.39 10.31* 27.15 11.14* 18.16 9.76* 14.58 2.78* 35.74 12.39 132.71 23.89* 164.97 31.03*

34.66 17.47 5.22 3.63 39.88 19.46 102.63 19.46 142.51 22.74 0.15 0.31 0.05 0.17 0.85 2.57 11.56 5.21 12.41 6.68 12.61 6.84 34.81 17.56 5.27 3.80 40.07 19.67 115.05 26.67 132.73 33.80

time. Therefore, even if a patient spends an additional 15 to 30 minutes in the PACU, institutional costs may not be affected unless overtime cost is incurred. Fast tracking may also permit the use of fewer nurses and a mix of lower-wage nursing aides with registered nurses. With the exception of two patients who required a rescue with general anesthesia, all patients in the IHNB-MAC group met the PACU discharge criteria prior to leaving the OR and were able to bypass the PACU, contributing to a shorter time to discharge home compared with general and spinal anesthesia groups. However, a criticism of the study is that all patients receiving general anesthesia were required to be admitted to the PACU. If nursing practices mandate a minimum stay in the various recovery areas, there may not be any financial benefit to an institution from the faster recovery profiles associated with the newer anesthetic drugs. The time-to-home readiness is a clinical determination indicating completion of the early recovery process. Factors contributing to delays in the time-tohome readiness include drowsiness, nausea, vomiting, inability to void, postural hypotension, prolonged motor blockade, and administrative (and social) delays (1,6 8,11,12,16). The longer time-to-home readiness with spinal (versus general) anesthesia is probably related to the residual motor and sympathetic blockade. Even with an IHNB, ambulation can be delayed

by transient femoral nerve palsy when the local anesthetic solution is injected deep to the internal oblique muscle (17,18). Inadequate pain control in the postoperative period can also contribute to prolonging the time-to-home readiness and increasing patient dissatisfaction (19 21). The patients in the IHNB-MAC group were found to have lower pain scores even though the patients in the spinal and general anesthesia groups also received local anesthesia at the incision site. Previous studies (2,3,6,8) have reported longer times-tofirst analgesia after herniorrhaphy with the use of local infiltration, but these studies vary as to the technique of local anesthetic administration. Although patients receiving IHNB had lower discharge pain scores, their requirements for oral pain medications after discharge did not differ from the other two treatment groups. Spinal anesthesia can provide for a profound conduction block and preemptive analgesia while minimizing complications associated with general anesthesia (PONV, sore throat) (6,8,22). However, the popularity of spinal anesthesia for outpatient surgery has been tempered by concerns regarding transient radicular irritation, urinary retention, and postdural puncture headache (14,23,24). Transient radicular irritation occurs in up to 5% of patients receiving lidocaine, but appears in 1% receiving bupivacaine (23).

ANESTH ANALG 2000;91:876 81

AMBULATORY ANESTHESIA SONG ET AL. RECOVERY PROFILES, COSTS, AND INGUINAL HERNIA REPAIR

881

Although 24% of patients in the spinal anesthesia group complained of mild lumbar discomfort postoperatively, there were no reports of radiating back discomfort. Unfortunately, the residual motor and sympathetic blockade with bupivacaine led to a prolonged recovery and delayed discharge. General anesthesia remains the technique of choice for uncooperative or anxious patients, difficult repairs (reoperation after a mesh repair), and when a local anesthetic technique fails to provide adequate surgical conditions (4). In our study, two patients in the IHNBMAC group and one in the spinal anesthesia group required conversion to general anesthesia for completion of the procedure. Data from these patients were included in their original group assignment and the analysis was performed on an intention-to-treat basis. The rationale for this decision was based on the fact that it was reasonable to expect increased costs and decreased patient satisfaction in these subjects, and we felt that the study should reflect the real-world situation where failure of local anesthetic-based MAC techniques does occur. In conclusion, the use of IHNB with propofol sedation for outpatients undergoing inguinal herniorrhaphy resulted in a shorter time-to-home readiness, lower pain scores at discharge, greater patient satisfaction, and lower associated, incremental costs compared with general and spinal anesthesia. In situations where fast tracking can provide benefits for the patient and the health care system, this MAC technique would appear to offer advantages over both general and spinal anesthetic techniques for inguinal herniorrhaphy procedures.

References
1. Harrison CA, Morris S, Harvey JS. Effect of ilioinguinal and iliohypogastric nerve block and wound infiltration with 0.5% bupivacaine on postoperative pain after hernia repair. Br J Anaesth 1994;72:6913. 2. Ding Y, White PF. Post-herniorrhaphy pain in outpatients after preincision ilioinguinal-hypogastric nerve block during monitored anaesthesia care. Can J Anaesth 1995;42:125. 3. Glassow F. Inguinal hernia repair using local anesthesia. Ann Roy Coll Surg Engl 1984;66:3827.

4. Schumpelick V, Treutner KH, Arlt G. Inguinal hernia repair in adults. Lancet 1994;344:3759. 5. Teasdale C, McCrum A, Williams NB, Horton RE. A randomized controlled trial to compare local with general anesthesia for short stay hernia repair. Ann Roy Coll Surg 1982;64:238 42. 6. Young DV. Comparison of local, spinal, and general anesthesia for inguinal herniorrhaphy. Am J Surg 1987;163:560 3. 7. Behnia R, Hashemi F, Stryker SJ, et al. A comparison of general versus local anesthesia during inguinal herniorrhaphy. Surgery 1992;174:277 80. 8. Tverskoy M, Cozacov C, Ayache M, et al. Postoperative pain after inguinal herniorrhaphy with different types of anesthesia. Anesth Analg 1990;70:29 35. 9. White PF, Song D. New criteria for fast tracking after outpatient anesthesia: A comparison with the Aldretes scoring system. Anesth Analg 1999;88:1069 72. 10. Rosenberg MK, Bridge P, Brown M. Cost comparison: a desflurane- versus a propofol-based general anesthetic technique. Anesth Analg 1994;79:8525. 11. Dierking GW, Ostergaard E, Ostergardtt HT, Dahl JB. The effect of wound infiltration with bupivacaine versus saline on postoperative pain and opioid requirements after herniorrhaphy. Acta Anaesthesiol Scand 1994;38:289 92. 12. Nehra D, Gemmell L, Pye JK. Pain relief after inguinal hernia repair: a randomized double blind study. Br J Surg 1995;82: 12457. 13. Flanagan L, Bascom JU. Repair of the groin: outpatient approach with local anesthesia. Surg Clin North Am 1984;64:257 67. 14. Finley RK, Miller SF, Jones LM. Elimination of urinary retention following inguinal herniorrhaphy. Am Surg 1991;57:486 9. 15. Watcha MF, White PF. Economics of anesthetic practice. Anesthesiology 1997;86:1170 96. 16. Callesan T, Kehlet H. Postherniorrhaphy pain. Anesthesiology 1997;87:1219 30. 17. Price R. Transient femoral nerve palsy complicating ilioinguinal nerve blockade for inguinal herniorrhaphy. Br J Surg 1995;82: 137 8. 18. Rosario DJ, Skinner PP, Raftery AT. Transient femoral nerve palsy complicating preoperative ilioinguinal nerve blockade for inguinal herniorrhaphy. Br J Surg 1994;81:897. 19. Tong D, Chung F, Wong D. Predictive factors in global and anesthesia satisfaction in ambulatory surgical patients. Anesthesiology 1997;87:856 64. 20. Marshall SI, Chung F. Discharge criteria and complications after ambulatory surgery. Anesth Analg 1999;88:508 17. 21. Korttila K. Recovery from outpatient anaesthesia, factors affecting outcome. Anaesthesia 1995;50(Suppl):22 8. 22. Ryan JA, Adye BA, Jolly PC, Mulroy MF. Outpatient inguinal herniorrhaphy with both regional and local anesthesia. Am J Surg 1984;148:313 6. 23. Pollack JE, Neal JM, Stephenson CA, Wiley CE. Prospective study of the incidence of transient radicular irritation in patients undergoing spinal anesthesia. Anesthesiology 1996;84:13617. 24. Halpern S, Preston R. Post dural puncture headache and spinal needle design metaanalyses. Anesthesiology 1994;81:1376 83.

Clinical review

Recent advances Management of patients in fast track surgery


Douglas W Wilmore, Henrik Kehlet
Surgery is slowly undergoing revolutionary changes due to newer approaches to pain control, the introduction of techniques that reduce the perioperative stress response, and the use of minimally invasive operations. Subsequently, many surgical procedures (such as arthroscopic surgery, laparoscopic cholecystectomy, eye surgery, sterilisation procedures, herniorrhaphy, and cosmetic operations) are routinely performed on an outpatient basis. Recently published pilot studies suggest that when these newer approaches are used in patients undergoing more complex elective surgical procedures, postoperative complications can be reduced, length of hospital stay decreased, and time to recovery shortened. This review of recent advances made in this newly developing specialty of fast track surgery will emphasise techniques that facilitate early recovery after major surgical procedures.

Recent advances
Newer techniques in surgery and anaesthesia that reduce the postoperative stress response are improving surgical outcome Use of these methods in day surgical units will be extended to more complex surgical procedures, thus decreasing length of time in hospital Regional anaesthesia and minimally invasive operative techniques are central to these changes Shortened postoperative recovery should be the focus of rehabilitation care units, which optimise pain relief, mobilisation, and nutrition Early patient discharge will be accompanied by functional recovery and presumably less morbidity

Laboratories for Surgical Metabolism and Nutrition, Department of Surgery, Brigham and Womens Hospital, Harvard Medical School, Boston, MA 02215, USA Douglas W Wilmore Frank Sawyer professor of surgery Department of Surgical Gastroenterology, Hvidovre University Hospital, Hvidovre, Denmark Henrik Kehlet professor of surgery Correspondence to: D Wilmore dwilmore@ partners.org
BMJ 2001;322:4736

What is fast track surgery?


Fast track surgery combines various techniques used in the care of patients undergoing elective operations. The methods used include epidural or regional anaesthesia, minimally invasive techniques, optimal pain control, and aggressive postoperative rehabilitation, including early enteral (oral) nutrition and ambulation. The combination of these approaches reduces the stress response and organ dysfunction and therefore greatly shorten the time required for full recovery. Recent advances in understanding perioperative pathophysiology have indicated that multiple factors contribute to postoperative morbidity, length of stay in hospital, and convalescence (fig 1). Major improvements in surgical outcome may therefore require multifaceted interventions (fig 2). Ambulatory surgery has become routine for many procedures with a well documented record for safety and low morbidity, even in patients at high risk.1 2 Studies of fast track surgery have evaluated somewhat similar approaches toward larger operations which carry more risk (box). Preliminary results from predominantly non-randomised trials have been positive (table). These studies have included high risk elderly patients undergoing operations such as segmental colonic resection, prostatectomy, and aortic aneurysmectomy. These preliminary data indicate topics for further randomised trials; the data need to be confirmed and extended to include end points of reduced costs, preserved safety, and patient satisfaction.
BMJ VOLUME 322 24 FEBRUARY 2001 bmj.com

We searched Medline from 1980 to the present and reviewed the articles identified. This information was supplemented with our own research on the mediators of the stress response in surgical patients, the use of epidural anaesthesia in elective operations, and pilot studies of fast track surgical procedures with the multifaceted approach.12

Preoperative evaluation and education


Before any operation, including fast track surgery, organ function should be optimised for patients with cardiac disease, chronic obstructive lung disease,
Surgery Pain Stress response/organ dysfunction Nausea, vomiting, ileus Hypoxaemia, sleep disturbances Fatigue Immobilisation, semistarvation Drains/nasogastric tubes, restrictions Delayed recovery

Fig 1 Factors contributing to postoperative morbidity

473

Clinical review
improves outcome.15 Patients can access information on specific clinical procedures on www.facs.org/ public_info/operation/aboutbroch.html, which is provided by the American College of Surgeons.

Staff training/reorganisation and procedure specific care plans

Preoperative information and optimisation of organ function

Stress reduction Regional anaesthesia Minimal invasive operations Normothermia Pharmacological modifiers

Effective pain relief and prophylaxis for nausea and vomiting

Modification of perioperative care Early mobilisation Minimal use of tubes, drains, and catheters Oral nutrition

Optimising anaesthesia
Recent developments in techniques in anaesthesia have optimised conditions for surgeons to operate while allowing for very early recovery of vital organ function after major procedures. Thus, the introduction of rapid short acting volatile anaesthetics (for example, desflurane and sevoflurane), opioids (for example, remifentanil), and muscle relaxants have facilitated expansion of ambulatory surgery for minor to moderate procedures (see box). However, the same techniques may be used to facilitate early recovery and decreased need for prolonged monitoring and stay in recovery and high dependency wards after major procedures, although this issue has been less explored and documented than the use of such techniques for minor procedures.16 The use of anaesthetic techniques that provide for minimal carryover of opioid effects into the recovery period, supported by other non-opioid analgesic methods (see below), may minimise postoperative complications and facilitate recovery after major procedures. Most, if not all, postoperative organ dysfunction and morbidity associated with major operative procedures may be related to changes induced by stress caused by the operation. Neural blockade techniques have been developed in recent years to provide attenuation of the surgical stress response, thereby reducing postoperative organ dysfunction and allowing early recovery.17 After experimental studies showed that the peripheral and central nervous system was crucial in the initiation of the endocrine-metabolic response to injury, a vast amount of research has shown that regional anaesthetic techniques that use local anaesthetics can reduce the classic pituitary, adrenocortical, and sympathetic responses to surgery.17 Neurogenic blockade (either by administering a local anaesthetic in the spinal or epidural space or by using local anaesthetic techniques that block the nerve impulses from an area) improves postoperative nitrogen economy and glucose intolerance but does not modify inflammatory or immunological responses. Relevant to clinical care, continuous neural blockade for 24 to 48 hours is necessary for a pronounced reduction in perioperative stress in major surgery.17 Moreover, the systemic effects of local or regional anaesthesia/analgesia on the stress response are greatest in procedures on the lower body (lower extremities or pelvis) compared with upper abdominal and thoracic operations. The effects of regional anaesthetic techniques are manifest by improved pulmonary function, decreased cardiovascular demands, reduced ileus, and improved pain relief.17 A recent meta-analysis of regional anaesthetic studies showed a 30% reduction in morbidity compared with general anaesthesia.18
BMJ VOLUME 322 24 FEBRUARY 2001 bmj.com

Fast track surgery

Documentation Morbidity Safety Cost Patient satisfaction

Fig 2 Interventions needed for major improvement in surgical outcome

diabetes mellitus, and other disorders, according to current recommendations. Pharmacological means have been used to enforce abstinence in alcohol misusers, and this has resulted in lower morbidity and enhanced recovery in such patients.13 Prolonged (one to two months) cessation of smoking in the preoperative period should also be encouraged to reduce postoperative respiratory complications. Education of patients about perioperative care before the operation reduces the need for pain relief,14 can include instruction on relaxation techniques which can be used after the operation, reduces anxiety, and

Recent developments on fast track surgery from single centre studies


Operation Laparoscopic cholecystectomy Laparoscopic or vaginal hysterectomy Laparoscopic gastro-oesophageal reflux surgery Elective surgery for aortic aneurysm Carotid endarterectomy Mastectomy Lung lobectomy Prostatectomy Partial colectomy Hospital stay Ambulatory procedure3 Ambulatory procedure, 1 day4 Ambulatory procedure, 1 day5 3-4 days6 1-2 days7 Ambulatory procedure, 1 day8 1-2 days9 1-2 days10 2 days11

Examples of fast track surgery


Ambulatory or 24 hour surgery Extensive knee and shoulder reconstruction (laparoscopy/endoscopy) Vaginal hysterectomy Gastric fundoplication (laparoscopy/endoscopy) Splenectomy (laparoscopy/endoscopy) Adrenalectomy (laparoscopy/endoscopy) Donor nephrectomy (laparoscopy/endoscopy) Mastectomy Cholecystectomy (laparoscopy/endoscopy) Short stay surgery1 to 4 days Colectomy Total hip and knee replacement Aortic aneurysmectomy Pneumonectomy and lobectomy Radical prostatectomy Peripheral vascular reconstruction

474

Clinical review Operative techniques


Minimal invasive surgery The use of minimal invasive abdominal surgical techniques, such as laparoscopic cholecystectomy, have not reduced the early endocrine mediated metabolic response to surgery, but this approach has been associated with a decrease in various inflammatory responses and immunodysfunctions.19 Pulmonary function seems to be improved and postoperative ileus reduced with minimal invasive approaches.19 20 Other studies have reported less pain, shorter hospital stays, and reduced morbidity, not only in abdominal surgery but also in cardiothoracic, vascular, cerebral, and major orthopaedic procedures. The scientific basis for these effects remains incompletely understood, and more basic studies are necessary to improve our understanding of the influence of minimal invasive surgery on postoperative responses. Intraoperative normothermia Operating rooms are cold. Patients are inadequately clothed and receive anaesthetics which hamper their homeostatic defences to cold. As a result, patients undergoing operations lasting over two hours often become hypothermic, with a fall of core temperature of 2-4C. During rewarming cortisol and catecholamines are released, which augment the stress response of the operation.21 Keeping patients warm has been associated with a threefold decrease in the rate of wound infection, a reduction in operative blood loss, a decrease in untoward cardiac events, including ventricular tachycardia, and a reduction in nitrogen excretion and patient discomfort.21 Maintenance of a normal temperature during surgery is central to reducing the stress of the surgical procedure and reducing the risk of organ dysfunction. cal treatment should be used if nausea and vomiting are present. Postoperative pain should be vigorously treated as it may amplify the surgical stress responses and organ dysfunction and prolong recovery.26 Principles for optimising treatment of postoperative pain have been developed, providing pain relief which allows early movement. Improvement of pain management includes education of staff and patients, establishment of an acute pain service, and the use of multifaceted analgesic intervention.27 After operations of minor to moderate size patients should receive non-opioid analgesics, such as non-steroidal anti-inflammatory agents, to avoid side effects related to use of opioid drugs, which prolong recovery.26 Major surgical procedures with high intensity pain and subsequent organ dysfunction induced by stress require the use of invasive analgesic methods, such as continuous epidural analgesia, to hasten recovery.24 Optimal management of acute pain after major procedures is a prerequisite for fast track surgery and should be used for all surgical patients. Nausea, vomiting, and ileus The ability to resume a normal diet is essential for a successful fast track surgical programme after both minor and major procedures. Principles for rational prophylaxis and treatment of nausea and vomiting have been developed,28 29 and several agents including droperidol, antiserotonergic drugs, and analgesic regimens with reduced use of opioid drugs will reduce these symptoms. The use of multifaceted regimens for nausea and vomiting in combination with dexamethasone requires further evaluation. Postoperative ileus, which is predominantly caused by a combination of inhibitory neural sympathetic visceral reflexes and the intestinal inflammatory response, may be considerably alleviated by a combination of epidural local anaesthetics, analgesia with reduced used of opioid drugs, minimally invasive surgery, and pharmacotherapy.17 23 Preliminary studies show that such regimens, when combined with early enteral nutrition, may almost completely prevent paralytic ileus after colonic resection.11 25 The second to fifth postoperative day Recovery from an operation depends on several factors, including the resolution of pain and fatigue. Fatigue in the early postoperative period is related to altered sleep within the hospital setting because of noise, environmental disturbances, drugs, and possibly inflammatory factors.30 31 Loss of muscle strength and loss of weight because of reduced food intake have been related to fatigue, which occurs after a week or so.24 Reduction of surgical stress, early enteral nutrition, and mobilisation are therefore important interventions which counteract fatigue and aid recovery.

Postoperative care
The first 24 hours Nasogastric tubes should not be used routinely in patients undergoing elective gastrointestinal surgery. A large meta-analysis of 26 randomised trials concluded that routine use may, in fact, be detrimental by increasing the incidence of pneumonia and delaying early enteral feeding by nasogastric tube.22 Likewise, randomised trials of drains show little benefit after cholecystectomy, joint replacements, colon resection, thyroidectomy, and radical hysterectomy.23 24 Drains limit formation of seroma after mastectomy, but such wound drainage does not limit discharge from hospital. Bed rest is undesirable as it increases muscle loss and weakness, impairs pulmonary function, and predisposes to venous stasis and thromboembolism.24 All efforts should be made to enforce postoperative movement, which is possible with adequate pain relief. Oral intake is commonly limited in the postoperative period. Presently there are no available clinically effective drugs that enhance gastric emptying,25 and with the attenuation of ileus associated with epidural anaesthesia, oral intake can often be successfully initiated six hours after surgery, even after colonic operations which use an anastomosis.11 PharmacologiBMJ VOLUME 322 24 FEBRUARY 2001 bmj.com

The future
The initial promising results reported from fast track programmes raise the question of whether our traditional system of surgical care needs to be modified to improve surgical outcome. Shortened postoperative recovery may not necessarily require dependency on
475

Clinical review
traditional surgical units, which rely on monitoring and high tech intervention, but rather we may need to emphasise postoperative rehabilitation care units which optimise pain relief, mobilisation, and nutrition. Further developments in the specialty of fast track surgery will require more effective methods for reduction of perioperative stress, such as blockade32 and improved combinations of analgesia and anaesthesia. In addition, more sophisticated approaches toward minimally invasive surgery and possibly pharmacological modification of the inflammatory response may be necessary. Integration of these approaches with aggressive rehabilitative techniques is also required. In the future, the trend will be for shorter recovery periods after major operations. Importantly, the increased use of fast track surgery with shorter hospital stays will not necessarily lead to an increased burden on general practitioners as the patients will be discharged without the postoperative impairment of function usually observed and hopefully with less morbidity. Thus, with continued understanding of perioperative pathophysiology and improvements in perioperative care, it may not be unrealistic in the next few years for the insertion of a hip prosthesis, the excision of a large cancer, or the repair of an aortic aneurysm to be performed as day surgery.
Competing interests: None declared.
1 2 3 4 5 Warner MA, Shields SE, Chute CG. Major morbidity and mortality within 1 month of ambulatory surgery and anesthesia. JAMA 1993;270:1337-41. Mesei G, Chung F. Return hospital visits and hospital readmissions after ambulatory surgery. Ann Surg 1999;230:721-7. Mjland O, Raeder J, Aasboe V, Trondsen E, Buanes T. Outpatient laparoscopic cholecystectomy. Br J Surg 1997;84:958-61. Bran DF, Spellman JR, Summitt RL Jr. Outpatient vaginal hysterectomy as a new trend in gynecology. AORN J 1995;62:810-4. Trondsen E, Mjaland O, Raeder J, Buanes T. Day-case laparoscopic fundoplication for gastro-oesophageal reflux disease. Br J Surg 2000;87:1708-11. Podore PC, Throop EB. Infrarenal aortic surgery with a 3-day hospital stay: a report on success with a clinical pathway. J Vasc Surg 1999;29:787-92. Collier PE. Are one-day admissions for carotid endarterectomy feasible? Am J Surg 1995;170:140-3. Coveney E, Weltz CR, Greengrass R, Iglehart JD, Leight GS, Steele SM, et al. Use of paravertebral block anaesthesia in the surgical management of breast cancer: experience in 156 cases. Ann Surg 1998;227:496-501. Tovar EA, Roethe RA, Weissig MD, Lloyd RE, Patel GR. One-day admission for lung lobectomy: an incidental result of a clinical pathway. Ann Thorac Surg 1998;65:803-6. 10 Kirsh EJ, Worwag EM, Sinner M, Chodak GW. Using outcome data and patient satisfaction surveys to develop policies regarding minimum length of hospitalization after radical prostatectomy. Urology 2000;36:101-7. 11 Basse L, Hjort Jakobsen D, Billesbolle P, Werner M, Kehlet H. A clinical pathway to accelerate recovery after colonic resection. Ann Surg 2000;232:51-7. 12 Nierman E, Zakrzewski K. Recognition and management of preoperative risk. Rheum Dis Clin North Am 1999;25:585-622. 13 Tonnesen H, Rosenberg J, Nielsen HJ, Rasmussen V, Hauge C, Pedersen IK, et al. Effect of preoperative abstinence on poor postoperative outcome in alcohol misusers: a randomised controlled trial. BMJ 1999;318:1311-6. 14 Egbert LD, Bant GE, Welch CE, Bartlett MK. Reduction of postoperative pain by encouragement and instruction of patients. N Engl J Med 1964;207:824-7. 15 Daltroy LH, Morlino CI, Eaton HM, Poss R, Liang MH. Preoperative education for total hip and knee replacement patients. Arthritis Care Res 1998;11:469-78. 16 White, P.F. Ambulatory anesthesia advances into the new millennium. Anesth Analg 2000;90:1234-35. 17 Kehlet, H. Modification of responses to surgery by neural blockade: clinical implications. In: Cousins MJ, Bridenbaugh PO, eds. Neural blockade in clinical anesthesia and management of pain . Philadelphia: JB Lippincott, 1998:129-75. 18 Rodgers A, Walker N, Schug S, McKee H, van Zundert A, Dage D, et al. Reduction of postoperative mortality and morbidity with epidural or spinal anesthesia: results from an overview of randomised trials. BMJ 2000;321;1493-97. 19 Kehlet H. Surgical stress response: does endoscopic surgery confer an advantage? World J Surg 1999;23:801-7. 20 Shea JA, Berlin JA, Bachwich DR, Staroscik RN, Malet PF, McGuckin M, et al. Indications for and outcomes of cholecystectomy: a comparison of the pre and postlaparoscopic eras. Ann Surg 1998;227:343-50. 21 Sessler DI. Mild operative hypothermia. N Engl J Med 1997;336:1730-7. 22 Cheatham ML, Chapman WC, Key SP, Sawyers JL. A meta-analysis of selective versus routine nasogastric decompression after elective laparotomy. Ann Surg 1995;221:469. 23 Kehlet H. Acute pain control and accelerated postoperative surgical recovery. Surg Clin North Am 1999;79:431-43. 24 Kehlet H. Multimodal approach to control postoperative pathophysiology and rehabilitation. Br J Anaesth 1997;78:606-17. 25 Holte K, Kehlet H. Postoperative ileus: a preventable event. Br J Surg 2000;87:1480-93. 26 Power I, Barratt S. Analgesic agents for the postoperative period. Nonopioids. Surg Clin North Am 1999;79:275-97. 27 McQuay H, Moore A, Justins D. Treating acute pain in hospital. BMJ 1997:314:1531-5. 28 Watcha MF. The cost-effective management of postoperative nausea and vomiting. Anesthesiology 2000;92:931-3. 29 Strunin L, Rowbotham D, Miles A, eds. The effective management of postoperative nausea and vomiting. London: Aesculapius Medical Press, 1999:1-42. 30 Kehlet H, Rosenberg J. Surgical stress: pain, sleep and convalescence. In: Kinney JM, Tucker HN, eds. Physiology, stress and malnutrition: functional correlates, nutritional intervention. New York: Lippincott-Raven, 1997:95-112. 31 Spth-Schwalbe E, Hansen K, Schmidt F, Schrezenmeier H, Marchall L, Burger K, et al. Acute effects of recombinant human interleukin-6 on endocrine and central nervous sleep functions in healthy men. J Clin Endocrinol Metab 1998;83:1573-9. 32 Poldermans D, Boersma E, Bax JJ, Thomson IR, van de Ven LL, Blankensteijn JD, et al. The effect of bisoprolol on perioperative mortality and myocardial infarction in high-risk patients undergoing vascular surgery. N Engl J Med 1999;341:1789-94.

6 7 8

(Accepted 21 December 2000)

Changing stockings
Remember who commended thy yellow stockings, and wished to see thee ever cross-gartered: I say, remember. Go to, thou art made, if thou desirest to be so; if not, let me see thee a steward still, the fellow of servants, and not worthy to touch Fortunes fingers. Farewell. William Shakespeare, Twelfth Night I was not sure whether I had got it right. Yes, the patient seems to change stockings, the nurse repeated. I was on call and as a junior doctor I was becoming used to unusual requests. Being a foreigner made it worse, because often I did not even understand what they were talking about. I did not have any idea what the nurse wanted me to do with that information. All I could picture was Malvolios yellow cross-gartered stockings, but I still could not make any sense out of it. So I asked and the answer was frightening. It means the patient is going to die soon, she stated in a tone that indicated how she disapproved of my lack of common knowledge. And her prognosis was correct: I found the patient terribly ill. But what a strange way to characterise his final state, assuming that he was getting prepared (and dressed) for his own funeral. In Germany we call it biting the grass, crossing the River Jordan, or passing ones spoon, but that would not be considered adequate in such a situation. Was this the legendary black humour of which my father had warned me? What is it like to start practising medicine in a foreign language? one of our students asked me later. I thought that this incident was a good example of the unexpected difficulties you might encounter. Surprisingly he had never heard of the expression changing stockings. But on the other hand he was English and so also a foreigner in Scotland. To my list of pitfalls for beginners, I actually added the advice: If a nurse tells you a patient is changing stockings, get there right away. It took months until I finally discovered that people across the channel are not as cynical and strange as I had reported home: all the nurse had wanted to tell me was that our late patient was Cheyne-Stoke-ing. Martin Sielk general practitioner trainee, University of Dsseldorf

476

BMJ VOLUME 322

24 FEBRUARY 2001

bmj.com

1998 by International Anesthesia Research Society. Volume 86(4) April 1998 pp 837-844

Epidural Anesthesia and Gastrointestinal Motility [Review Article] Steinbrook, Richard A. MD Department of Anesthesia, Brigham and Women's Hospital and Harvard Medical School, Boston, Massachusetts. Accepted for publication December 30, 1997. Correspondence and reprint requests to Richard A. Steinbrook, MD, Department of Anesthesia, Brigham and Women's Hospital, 75 Francis St., Boston, MA 02115.

Postoperative ileus, a temporary inhibition of gastrointestinal function, is a universal complication after major abdominal surgery. Treatment for ileus is supportive and has changed little since Wangensteen's 1932 report [1] that nasogastric suction could delay or replace operative management of bowel obstruction, thereby reducing mortality. Gastric decompression, together with IV hydration and electrolyte replacement, remains the only proven therapy for ileus [2,3]. Liu et al. [4] suggest that epidural analgesia may significantly shorten the duration of postoperative ileus. The benefits of a reduction in ileus include decreased patient morbidity and potentially substantial cost-savings, as prolongation of hospitalization in the United States due to ileus has been estimated to cost $1,500 per patient or $750,000,000 annually [3]. Nevertheless, clinical guidelines currently promulgated by some consulting firms continue to state that "while epidural analgesia is effective for thoracic surgery and certain major musculoskeletal procedures, it has often been associated with prolonged ileus, delayed oral nutrition, and discharge in patients with gastrointestinal surgery" (Milliman and Robertson, Inc, Actuaries and Consultants, Seattle, WA, written communication, 1996). In this article, the pathophysiology of postoperative ileus is reviewed, and a framework for appreciating the theoretical basis for an effect of epidural anesthesia, especially thoracic epidural anesthesia, on ileus is provided. Potential risks and benefits of epidural anesthesia for bowel surgery are considered, including an examination of relevant animal studies. The major focus of this article is to review recent clinical studies comparing epidural analgesia with systemic analgesia, as well as to review studies comparing epidural narcotics with epidural local anesthetics with regard to postoperative ileus. Catheter location is discussed as a particularly important factor in determining the effects of epidural blockade on gastrointestinal motility. Finally, suggestions for future research are offered. Pathophysiology "On no subject in physiology do we meet with so many discrepancies of fact and opinion as in the physiology of the intestinal movements" [5]. Although intestinal motor activity may be normal after physical or chemical blockade of all neural input [6], contractile activity of the bowel is modulated by a variety of neural and humoral factors. Nearly 100 yr ago, Cannon and Murphy [7] demonstrated that opening the peritoneal cavity and manipulating the intestines resulted in a striking inhibition of contractile activity in the gastrointestinal tracts of dogs. The same authors also reported ileus associated with an extraabdominal procedure (crushing the testicles) in cats [8], whereas Meltzer and Auer [9] noted that ileus may follow various less noxious stimuli in rabbits. Parasympathetic stimulation increases gastrointestinal motility, but tonic inhibitory sympathetic control normally predominates. Thus, blockade of splanchnic nerves or spinal anesthesia results in increased motility or inhibits the

development of ileus, whereas vagotomy has little apparent effect. Although the autonomic nervous system has a major role in regulating gastrointestinal transit, other factors must also be involved. Factors that alter gastrointestinal motility in humans or animals are listed in Table 1.

Table 1. Factors that Alter Gastrointestinal Motility in Humans or Animals Typically, uncomplicated postoperative ileus is associated with restoration of motility in the stomach and small bowel within 24 h, whereas the colon recovers over 48-72 h [10,11]. Neely [12] suggested that the duration of postoperative ileus was related to the severity of the surgical procedure, but other authors' findings do not confirm this [13,14]. Other authors [3,15] use the terms paralytic or adynamic ileus to refer to more severe, prolonged inhibition of bowel function, as differentiated from the usual type of uncomplicated postoperative ileus that lasts no more than 3 days. The duration of postoperative ileus is increased by opioids [16]. The dose-dependent inhibitory effects of morphine and other opiates on motility [17] suggest a possible contributory role for endogenous opioids in the pathogenesis of postoperative ileus; however, the lack of effect of naloxone on postoperative bowel function in rats does not support this notion [18]. Inhaled anesthetics may decrease gastrointestinal motility, but motility consistently recovered within a matter of minutes after cessation of anesthesia in multiple animal studies [19]. Thus, it is unlikely that inhaled anesthetics are responsible for diminished gastrointestinal motility lasting much beyond the immediate postoperative period. Nitrous oxide may have longer-lasting deleterious effects on motility than do the volatile anesthetics. In a study of 40 patients undergoing elective major large bowel surgery under general anesthesia with isoflurane and fentanyl, Scheinin et al. [20] found significantly earlier return of bowel function, as assessed by the passage of flatus and feces, in the 20 patients randomly allocated to air compared with the 20 patients allocated to intraoperative nitrous oxide. The groups were comparable with respect to demographics and surgical procedures. The duration of postoperative hospitalization was significantly shorter for the air group (mean +/- SD; 10.0 +/- 1.3 vs 11.7 +/- 2.5 days; P < 0.05). IV infusion of lidocaine shortens the duration of postoperative ileus in humans [21]. In a double-blind study of patients undergoing cholecystectomy, the passage of radiopaque markers through the colon was significantly faster in the 15 patients who received IV lidocaine (100 mg bolus before anesthesia, continuous infusion at 3 mg/min for 24 h) than in the 15 patients who received IV saline. The authors speculate that systemic lidocaine may reduce postoperative peritoneal irritation, thereby suppressing inhibitory gastrointestinal reflexes; however, patients in the lidocaine group received significantly less postoperative narcotics, providing another explanation for the more rapid resolution of ileus.

Risks and Benefits Some authors have suggested that epidural anesthesia may be detrimental to the healing of a bowel anastomosis because of the increase in bowel motility. Carlstedt et al. [22] observed a significant increase in motility after the administration of atropine and neostigmine to reverse nondepolarizing muscle relaxants during epidural anesthesia; there was no increase in motility after atropine and neostigmine in the absence of epidural anesthesia. The authors warn that such an increase in intestinal motility "may expose a newly constructed colorectal anastomosis to undue strain in the immediate postoperative period" [22]. Despite the theoretical risk of increased motility secondary to anticholinesterase drugs during epidural anesthesia, disruption of colonic anastomoses during or immediately after epidural anesthesia has been reported in only three cases [23,24]; none involved neostigmine. Except for a statistically insignificant trend toward increased rates of anastomotic dehiscence in one study [25], there is substantial clinical and experimental evidence that epidural anesthesia/analgesia is safe for patients undergoing bowel resection with anastomosis. Furthermore, by increasing blood flow to the colon [26], epidural anesthesia with postoperative epidural analgesia may promote anastomotic healing. A study in animals by Schnitzler et al. [27] provides support for the safety of epidural analgesia after bowel anastomosis. After performing colorectal resection and anastomosis in 21 pigs, the authors administered epidural infusions of either bupivacaine, morphine, or saline for 48-72 h postoperatively. Colonic transit time, measured with radiopaque markers and serial radiographs, was accelerated with epidural bupivacaine (3.9 days) and epidural morphine (4 days) compared with epidural saline (6 days; P < 0.05). There were no significant differences in blood flow, intraluminal bursting pressure, or hydroxyproline content (a measure of wound healing), and there were no anastomotic complications. In another animal study, Udassin et al. [28] demonstrated beneficial effects of epidural anesthesia on ileus. These investigators measured the recovery of gastrointestinal motility in rats after a 30-min period of bowel ischemia. After a black test meal and a subsequent 90-min study period, they observed the fraction of the small bowel filled with the colored meal. Although control animals had 84.4% of the small bowel filled with the black meal, 30 min of ischemia resulted in pronounced adynamic ileus, with only 0.7% of the bowel filled with the marker. Lidocaine epidural anesthesia promoted the rapid resolution of ileus after ischemia, compared with epidural saline injection (lidocaine 60.3% filled, saline 30.9% filled). Aitkenhead et al. [29] reviewed the records of 68 patients who underwent large bowel anastomoses under spinal plus light general anesthesia, epidural plus light general anesthesia, or general anesthesia alone; postoperative analgesia was achieved with systemic narcotics. Early or late postoperative ileus (before or after the 4th postoperative day, respectively) occurred in 11.6% (early) and 11.6% (late) of patients in the spinal group, 12.0% and 4.0% of the epidural group, and 19.2% and 23.1% of the general anesthesia group. Anastomotic dehiscences occurred in 7.0% of patients in the spinal anesthesia group, 8.0% of patients in the epidural anesthesia group, and 23.1% of patients in the general anesthesia group. Although the differences observed in this retrospective study were not statistically significant, these investigators concluded that spinal or epidural anesthesia "may have had a beneficial effect on the anastomoses, since other factors were similar in the three groups." Consideration of the mechanisms and studies described above suggests a number of potentially desirable effects of epidural anesthesia on gastrointestinal motility (Table 2). By blocking thoracolumbar sympathetic nerves while leaving craniosacral parasympathetic nerves undiminished, epidural anesthesia-especially thoracic epidural anesthesia-would be expected to increase gastrointestinal motility. Furthermore, by substantially reducing or abolishing postoperative pain, epidural analgesia with a local anesthetic and/or narcotic decreases or eliminates the need for postoperative systemic opiates, thereby avoiding a major contributing factor to postoperative ileus. Additionally, to the extent that increased gastrointestinal blood flow [26] and systemic actions of local anesthetics [21] increase gastrointestinal motility, epidural analgesia may further reduce the duration of postoperative ileus.

Table 2. Mechanisms by Which Thoracic Epidural Anesthesia May Promote Gastrointestinal Motility Epidural Analgesia Compared with Systemic Analgesia Numerous recent studies have compared epidural analgesia and systemic analgesia with regard to the postoperative recovery of gastrointestinal function. Sixteen such studies published since 1977 are presented in Table 3, arranged in descending order by location of the epidural catheter. In all eight studies with epidural catheter placement above T12, gastrointestinal function recovered significantly more rapidly when epidural analgesia was used than when patients received systemic analgesics. Studies in which the epidural catheter was positioned at or below T12, or in which the location of the epidural catheter was not specified, were equally as likely to show faster recovery of gastrointestinal function with epidural analgesia as with systemic analgesia. In no case was systemic analgesia associated with more rapid recovery of gastrointestinal motility.

Table 3. Studies Comparing Postoperative Epidural Analgesia with Systemic Analgesia One of the first studies to compare the gastrointestinal effects of epidural local anesthetics with systemic narcotics was that of Gelman et al. [30]. These authors monitored intestinal motility by external electroenterography (EEnG) in 30 patients after cholecystectomy under general anesthesia. In 21 patients, an epidural catheter was placed at T7-8 and was intermittently dosed with bupivacaine during or after surgery. Electrical activity, as assessed by using EEnG, was decreased for 3-4 days after surgery. Eighty percent of the time, EEnG activity increased after epidural injections of bupivacaine, but EEnG activity almost always decreased after an IV or IM nicomorphine injection. In a randomized study of 214 patients undergoing major abdominal operations, Seeling et al. [31] compared patients receiving thoracic (T7-11) epidural plus light general anesthesia followed by postoperative thoracic epidural analgesia (bupivacaine

0.25% plus fentanyl 2 mg/mL, 6-10 mL/h for 76 +/- 1.45 h) with a control group receiving general anesthesia alone and postoperative IV or IM piritramide. The time to first feces was shorter in the epidural group (79 +/- 1.51 vs 93 +/- 1.38 h), but time to hospital discharge was the same. Although analgesia and ability to cough were better in the epidural group, the incidence and severity of postoperative complications were the same in both groups. In a study designed to compare postoperative pulmonary complications in patients after major abdominal surgery, Jayr et al. [32] randomly allocated 153 patients to receive either general anesthesia with postoperative subcutaneous morphine or combined thoracic (T7-11) epidural-general anesthesia with postoperative thoracic epidural analgesia (bupivacaine 0.125%, 10 mL/h with morphine 0.25 mg/h) for 4 days. Recovery of intestinal gas transit was significantly earlier in the epidural group, but the duration of hospitalization was not different. In a randomized, double-blind study of 30 morbidly obese patients undergoing gastroplasty, Rawal et al. [33] compared the effects of thoracic (T8) epidural morphine 4 mg with morphine 0.1 mg/kg IM given on demand. Postoperative analgesia was better with epidural morphine, at significantly smaller total doses of morphine. Bowel function, as assessed by first flatus or feces, recovered sooner with epidural morphine. Duration of hospitalization was significantly shorter with epidural morphine (7.1 +/- 0.3 days) than with IM morphine (9.0 +/- 0.6 days; P < 0.05). In a study by Bredtmann et al. [25], 116 patients undergoing colonic resection and/or anastomosis were randomly allocated to receive thoracic (T8-9 or T9-10) epidural plus general anesthesia or to receive general anesthesia alone. The groups were comparable with respect to preoperative morbidity, as well as to surgical procedures. Patients in the epidural group received bupivacaine 0.75% during surgery and 0.25% continuously for 3 days to maintain blockade of T5-L2; control patients received systemic narcotics and other analgesics. Epidural patients had significantly lower pain scores and earlier bowel movements. Nevertheless, the authors of this study noted several disadvantages of epidural analgesia, including significantly more fevers, as well as statistically insignificant trends toward higher rates of rectal anastomotic breakdown, blood replacement, intensive care therapy, and longer hospitalization. Liu and colleagues [34] randomized 54 patients undergoing partial colectomy into four groups. All patients had a standardized general anesthetic, as well as standardized postoperative care. One group received IV morphine analgesia; the other groups received thoracic (T8-10) epidural morphine, bupivacaine, or both. Groups were similar with respect to patient demographics, type and duration of surgery, and blood loss and fluid replacement. Time to first flatus and time to fulfillment of predetermined discharge criteria were significantly shorter for patients in the bupivacaine (flatus 40 +/- 2 h, discharge 62 +/- 5 h) and bupivacaine plus morphine (43 +/- 4 h, 67 +/- 8 h) epidural groups than for those in the epidural morphine (71 +/- 4 h, 102 +/13 h) and the IV morphine (81 +/- 3 h, 96 +/- 7 h) groups. Analgesia with activity was also significantly better in the two epidural bupivacaine groups. In a retrospective review of 68 women who underwent radical hysterectomies, de Leon-Casasola et al. [35] compared bowel function recovery with postoperative continuous thoracic (T10-12) epidural analgesia (bupivacaine 0.05% with morphine 0.01% for approximately 4 days) with IV morphine via patient-controlled analgesia (PCA). The epidural group required fewer days of nasogastric therapy (4 +/- 3 vs 8 +/- 2), had shorter times to first flatus (4 +/- 3 vs 8 +/- 2 days), tolerated solid foods sooner (6 +/- 2 vs 11 +/- 3 days), and had a shorter duration of hospitalization (10 +/- 3 vs 14 +/- 4 days) than the PCA group. Total hospital room costs were significantly less for epidural patients ($4175 vs $5845). Wallin et al. [36] studied 30 patients undergoing elective cholecystectomy under general anesthesia; in 15 patients, an epidural catheter was inserted preoperatively at T12-L1 and dosed with 0.5% plain bupivacaine (18-20 mL, level from T2-4 to S3-5); postoperative sensory blockade was maintained with intermittent injections of 0.25% bupivacaine (10-14 mL every 3 h) for 24 h in the epidural group, and by IM pentazocine in the general anesthesia alone group. Colonic motility was evaluated by the transit of radiopaque markers on serial abdominal radiographs, by time to first flatus, and by time to first feces. Despite effective epidural blockade in 11 patients (as evidenced by lower blood glucose concentrations for 24 h after skin incision), there were no significant differences in the passage of radiopaque markers or in times to first flatus or feces. However, it is possible that the lack of benefit in the epidural group was due to the relatively low insertion site (T12-L1), as well as the short duration of epidural analgesia (24 h), compared with the much longer mean transit times for radiopaque markers (>60 h).

In a randomized study of 40 patients undergoing abdominal hysterectomy, Wattwil [37] compared epidural analgesia with bupivacaine with IM ketobemidone (i.e., a synthetic opioid). All patients received general anesthesia, but those in the epidural group (T12-L1 catheter, 0.5% bupivacaine to achieve at least a T6 level before induction of general anesthesia) received no intraoperative opioids. Postoperative analgesia was maintained with 0.25% bupivacaine at 8 mL/h for 26-30 h in the epidural group, and with ketobemidone in the general anesthesia alone group. Pain relief was significantly better in the epidural group. Despite the low catheter position and short duration of epidural infusion, gastrointestinal motility, as assessed by times to first flatus and feces and by radiopaque markers and serial radiographs, was significantly enhanced in the epidural group. Hjortso et al. [38] randomized 100 patients scheduled for elective abdominal surgery to either general anesthesia with postoperative IM morphine (4-8 mg every 4-6 h), or combined lumbar (L1-2) epidural-general anesthesia with postoperative epidural analgesia. Epidural catheters were dosed preoperatively with sufficient 1.5% etidocaine to block T4-S5; postoperative epidural analgesia was achieved with 0.5% bupivacaine, 5 mL/4 h for 24 h, with morphine 4 mg/12 h for 72 h. Postoperative pain relief, assessed retrospectively on the 5th postoperative day, was better in the epidural group; nonetheless, there were no significant differences in a variety of postoperative complications. There also were no significant differences in recovery of bowel function as assessed by flatus, feces, and food intake. This is the only large, randomized, prospective study that has not found any advantage for epidural analgesia with regard to recovery of bowel function; however, the epidural insertion site was low (L1-2) and the dose of bupivacaine was small (total 30 mL over 24 h). Ahn et al. [39] randomly allocated 30 patients undergoing surgery of the left colon and/or rectum to postoperative lumbar (L23) epidural analgesia (0.25% bupivacaine, 8- to 15-mL boluses for 48 h) or to a control group receiving IV pentazocine. One hour after surgery, barium was injected into a duodenal tube; transit time was measured using serial radiographs. Compared with the control group, the epidural group had significantly shorter barium transit times through the small (12 vs 24 h for epidural versus control) and large intestines (35 vs 150 h), as well as significantly shorter times to first flatus and feces. Scheinin et al. [40] randomly allocated 60 patients undergoing colonic surgery to one of four groups with regard to postoperative pain control: 1) a control group receiving IM oxycodone on request; 2) an epidural group receiving an epidural bupivacaine infusion (0.25%, 4-6 mL/h for 48 h); 3) an epidural group receiving epidural morphine boluses (2-6 mg/d for 48 h); or 4) an epidural group receiving an epidural morphine infusion (2-6 mg/d for 48 h). Bowel movements occurred on the 2nd postoperative day in the epidural bupivacaine group, significantly earlier than all other groups (4th postoperative day). Lehman and Wiseman [41] reviewed the hospital courses of 102 patients who underwent elective colonic surgery. All patients received general anesthesia; 41 patients received postoperative epidural analgesia with narcotics alone or together with local anesthetic for an average of 3.4 +/- 1.4 days (range 1-7 days), whereas 61 patients received postoperative parenteral narcotics or ketorolac. There were no significant differences in duration of ileus or length of hospital stay in this retrospective study. The site of epidural catheterization was not identified. Morimoto et al. [42] reviewed the records of 85 patients who underwent proctocolectomy with ileal pouch-anal canal anastomosis at the Mayo Medical Center. Postoperative pain was treated with systemic morphine in 41 patients and with epidural fentanyl (bolus 1 [micro sign]g/kg, infusion 1 [micro sign]g [center dot] kg-1 [center dot] h-1 for 3.1 +/- 1.2 days) in 44 patients. Patients in the epidural group had shorter times to first feces (3.5 +/- 1.2 vs 4.3 +/- 1.3 days) and to first oral intake (4.5 +/- 0.9 vs 6.2 +/- 3.2 days) and had shorter total hospital stays (9.6 +/- 2.8 vs 12.1 +/- 4.4 days). Epidural patients also experienced significantly less need for nasogastric suction (61% vs 90% of patients, duration 1.9 +/- 1.7 vs 4.1 +/- 2.3 days) and IV fluids (6.6 +/- 2.0 vs 9.9 +/- 4.6 days), although the criteria used to determine need were not reported. Kanazi et al. [43] came to a different conclusion from reviewing the records of 50 patients who underwent colectomy with ileal pouch-anal anastomosis at the University of Nebraska Medical Center. All patients received general anesthesia; postoperative pain was managed with epidural medications (local anesthetic and narcotic for 24 h, then fentanyl or morphine) in 23 patients; 27 patients received parenteral analgesics only. Although the pain scores were significantly lower in the epidural group, there were no significant differences in duration of nasogastric suction (4.1 +/- 1.8 vs 4.6 +/- 4.1 days, epidural versus parenteral), number of patients requiring tube reinsertion, or time to tolerating liquid (4.8 +/- 1.9 vs 5.1 +/- 4.4 days) or regular (7.0 +/- 2.5 vs 7.7 +/- 5.3 days) diet. The mean hospital stay was 10.5 +/- 3.6 days for the epidural group, similar to the 12.6 +/- 6.9 days for the parenteral group.

In a retrospective comparison of thoracic epidural analgesia, lumbar epidural analgesia, and IV morphine via PCA, Scott et al. [44] observed the best pain control and fastest resolution of ileus in the thoracic epidural group. Patients undergoing restorative proctocolectomy under general anesthesia received intra-and postoperative analgesia with narcotics or local anesthetic-narcotic mixtures via either a thoracic (T6-10, n = 53) or lumbar (L2-4, n = 51) epidural catheter; a third group did not receive epidurals and had postoperative pain control with IV morphine via PCA (n = 75). Thoracic epidural catheters were infused for a longer period of time (3.7 +/- 1.8 days) than were lumbar catheters (2.0 +/- 1.2 days; P < 0.05), and smaller doses of morphine were used with thoracic catheters (0.25 mg/h) than with lumbar catheters (0.35 mg/h). Nevertheless, the pain scores (daily visual analog scale) were lowest in the thoracic epidural group. Bowel sounds returned 2.45 +/- 1.19 days postoperatively in the thoracic group, significantly earlier than in the lumbar (3.17 +/- 1.18 days) or PCA (2.96 +/- 1.14 days) groups (P < 0.05). Similarly, patients with thoracic catheters had stool outputs greater than 50 mL/8 h 3.4 +/- 1.7 days postoperatively versus 4.0 +/- 1.5 days postoperatively for patients with lumbar catheters and 4.3 +/- 1.3 days postoperatively for patients with PCA (P < 0.05). There were no significant differences in postoperative length of stay. This study provides direct evidence for the importance of catheter location in determining the effects of epidural analgesia on postoperative gastrointestinal motility. Epidural Local Anesthetics Compared with Epidural Narcotics Studies evaluating postoperative gastrointestinal function comparing epidural local anesthetics with epidural narcotics are presented in Table 4. In all studies with epidural catheter placement above T12, gastrointestinal motility was greater with the use of epidural local anesthetics compared with epidural narcotics.

Table 4. Studies Comparing Epidural Local Anesthetics with Epidural Narcotics In a study in healthy volunteers, Thoren and Wattwil [45] compared acetaminophen absorption, a measure of the rate of gastric emptying, after thoracic (T4) epidural injection of either 4 mg of morphine or 0.5% bupivacaine (sufficient to block at least T610). Compared with control acetaminophen absorption studies without epidural injection, epidural analgesia with morphine significantly delayed gastric emptying (lower mean and maximal serum acetaminophen concentrations, longer time to peak concentration, smaller area under concentration-time curve), whereas acetaminophen absorption after epidural bupivacaine was the same as that after control. In a small, randomized study in 14 patients after open cholecystectomy, Thorn et al. [46] compared gastroduodenal myoelectric activity and acetaminophen absorption during thoracic epidural analgesia with bupivacaine (0.25% at 8.0 +/- 0.9 mL/h) or morphine (4 mg plus 2 mg as needed). Pain relief (visual analog scale) at rest was the same in both groups. Acetaminophen

absorption was significantly delayed in the epidural morphine group. Furthermore, epidural morphine was associated with significant changes in gastroduodenal myoelectric activity compared with epidural bupivacaine. Thoren et al. [47] compared low thoracic (T12-L1) epidural bupivacaine (0.5% intraoperatively, 0.25% postoperatively for 42 h) with epidural morphine (4 mg, then 2 mg bolus as needed up to 42 h) in 22 patients undergoing abdominal hysterectomies under general anesthesia. The epidural bupivacaine patients had significantly better pain relief, earlier flatus (22 +/- 16 vs 56 +/- 22 h, P < 0.001), earlier feces (57 +/- 44 vs 92 +/- 22 h, P < 0.05), and earlier and greater intake of oral fluids. In a randomized study of 29 patients undergoing elective major abdominal surgery, Bisgaard et al. [48] compared lumbar (L24) epidural analgesia with bupivacaine plus morphine as a continuous infusion for 3-6 days with epidural morphine boluses for 48 h. Although pain relief was better with the combination of bupivacaine plus morphine, there were no differences in colonic motility, as assessed by first flatus, first feces, and radiopaque markers. Conclusions Thoracic epidural anesthesia with postoperative thoracic epidural analgesia has been shown to have beneficial effects on postoperative pain and recovery of bowel function after major abdominal surgery; lumbar epidural blockade is not as consistently effective. Local anesthetics and local anesthetic-narcotic mixtures seem to be more effective with fewer undesirable side effects than epidural narcotics alone; however, published studies are limited by relatively small numbers of subjects, as well as by lack of documentation of the level of epidural blockade or degree of analgesia or sympathectomy. Future studies should include documentation of the level of epidural blockade, and might include measurements of intestinal blood flow and motility. Additional studies are required to determine the ideal drugs for epidural infusion, optimal timing of administration (i.e., when to start, as well as how long to continue), and differences, if any, in outcome measures such as patient satisfaction and time to return to work. Simon Gelman, MD, PhD, provided a critical review of the manuscript. REFERENCES 1. Wangensteen OH. The early diagnosis of acute intestinal obstruction with comments on pathology and treatment: with a report on successful decompression of three cases of mechanical small bowel obstruction by nasal catheter siphonage. West J Surg Obstet Gynecol 1932;40:1-17. [Context Link] 2. Heimbach DM, Crout JR. Treatment of paralytic ileus with adrenergic neuronal blocking drugs. Surgery 1971;69:582-7. [Medline Link] [Context Link] 3. Livingston EH, Passaro EP. Postoperative ileus. Dig Dis Sci 1990;35:121-32. [Medline Link] [Context Link] 4. Liu S, Carpenter RL, Neal JM. Epidural anesthesia and analgesia: their role in postoperative outcome. Anesthesiology 1995;82:1474-506. [Fulltext Link] [Medline Link] [Context Link] 5. Bayliss WM, Starling EH. The movements and innervation of the small intestine. J Physiol (Lond) 1899;24:99-143. [Context Link] 6. Ochsner A, Gage IM, Cutting RA. Comparative value of splanchnic and spinal analgesia in the treatment of experimental ileus. Arch Surg 1930;20:802-31. [Context Link] 7. Cannon WB, Murphy FT. The movement of the stomach and intestine in some surgical conditions. Ann Surg 1906;43:51236. [Context Link] 8. Cannon WB, Murphy FT. Physiologic observations on experimentally produced ileus. JAMA 1907;49:840-3. [Context Link]

9. Meltzer SJ, Auer J. Peristaltic movements of the rabbit's cecum and their inhibition, with demonstration. Proc Soc Exp Biol Med 1907;4:37-40. [Context Link] 10. Wells C, Rawlinson K, Tinckler L, et al. Ileus and postoperative intestinal motility. Lancet 1961;2:136-7. [Context Link] 11. Woods JH, Erickson LW, Condon RE, et al. Postoperative ileus: a colonic problem? Surgery 1978;84:527-33. [Medline Link] [Context Link] 12. Neely J. The effects of analgesic drugs on gastrointestinal motility in man. Br J Surg 1969;56:925-9. [Medline Link] [Context Link] 13. Graber JN, Schulte WJ, Condon RE, Cowles VE. Relationship of duration of postoperative ileus to extent and site of operative dissection. Surgery 1982;92:87-92. [Medline Link] [Context Link] 14. Wilson JP. Postoperative motility of the large intestine in man. Gut 1975;16:689-92. [Medline Link] [Context Link] 15. Cathpole BN. Ileus: use of sympathetic blocking agents in its treatment. Surgery 1969;66:811-20. [Medline Link] [Context Link] 16. Yukioka H, Bogod DG, Rosen M. Recovery of bowel motility after surgery: detection of time of first flatus from carbon dioxide concentration and patient estimate after nalbuphine and placebo. Br J Anaesth 1987;59:581-4. [Medline Link] [Context Link] 17. Weisbrodt NW, Sussman SE, Stewart JJ, Burks TF. Effect of morphine sulfate on intestinal transit and myoelectric activity of the small intestine of the rat. J Pharmacol Exp Ther 1980;214:333-8. [Medline Link] [Context Link] 18. Howd RA, Adamovics J, Palekar A. Naloxone and intestinal motility. Experientia 1978;34:1310-1. [Medline Link] [Context Link] 19. Condon RE, Cowles V, Ekbom GA, et al. Effects of halothane, enflurane and nitrous oxide on colon motility. Surgery 1987;101:81-5. [Medline Link] [Context Link] 20. Scheinin B, Lindgren L, Scheinin TM. Peroperative nitrous oxide delays bowel function after colonic surgery. Br J Anaesth 1990;64:154-8. [Medline Link] [Context Link] 21. Rimback G, Cassuto J, Tollesson P-O. Treatment of postoperative paralytic ileus by intravenous lidocaine infusion. Anesth Analg 1990;70:414-9. [Medline Link] [Context Link] 22. Carlstedt A, Nordgren S, Fasth S, et al. Epidural anaesthesia and postoperative colorectal motility: a possible hazard to a colorectal anastomosis. Int J Colorect Dis 1989;4:144-9. [Medline Link] [Context Link] 23. Treissman D. Disruption of colonic anastomosis associated with epidural anesthesia. Reg Anesth 1980;5:22-3. [Context Link] 24. Bigler D, Hjortso MC, Kehlet H. Disruption of colonic anastomosis during continuous epidural analgesia: an early postoperative complication. Anaesthesia 1985;40:278-80. [Medline Link] [Context Link] 25. Bredtmann RD, Herden HN, Teichmann W, et al. Epidural analgesia in colonic surgery: results of a randomized prospective study. Br J Surg 1990;77:638-42. [Medline Link] [Context Link] 26. Johansson K, Ahn H, Lindhagen J, Tryselius U. Effect of epidural anaesthesia on intestinal blood flow. Br J Surg 1988;75:73-6. [Medline Link] [Context Link]

27. Schnitzler M, Kilbride M, Senagore A. Effect of epidural analgesia on colorectal anastomotic healing and colonic motility. Reg Anesth 1992;17:143-7. [Medline Link] [Context Link] 28. Udassin R, Eimerl D, Schiffman J, Haskel Y. Epidural anesthesia accelerates the recovery of postischemic bowel motility in the rat. Anesthesiology 1994;80:832-6. [Medline Link] [Context Link] 29. Aitkenhead AR, Wishart HY, Peebles Brown DA. High spinal nerve block for large bowel anastomosis. Br J Anaesth 1978;50:177-83. [Medline Link] [Context Link] 30. Gelman S, Feigenberg Z, Dintzman M, Levy E. Electroenterography after cholecystectomy: the role of high epidural analgesia. Arch Surg 1977;112:580-3. [Medline Link] [Context Link] 31. Seeling W, Bruckmooser K-P, Hufner C, et al. Continuous thoracic epidural analgesia does not diminish postoperative complications after abdominal surgery in patients at risk. Anaesthesist 1990;39:33-40. [Medline Link] [Context Link] 32. Jayr C, Thomas H, Rey A, et al. Postoperative pulmonary complications: epidural analgesia using bupivacaine and opioids versus parenteral opioids. Anesthesiology 1993;78:666-76. [Medline Link] [Context Link] 33. Rawal N, Sjostrand U, Christoffersson E, et al. Comparison of intramuscular and epidural morphine for postoperative analgesia in the grossly obese: influence on postoperative ambulation and pulmonary function. Anesth Analg 1984;63:583-92. [Medline Link] [Context Link] 34. Liu SS, Carpenter RL, Mackey DC, et al. Effects of perioperative analgesic technique on rate of recovery after colon surgery. Anesthesiology 1995;83:757-65. [Fulltext Link] [Medline Link] [Context Link] 35. de Leon-Casasola OA, Karabella D, Lema MJ. Bowel function recovery after radical hysterectomies: thoracic epidural bupivacaine-morphine versus intravenous patient-controlled analgesia with morphine-a pilot study. J Clin Anesth 1996;8:8792. [Medline Link] [Context Link] 36. Wallin G, Cassuto J, Hogstrom S, et al. Failure of epidural anesthesia to prevent postoperative paralytic ileus. Anesthesiology 1986;65:292-7. [Medline Link] [Context Link] 37. Wattwil M, Thoren T, Hennerdal S, Garvill J-E. Epidural analgesia with bupivacaine reduces postoperative paralytic ileus after hysterectomy. Anesth Analg 1998;68:353-8. [Medline Link] [Context Link] 38. Hjortso NC, Neumann P, Frosig F, et al. A controlled study on the effect of epidural analgesia with local anaesthetics and morphine on morbidity after abdominal surgery. Acta Anaesthesiol Scand 1985;29:790-6. [Medline Link] [Context Link] 39. Ahn H, Bronge A, Johansson K, et al. Effect of continuous postoperative epidural analgesia on intestinal motility. Br J Surg 1988;75:1176-8. [Medline Link] [Context Link] 40. Scheinin B, Asantila R, Orko R. The effect of bupivacaine and morphine on pain and bowel function after colonic surgery. Acta Anaesthesiol Scand 1987;31:161-4. [Medline Link] [Context Link] 41. Lehman J, Wiseman J. The effect of epidural analgesia on the return of peristalsis and the length of stay after elective colonic surgery. Am Surg 1995;61:1009-12. [Medline Link] [Context Link] 42. Morimoto H, Cullen JJ, Messick JM Jr, Kelly KA. Epidural analgesia shortens postoperative ileus after ileal pouch-anal canal anastomosis. Am J Surg 1995;169:79-83. [Fulltext Link] [Medline Link] [Context Link] 43. Kanazi GE, Thompson JS, Boskouski NA. Effect of epidural analgesia on postoperative ileus after ileal pouch-anal anastomosis. Am Surg 1996;62:499-502. [Medline Link] [Context Link]

44. Scott AM, Starling JR, Ruscher AE, et al. Thoracic versus lumbar epidural anesthesia's effect on pain control and ileus resolution after restorative proctocolectomy. Surgery 1996;120:688-97. [Medline Link] [Context Link] 45. Thoren T, Wattwil M. Effects on gastric emptying of thoracic epidural analgesia with morphine or bupivacaine. Anesth Analg 1988;67:687-94. [Medline Link] [Context Link] 46. Thorn S-E, Wickbom G, Philipson L, et al. Myoelectric activity in the stomach and duodenum after epidural administration of morphine of bupivacaine. Acta Anaesthesiol Scand 1996;40:773-8. [Medline Link] [Context Link] 47. Thoren T, Sundberg A, Wattwil M, et al. Effects of epidural bupivacaine and epidural morphine on bowel function and pain after hysterectomy. Acta Anaesthesiol Scand 1989;33:181-5. [Medline Link] [Context Link] 48. Bisgaard C, Mouridsen P, Dahl JB. Continuous lumbar epidural bupivacaine plus morphine versus epidural morphine after major abdominal surgery. Eur J Anaesthesiol 1990;7:219-25. [Context Link]

Accession Number: 00000539-199804000-00029 Copyright (c) 2000-2001 Ovid Technologies, Inc. Version: rel4.3.0, SourceID: 1.5031.1.149

British Journal of Surgery 2000 Blackwell Science Ltd. Volume 87(11) November 2000 pp 1480-1493

Postoperative ileus: a preventable event [Review] Holte, K.; Kehlet, H. Department of Surgical Gastroenterology, Hvidovre University Hospital, DK-2650 Hvidovre, Denmark Correspondence to: Dr H. Kehlet Paper accepted 14 June 2000 Abstract Background: Postoperative ileus has traditionally been accepted as a normal response to tissue injury. No data support any beneficial effect of ileus and indeed it may contribute to delayed recovery and prolonged hospital stay. Efforts should, therefore, be made to reduce such ileus. Methods: Material was identified from a Medline search of the literature, previous review articles and references cited in original papers. This paper updates knowledge on the pathophysiology and treatment of postoperative ileus. Results and conclusion: Pathogenesis mainly involves inhibitory neural reflexes and inflammatory mediators released from the site of injury. The most effective method of reducing ileus is thoracic epidural blockade with local anaesthetic. Opioid-sparing analgesic techniques and non-steroidal anti-inflammatory agents also reduce ileus, as does laparoscopic surgery. Of the prokinetic agents only cisapride is proven beneficial; the effect of early enteral feeding remains unclear. However, postoperative ileus may be greatly reduced when all of the above are combined in a multimodal rehabilitation strategy.

Introduction Postoperative ileus is generally defined as a transient impairment of bowel motility after abdominal surgery or other injury. Clinically, it is characterized by bowel distension, lack of bowel sounds, and lack of passage of flatus and stool. Symptoms include nausea, vomiting and stomach cramps, and ileus is thus a major contributory factor to postoperative discomfort. Resumption of a regular diet and mobilization is delayed, and hospital stay is thereby prolonged. Ileus has traditionally been accepted as an obligatory physiological response to abdominal surgery, but the purpose of this response in the elective surgical setting has not been established and no data suggest a beneficial effect of the delayed postoperative recovery of gastrointestinal motility. Various pathogenetic mechanisms have been proposed and several pharmacological interventions have been employed to resolve ileus, but so far no single technique or agent has been found effectively to eliminate the problem. In recent reviews 1-5 both pathophysiology and pharmacological treatment have been described, but none of these papers has addressed recent advances in perioperative management (including choice of anaesthesia, pain management, nutrition and mobilization) in an integrated multimodal approach. The present review updates knowledge on the pathophysiology and pharmacological treatment of postoperative ileus, and brings these data into the broader context of a multimodal anaesthetic, analgesic, pharmacological and nutritional rehabilitation strategy in an attempt to reduce ileus and improve postoperative outcome. Definitions and methods of assessment

Not all segments in the gastrointestinal tract are equally affected by postoperative ileus. The average paralytic state lasts between 0 and 24 h in the small intestine, 24 and 48 h in the stomach, and between 48 and 72 h in the colon after major abdominal surgery. The effective duration of ileus is, therefore, mainly dependent on the return of colonic motility and, in particular, motility of the left colon 6,7. Ileus most commonly occurs after intraperitoneal operations, but it may also occur after retroperitoneal and extra-abdominal surgery. The duration is related to the anatomical location of surgery and the longest duration is encountered after operation involving the colon 8,9. The difference between right-sided and left-sided colonic procedures on the duration of ileus remains uncertain 10-12. In experimental and clinical studies ileus has been demonstrated to be related to the degree of surgical manipulation and the magnitude of the inflammatory response 13. The definition of ileus and methods of assessment are not well defined. As an objective indicator of resolution, assessment of electrical activity has been widely used, focusing on either the return of the migrating myoelectric complex (MMC) or qualitative changes in MMC patterns. However, the MMC reflects mostly fasted state activity and some investigators have found no correlation between return of the MMC or specific MMC pattern and the clinical resolution of postoperative ileus 6,14. A correlation between some of the widely used clinical endpoints, such as bowel sounds, passage of flatus and stool, is also controversial. Bowel sounds are non-specific because they may originate in the small bowel as well as in the large bowel, and also require frequent auscultation for assessment. Passage of flatus is highly dependent on reporting by patients, and the correlation between passage of flatus and propulsive bowel movements is unclear 6. Passage of stool, although manifest as a clinical sign, is not specific, as it may indicate only distal bowel emptying and not necessarily the function of the entire gastrointestinal tract. Other frequently used endpoints include measurements of intraluminal pressure, migration of radio-opaque markers and non-invasive electrical measurements, such as percutaneous registration of electrical activity. Several investigators have found the clinical resolution of ileus to be relatively independent of these technical variables 15,16. As no single objective variable has yet been found accurately to predict resolution of ileus, the most adequate definition of resolution probably depends on a combined functional outcome of normalization of food intake and bowel function. Pathogenesis of postoperative ileus Inhibitory neural reflexes Several inhibitory reflexes in the gastrointestinal tract have been proposed to mediate postoperative ileus, including somatovisceral and viscerovisceral reflexes 1. Three anatomically distinguishable reflexes seem to be involved: ultrashort reflexes confined to the wall of the gut, short reflexes involving the prevertebral ganglia, and long reflexes involving the spinal cord. The long reflexes are probably of most importance, since several experimental studies have shown spinal anaesthesia, abdominal sympathectomy and other nerve-cutting techniques to prevent or reduce the development of ileus 2,3. Other experimental studies have demonstrated that selective degradation of splanchnic afferent neurones with capsaicin reduces ileus 17,18. This applies to both systemic administration and direct application at the prevertebral ganglia; perivagal administration is without effect 17. Furthermore, ablation of vagal fibres may not influence gastrointestinal transit after injury in contrast to spinal afferent fibre ablation 17. Several experimental studies suggest that the afferent reflexes originate primarily from the peritoneum and that skin incisions alone, unlike incisions through the peritoneum, do not provoke ileus 19,20. In summary, inhibitory sympathetic reflexes are of major importance in the pathogenesis of ileus. This has substantial clinical implications as these reflexes are subject to modification by epidural blockade. Neurotransmitters and inflammatory factors A surgical operation elicits a stress response that is generally considered to be of combined endocrine and inflammatory origin. Although many neurotransmitters and neuropeptides are found locally in the gastrointestinal tract and might possibly contribute to ileus, they may also be released systemically following noxious stimuli or local inflammatory responses (the wound). So far, few comparative studies are available, but plasma changes in motilin and substance P may be related to depressed postoperative gastrointestinal motility 21. These findings, however, do not exclude the importance of local release of these substances. Numerous transmitters and peptides are involved in regulating gastrointestinal motility and so may be involved in ileus. Nitric oxide, vasoactive intestinal peptide (VIP) and substance P have been established as inhibitory neurotransmitters in the intrinsic gut nervous system. Experimental studies have shown that VIP and substance P receptor antagonists, as well as inhibitors of

nitric oxide synthesis, improve postoperative gastrointestinal transit 19,22,23. Induced endotoxaemia in dogs leads to increased concentrations of products of VIP and nitric oxide synthesis combined with decreased gastrointestinal motility 24,25. Calcitonin gene-related peptide also inhibits postoperative gastric emptying and gastrointestinal transit, acting on specific peripheral receptors, possibly located in splanchnic afferent nerves or ganglia 17,26. Several studies have demonstrated that corticotrophin releasing factor (CRF) is involved in the pathogenesis of ileus, as intracisternal and intraventricular injection of CRF may delay postoperative gastrointestinal transit, which subsequently may be reversed by administration of specific CRF antagonists 27,28. Finally, opioids are well established as modulators of transmission in the central and peripheral nervous systems, leading to inhibition of gastric emptying and non-propulsive smooth muscle contraction with an increase in intraluminal pressure throughout the gastrointestinal tract 29,30. This effect is predominantly mediated by mu receptor agonists. Recent experimental studies have shown that selective peripheral kappa agonists may reverse ileus and gut paralysis following surgery or chemical peritonitis 31-33. Interestingly, experimental studies have also documented improved visceral pain relief by kappa agonists 32. Little work has focused on the role of local inflammatory responses and cellular events in mediating ileus. Bauer's group has demonstrated, in a series of experimental studies, that the local inflammatory response is related to the extent of surgical trauma and degree of ileus 13. The effect may be mediated by leucocyte-derived nitric oxide 23, and is prevented by reducing the number of inflammatory cells locally 34. Furthermore, the paralytic gut response to surgery seems to be biphasic, consisting of a short temporary initial paralysis, followed by a longer-lasting impairment of muscle activity paralleling the local tissue concentration of inflammatory cells 35. In summary, various neurotransmitters and inflammatory factors are known to be involved in the pathogenesis of postoperative ileus. However, their relative roles, as well as the hierarchical order and cooperation of these substances in the initiation and resolution of ileus, remain obscure. Perioperative management Anaesthesia All anaesthetics used for induction or maintenance of general anaesthesia may depress gastrointestinal motility 36. However, the choice of general anaesthetic technique may have insignificant effects on ileus as even prolonged general anaesthesia in surface surgery does not lead to any clinically relevant reduction in bowel motility. The potential contributory effect of nitrous oxide to ileus has been evaluated in several clinical trials comparing nitrous oxide with isoflurane or propofol in abdominal surgery; it has not been found to be of clinical significance 37-39. A single dose of neural blockade with spinal or epidural anaesthetic alone or as a supplement to general anaesthesia does not influence the duration of ileus 40. The role of intraoperative short-acting opioids (alfentanil, remifentanil) on ileus is unknown, but is unlikely to be of clinical significance. Analgesia As noted above, it is well established that a surgical noxious stimulus leads to activation of inhibitory sympathetic splanchnic reflexes, and that the choice of analgesic technique may affect several aspects of the surgical stress response, including these inhibitory reflexes 40. After major abdominal surgery effective dynamic pain relief may be obtained only by a continuous epidural infusion that includes local anaesthetics 41. Analgesic treatment that includes opioids may prolong ileus, and the use of opioid-sparing analgesia with non-steroidal anti-inflammatory drugs (NSAIDs) or other analgesics (balanced analgesia) may reduce ileus. Epidural analgesia Theoretically, epidural blockade with local anaesthetics may improve postoperative ileus by several mechanisms: blockade of afferent and efferent inhibitory reflexes, efferent sympathetic blockade with concomitant increase in splanchnic blood flow, and anti-inflammatory effects via systemic absorption of local anaesthetics 42,43. Several randomized studies in patients undergoing abdominal procedures have evaluated the effect of epidural thoracic local anaesthetics compared with systemic opioids 44-51 (Fig. 1). Thus, in six of eight studies epidural bupivacaine significantly reduced ileus. In one statistically negative study 44 the duration of epidural analgesia was only 24 h, whereas in all other studies it was administered for between 48 and 72 h. In the other small, statistically negative study 51 a low thoracic (Th9-12) epidural blockade was used, which may not permit sufficient dermatomal blockade of noxious stimuli to improve gastrointestinal motility.

Fig. 1 Randomized clinical trials assessing the effect of epidural local anaesthetic versus systemic opioid on postoperative ileus. In all studies colonic surgery was performed, except Wallin et al.44 (cholecystectomy) and Wattwil et al.47 (gynaecological surgery). The various endpoints (defaecation (D), combination score (C) and flatus (F)) used to assess resolution of postoperative ileus are indicated. The combination score is defined as a combination of flatus and defaecation. In studies with assessments of both flatus and defaecation, defaecation was given a higher priority than flatus. *P<0.05 Compared with epidural opioid or epidural bupivacaine plus low-dose opioid, epidural bupivacaine also led to a reduction in the duration of ileus 45,50,52,53 (Fig. 2). Furthermore, the combination of epidural opioid plus low-dose bupivacaine may also reduce ileus compared with epidural or systemic opioid 50,53-58 (Fig. 3). In one randomized study in hip surgery 59, the combination of intraoperative plus postoperative lumbar epidural local anaesthetics reduced ileus compared with general anaesthesia plus postoperative intravenously administered analgesia.

Fig. 2 Randomized clinical trials assessing the effect of epidural local anaesthetic versus epidural opioid or epidural local anaesthetic plus opioid on ileus. Two studies 45,50 were colonic and two studies 52,53 were gynaecological. The various endpoints (combination score (C), defaecation (D) and flatus (F)) used to assess resolution of ileus are indicated. The combination score is defined as a combination of flatus and defaecation. In studies with assessments of both flatus and defaecation, defaecation was given a higher priority than flatus. *P < 0.05

Fig. 3 Randomized clinical trials assessing the effect of a combination of epidural local anaesthetic and opioid versus epidural or

systemic opioid on ileus. In the studies demonstrating a positive effect of epidural local anaesthetic, thoracic epidural infusion was used. In the remaining four studies, epidural blockade was lumbar or not specified precisely. In all studies except that of Asantila et al.53 (hysterectomy), major abdominal (not specified) and/or colonic procedures were performed. The various endpoints (defaecation (D), combination score (C) and flatus (F)) used to assess resolution of ileus are indicated. The combination score is defined as a combination of flatus and defaecation. In studies with assessments of both flatus and defaecation, defaecation was given a higher priority than flatus. *P < 0.05 The established positive effect of epidural local anaesthetic administration is related to the segmental visceral afferent/efferent blockade which, in abdominal surgery, can be obtained only by thoracic application of the local anaesthetic. Not surprisingly, studies using lumbar or low-thoracic epidural administration of local anaesthetics have not demonstrated the positive effects of epidural analgesia on ileus 44,51,54,55. Experimental studies have shown that local anaesthetics inhibit the development of chemical peritonitis following hydrochloric acid administration 60 and shorten bowel paralysis after ischaemia 61. However, clinical studies using intraperitoneal or systemic (intravenous) administration of local anaesthetics have been inconclusive with regard to both analgesic effects 62 and to any potentially advantageous effect on the resolution of postoperative ileus 16,62,63. Opioid analgesia It is well known from experimental studies that opioids have a profound inhibitory effect on resting and post-traumatic gastrointestinal motility 30,31. These effects are primarily seen during systemic opioid administration with intravenous patientcontrolled analgesia, conventional intramuscular opioid administration or epidural opioid administration 11,29. Pain relief apart, there is a major difference between the effects on ileus of these opioid administration techniques and techniques involving local anaesthetics. Findings from experimental studies demonstrate that kappa opioid agonists may provide sufficient visceral analgesia 32 as well as reduce ileus 31,33, but this has not been studied in the clinical setting. The effects of a kappa opioid agonist (fedotozine) on pain and bowel function in a nonsurgical setting, irritable bowel syndrome, have been negative 64. Recent studies in chronic methadone users suggest that peripheral opioid antagonists (methylnaltrexone) may reverse the opioid-mediated inhibition of gastrointestinal function 65, with possible future implications in the surgical setting. Opioid-sparing analgesia Based on knowledge of the inhibitory effects of opioids on gut motility, various opioid-sparing analgesic techniques have been developed to avoid the undesirable sequelae of opioid administration in the postoperative period (drowsiness, nausea and vomiting, inhibition of bladder function)66. One of the most established techniques of obtaining a 20-30 per cent sparing of opioid is pain relief with NSAIDs 67. In both experimental and clinical studies, the administration of NSAIDs and the subsequent opioid-sparing effect has, in most but not all studies, resulted in less nausea and vomiting and improvement in overall gastrointestinal motility 15,20,68-70. Similar effects have been obtained with different types of NSAIDs, although no data are available to compare cyclo-oxygenase (COX) 1 versus COX-2 inhibitors. The advantageous effect of NSAIDs, apart from the sparing of opioid, may also be related to a direct anti-inflammatory effect mediated by the inhibition of prostaglandin synthesis 68. The effect of other types of opioid-sparing analgesia (ketamine, paracetamol, tramadol, etc.) has not been finally established. Nasogastric intubation The insertion of a nasogastric tube has been the traditional supportive treatment for postoperative ileus, but it does not shorten time to first bowel movement or time to effective oral food intake 71,72. Many randomized clinical studies and a meta-analysis have concluded that a nasogastric tube should not be used routinely, and that unnecessary use may contribute to postoperative morbidity such as atelectasis, pneumonia and fever 71,72. Mobilization

Contrary to popular belief, physical exercise does not improve colonic motility in healthy volunteers 73. Neither does mobilization of itself shorten the duration of postoperative ileus; recovery of myoelectrical activity is similar in patients mobilized from day 1 compared with day 4 after major abdominal surgery 74. However, the commonly practised prolonged immobilization after surgery has never been proven to be beneficial 75. In fact, prolonged bedrest may enhance the risk of postoperative complications and prolong recovery 41,76, and so should be avoided for reasons other than recovery of gastrointestinal function. Early postoperative oral feeding Food intake elicits a reflex response that is propulsive in action. Several intestinointestinal reflexes connecting various parts of the gastrointestinal tract respond to food intake, producing coordinated propulsive activity 77. In addition, the presence of food stimulates the secretion of various intestinal hormones, with an overall stimulating effect on gastrointestinal motility 77. Traditionally, oral feeding has been delayed after laparotomy until any ileus has resolved clinically. At that point a liquid diet is administered, gradually progressing to solid food. Patients undergoing major abdominal procedures often have at least 4 or 5 days of semistarvation or total starvation. This is unfortunate because semistarvation contributes to catabolism and so to fatigue and prolonged convalescence 41. Furthermore, several randomized clinical studies have shown that early enteral nutrition may improve immune function and reduce postoperative and posttraumatic infectious complications 78. Early uncontrolled studies by Moss 79 found that ileus was resolved within 48 h when patients received immediate enteral nutrition. Unfortunately, it has been only recently that the traditional restriction of early post-operative oral intake has been questioned seriously; several trials have shown the institution of early enteral feeding both to be safe 80-86 and to reduce postoperative ileus (Fig. 4). One study of 80 patients undergoing colorectal surgery and randomized to either early enteral feeding or conventional fasting showed that the early fed group tolerated a regular diet 3 days before the later fed group. The former were discharged from hospital 2 days before the traditionally fed group 88. Another study of 96 gynaecological patients demonstrated that an earlier fed group tolerated solid diet 1 day before a group receiving traditional feeding; the former were discharged 1 day sooner too 87. Two other studies involving gynaecological patients have similarly demonstrated a reduction in ileus in early fed groups 89,90, although in one of these studies 89 the positive effect was noted in only some of the assessed variables (return of bowel sounds and hospital stay).

Fig. 4 Randomized clinical trials assessing the effect of early enteral nutrition versus traditional feeding (consisting of no oral intake until clinical resolution of ileus or of only liquid oral intake) on resolution of ileus, except Heslin et al.91, which is

discussed in the text. Three studies 87,89,90 were performed in gynaecological patients and the rest in abdominal (not specified or colonic) surgery. As indicators of ileus various endpoints (combination score (C), time to tolerate regular food (I), passage of flatus (F) and defaecation (D)) were used. The combination score is defined as a combination of flatus and defaecation. In studies with assessments of both flatus and defaecation, defaecation was given a higher priority than flatus. *P < 0.05 In contrast, two other recent trials 91,92 investigating the effect of early oral feeding have demonstrated no difference in either length of ileus or hospital stay. In one of these studies impairment of pulmonary function and postoperative mobilization with early oral feeding was noted 92. Of importance, however, in these and other oral feeding studies is that the type of analgesia was rarely reported and, when mentioned, usually consisted solely of opioids and rarely of epidural local anaesthetics 80-85. Thus the established stimulatory effect of solid food on intestinal motility may not have been obvious as the profound inhibitory effect of opioids on motility may have dominated. The proven beneficial effects of continuous epidural local anaesthetics, opioid-sparing analgesics and cisapride have unfortunately not been incorporated in previous controlled clinical studies of early enteral nutrition. Future well designed studies are required to evaluate the potential benefits of early enteral nutrition when combined with regimens aimed at early resolution of postoperative ileus. Laparoscopic surgery Experimental studies comparing laparoscopic with open cholecystectomy have shown that the laparoscopic approach improves gastric emptying and fasted state gastrointestinal motility 14,93. However, ileus is clinically non-existent after laparoscopic cholecystectomy 94 and so these results are not applicable to major abdominal procedures. In experimental studies in colonic surgery, the laparoscopic technique leads to earlier return of gastrointestinal motility as well as earlier normalization of myoelectrical activity and bowel movement compared with open colectomy 95,96. The mechanisms involved may include reduced activation of inhibitory reflexes and local inflammation due to a reduction in surgical trauma 13. In clinical studies, earlier resolution of ileus has been demonstrated following laparoscopic colonic surgery. Four prospective randomized clinical studies comparing laparoscopic and open colonic techniques have shown a reduction in postoperative ileus with the laparoscopic method in two of three studies in which it was assessed 97-99 (Fig. 5). However, in a broader context the difference between laparoscopic and conventional abdominal surgery with respect to ileus remains to be clarified; expectations may have led to treatment bias favouring earlier food intake and less routine use of nasogastric tubes in laparoscopic groups.

Fig. 5 Randomized clinical trials comparing the effect of laparoscopic versus open colonic surgery on ileus. F=flatus, D=defaecation. *P < 0.05

Prokinetic agents Cisapride Cisapride enhances acetylcholine release from the intrinsic plexus and acts as a serotonin receptor agonist; it may stimulate all aspects of gastrointestinal motility 100. Nine randomized controlled clinical trials have been performed on postoperative patients 101-109 (Fig. 6). In one study, however, patients (neonates) were included only after an initial 10-day period of ileus 109. In four studies cisapride significantly reduced ileus; it was administered intravenously in three of them. In one of these studies 107 43 per cent of patients undergoing various abdominal operations had flatus 2 h after the intravenous administration of cisapride 4 mg, compared with 12 per cent in the control group. In another study on postoperative patients 108 55 per cent had flatus 4 h after the intravenous administration of cisapride 8 mg, compared with 29 per cent of patients receiving placebo. In only one clinical trial 105 was cisapride administered orally; ileus was reduced from 4.8 to 3.7 days. In three of the four remaining (negative) studies, cisapride was administered rectally. These results suggest that the beneficial effect of cisapride on ileus depends on the route of its administration, favouring intravenous or, possibly, oral administration in the postoperative period. However, adverse cardiac effects may occur with cisapride, which may therefore be contraindicated in high-risk patients 110.

Fig. 6 Randomized clinical trials evaluating the effect of cisapride on a duration of ileus and b percentage of subjects with resolved ileus. Various indicators of postoperative ileus (flatus (F) and defaecation (D)) and were used as well as various doses of drug (4-30 mg administered for between 1 and 7 days) and various routes of administration (intravenous, rectal and oral). In all studies, patients underwent various abdominal (not specified) or colonic procedures. In studies with assessments of both flatus and defaecation, defaecation was given a higher priority than flatus. *P < 0.05 Ceruletide Ceruletide is a synthetic peptide whose cholecystokinin antagonist activity may stimulate gastrointestinal motility. In two clinical placebo-controlled studies ceruletide has been found to reduce ileus slightly 111,112. However, side-effects such as nausea and vomiting, which may require additional antiemetic treatment, limit the potential use of ceruletide. Further studies are necessary on other cholecystokinin antagonists before clinical recommendations can be made. Erythromycin Erythromycin stimulates the motilin receptor and so may induce a MMC. However, the only existing prospective clinical randomized study found no effect of postoperatively administered erythromycin on ileus after abdominal surgery 113.

Metoclopramide Metoclopramide may potentially influence gastrointestinal motility by acting as a dopamine antagonist, as well as by direct and indirect effects on cholinergic and serotonergic receptors throughout the gastrointestinal tract. Six controlled clinical studies exist in patients undergoing abdominal surgery to assess the effect of metoclopramide on ileus. In three of the studies endpoints such as bowel sounds, food intake and defaecation were assessed 114-116 (Fig. 7). Of the remaining three studies, one assessed ileus with a functional combination score 117. In another 118, the study population was divided according to the median drug dose received, and in the third study 119 the number of patients with bowel sounds in each group was counted after 2 days. Despite the variety of endpoints, none of these studies has demonstrated a significant effect of metoclopramide on the resolution of postoperative ileus.

Fig. 7 Effect of metoclopramide on ileus. Three of the existing six clinical trials are shown. The remaining three studies are discussed only in the text, owing to the great variety of endpoints assessed and statistical analyses. Surgery was aortic 114, various abdominal 115 or cholecystectomy 116. Indicators of ileus were flatus (F), time to solid food intake (I) and defaecation (D) Somatostatin Somatostatin is known to inhibit the secretion of gastrointestinal hormones. In experimental studies the somatostatin analogue, octreotide, has been shown to enhance colonic transit and to restore interdigestive myoelectrical complexes 21,120. The effect of somatostatin analogues on postoperative ileus has not been investigated in clinical studies. Adrenoblockers Although inhibitory reflexes mediated by sympathetic visceral afferents are important in the development of ileus, the effect of adrenoblocking agents in reducing ileus in the clinical setting has been rather limited 121,122. It is probably clinically insignificant. Laxatives Despite the wide use of laxatives with prokinetic effect in non-surgical patients, no randomized clinical study exists to assess the possible beneficial effect of postoperative laxatives on ileus. In fact only one clinical study (unblinded and nonrandomized) has assessed the effect of postoperatively administered laxatives (magnesia and biscolic suppositories) and this was in 20 consecutive

gynaecological patients 123. The duration of ileus in the treated group (time to flatus and bowel movement) was 3 days. Furthermore a 50 per cent reduction in length of hospital stay (4 versus 8 days) was noted. Conclusion Several prokinetic drugs are available, but so far none has been found effective in the elimination or reduction of postoperative ileus, with the exception of cisapride. However, as a sole agent, cisapride is likely to be of limited clinical relevance, although it may be useful in conjunction with other techniques. The future In the future we must consider a multimodal concept of controlling perioperative pathophysiology, postoperative ileus and rehabilitation. The pathogenesis of ileus involves mainly inhibitory neural reflexes, and inflammatory mediators and neurotransmitters from the site of intestinal injury. Effective pain relief with continuous thoracic epidural local anaesthetics is the most important method of reducing ileus. It appears that several techniques, when studied individually, may reduce the duration of ileus without eliminating the problem. A rational treatment would therefore be a combination of avoidance of the routine use of nasogastric tubes, the use of a continuous thoracic epidural local anaesthetic regimen for at least 48 h, the use of additional opioid-sparing analgesia, and routine early oral nutrition. Other additional methods may be supportive, for instance cisapride and the use of a laparoscopic technique. Unfortunately, no randomized studies are available to document the validity of this multimodal concept, but several case series suggest that such an approach may significantly reduce, or even eliminate, ileus after colonic surgery. Sixty consecutive patients scheduled for elective open colorectal surgery participated in an accelerated rehabilitation programme. Their median age was 74 years and 33 per were American Society of Anesthesiologists grade III or IV. Analgesia consisted of continuous thoracic epidural blockade with local anaesthetics over 48 h, combined with systemic paracetamol and ketorolac. Nasogastric tubes were removed immediately after surgery and oral food intake was started on the day of operation. The patients were mobilized within 8 h of surgery. The median hospital stay of these unselected patients was 2 days and ileus was considered to be significantly reduced, as 95 per cent of the patients tolerated normal food and had defaecation within 48 h of operation 124,125. Similar results have been found in a group of 17 patients undergoing elective colonic surgery; 12 passed stool within 48 h, despite the fact that several patients received morphine after operation 126. Thirty-nine patients undergoing laparoscopic colonic surgery participated in a similar programme and 35 had normal bowel function on day 2 after operation 127,128. The combination of epidural anaesthesia and analgesia and of a standardized care protocol has also been found to reduce the duration of postoperative ileus and hospital stay by 1 day, compared with other analgesic techniques, in patients undergoing elective colorectal surgery 50,129. In conclusion, these preliminary findings suggest that a multimodal rehabilitation approach may result in a major improvement in terms of early resolution of postoperative ileus. Further trials are needed to document efficacy with respect to different types of operation, safety and the potential for reducing overall postoperative morbidity and cost. Acknowledgements This work was supported by a grant from the Danish Research Council (28809). References 1 Bueno L, Fioramonti J, Delvaux M, Frexinos J. Mediators and pharmacology of visceral sensitivity: from basic to clinical investigations. Gastroenterology 1997; 112: 1714-43. [Medline Link] [Context Link] 2 Resnick J, Greenwald DA, Brandt LJ. Delayed gastric emptying and postoperative ileus after nongastric abdominal surgery: part I. Am J Gastroenterol 1997; 92: 751-62. [Medline Link] [Context Link] 3 Resnick J, Greenwald DA, Brandt LJ. Delayed gastric emptying and postoperative ileus after nongastric abdominal surgery: part II. Am J Gastroenterol 1997; 92: 934-40. [Medline Link] [Context Link]

4 Bungard TJ, Kale-Pradhan PB. Prokinetic agents for the treatment of postoperative ileus in adults: a review of the literature. Pharmacotherapy 1999; 19: 416-23. [Medline Link] [Context Link] 5 Thompson JS, Quigley EM. Prokinetic agents in the surgical patient. Am J Surg 1999; 177: 508-14. [Fulltext Link] [Medline Link] [Context Link] 6 Waldhausen JH, Shaffrey ME, Skenderis BS II, Jones RS, Schirmer BD. Gastrointestinal myoelectric and clinical patterns of recovery after laparotomy. Ann Surg 1990; 211: 777-84. [Medline Link] [Context Link] 7 Condon RE, Cowles VE, Ferraz AA, Carilli S, Carlson ME, Ludwig K et al. Human colonic smooth muscle electrical activity during and after recovery from postoperative ileus. Am J Physiol 1995; 269: G408-17. [Context Link] 8 Wilson JP. Postoperative motility of the large intestine in man. Gut 1975; 16: 689-92. [Medline Link] [Context Link] 9 Shibata Y, Toyoda S, Nimura Y, Miyati M. Patterns of intestinal motility recovery during the early stage following abdominal surgery: clinical and manometric study. World J Surg 1997; 21: 806-9. [Medline Link] [Context Link] 10 Roberts JP, Benson MJ, Rogers J, Deeks JJ, Williams NS. Characterization of distal colonic motility in early postoperative period and effect of colonic anastomosis. Dig Dis Sci 1994; 39: 1961-7. [Medline Link] [Context Link] 11 Petros JG, Realica R, Ahmad S, Rimm EB, Robillard RJ. Patient-controlled analgesia and prolonged ileus after uncomplicated colectomy. Am J Surg 1995; 170: 371-4. [Fulltext Link] [Medline Link] [Context Link] 12 Graber JN, Schulte WJ, Condon RE, Cowles VE. Relationship of duration of postoperative ileus to extent and site of operative dissection. Surgery 1982; 92: 87-92. [Medline Link] [Context Link] 13 Kalff JC, Schraut WH, Simmons RL, Bauer AJ. Surgical manipulation of the gut elicits an intestinal muscularis inflammatory response resulting in postsurgical ileus. Ann Surg 1998; 228: 652-63. [Medline Link] [Context Link] 14 Hotokezaka M, Mentis EP, Patel SP, Combs MJ, Teates CD, Schirmer BD. Recovery of gastrointestinal tract motility and myoelectric activity change after abdominal surgery. Arch Surg 1997; 132: 410-17. [Fulltext Link] [Medline Link] [Context Link] 15 Yee MK, Evans WD, Facey PE, Hayward MW, Rosen M. Gastric emptying and small bowel transit in male volunteers after i.m. ketorolac and morphine. Br J Anaesth 1991; 67: 426-31. [Medline Link] [Context Link] 16 Rimback G, Cassuto J, Tollesson PO. Treatment of postoperative paralytic ileus by intravenous lidocaine infusion. Anesth Analg 1990; 70: 414-19. [Medline Link] [Context Link] 17 Zittel TT, Lloyd KC, Rothenhofer I, Wong H, Walsh JH, Raybould HE. Calcitonin gene-related peptide and spinal afferents partly mediate postoperative colonic ileus in the rat. Surgery 1998; 123: 518-27. [Medline Link] [Context Link] 18 Holzer P, Lippe IT, Holzer-Petsche U. Inhibition of gastrointestinal transit due to surgical trauma or peritoneal irritation is reduced in capsaicin-treated rats. Gastroenterology 1986; 91: 360-3. [Medline Link] [Context Link] 19 De Winter BY, Robberecht P, Boeckxstaens GE, De Man JG, Moreels TG, Herman AG et al. Role of VIP1/PACAP receptors in postoperative ileus in rats. Br J Pharmacol 1998; 124: 1181-6. [Medline Link] [Context Link] 20 De Winter BY, Boeckxstaens GE, De Man JG, Moreels TG, Herman AG, Pelckmans PA. Differential effect of indomethacin and ketorolac on postoperative ileus in rats. Eur J Pharmacol 1998; 344: 71-6. [Medline Link] [Context Link]

21 Cullen JJ, Eagon JC, Kelly KA. Gastrointestinal peptide hormones during postoperative ileus. Effect of octreotide. Dig Dis Sci 1994; 39: 1179-84. [Medline Link] [Context Link] 22 Espat NJ, Cheng G, Kelley MC, Vogel SB, Sninsky CA, Hocking MP. Vasoactive intestinal peptide and substance P receptor antagonists improve postoperative ileus. J Surg Res 1995; 58: 719-23. [Medline Link] [Context Link] 23 Kalff JC, Schraut WH, Billiar TR, Simmons RL, Bauer AJ. Role of inducible nitric oxide synthase in postoperative intestinal smooth muscle dysfunction in rodents. Gastroenterology 2000; 118: 316-27. [Medline Link] [Context Link] 24 Cullen JJ, Caropreso DK, Hemann LL, Hinkhouse M, Conklin JL, Ephgrave KS. Pathophysiology of adynamic ileus. Dig Dis Sci 1997; 42: 731-7. [Medline Link] [Context Link] 25 Cullen JJ, Ephgrave KS, Caropreso DK. Gastrointestinal myoelectric activity during endotoxemia. Am J Surg 1996; 171: 5969. [Fulltext Link] [Medline Link] [Context Link] 26 Zittel TT, Reddy SN, Plourde V, Raybould HE. Role of spinal afferents and calcitonin gene-related peptide in the postoperative gastric ileus in anesthetized rats. Ann Surg 1994; 219: 79-87. [Medline Link] [Context Link] 27 Barquist E, Bonaz B, Martinez V, Rivier J, Zinner MJ, Tache Y. Neuronal pathways involved in abdominal surgery-induced gastric ileus in rats. Am J Physiol 1996; 270: R888-94. [Medline Link] [Context Link] 28 Martinez V, Rivier J, Wang L, Tache Y. Central injection of a new corticotropin-releasing factor (CRF) antagonist, astressin, blocks CRF- and stress-related alterations of gastric and colonic motor function. J Pharmacol Exp Ther 1997; 280: 754-60. [Medline Link] [Context Link] 29 Thorn SE, Wattwil M, Lindberg G, Sawe J. Systemic and central effects of morphine on gastroduodenal motility. Acta Anaesthesiol Scand 1996; 40: 177-86. [Medline Link] [Context Link] 30 Ferraz AA, Cowles VE, Condon RE, Schulte WJ. Opioid and nonopioid analgesic drug effects on colon contractions in monkeys. Dig Dis Sci 1995; 40: 1417-19. [Medline Link] [Context Link] 31 De Winter BY, Boeckxstaens GE, De Man JG, Moreels TG, Herman AG, Pelckmans PA. Effects of mu- and kappa-opioid receptors on postoperative ileus in rats. Eur J Pharmacol 1997; 339: 63-7. [Medline Link] [Context Link] 32 Friese N, Chevalier E, Angel F, Pascaud X, Junien JL, Dahl SG et al. Reversal by kappa-agonists of peritoneal irritationinduced ileus and visceral pain in rats. Life Sci 1997; 60: 625-34. [Medline Link] [Context Link] 33 Riviere PJ, Pascaud X, Chevalier E, Junien JL. Fedotozine reversal of peritoneal-irritation-induced ileus in rats: possible peripheral action on sensory afferents. J Pharmacol Exp Ther 1994; 270: 846-50. [Medline Link] [Context Link] 34 Kalff JC, Carlos TM, Schraut WH, Billiar TR, Simmons RL, Bauer AJ. Surgically induced leukocytic infiltrates within the rat intestinal muscularis mediate postoperative ileus. Gastroenterology 1999; 117: 378-87. [Medline Link] [Context Link] 35 Kalff JC, Buchholz BM, Eskandari MK, Hierholzer C, Schraut WH, Simmons RL et al. Biphasic response to gut manipulation and temporal correlation of cellular infiltrates and muscle dysfunction in rat. Surgery 1999; 126: 498-509. [Medline Link] [Context Link] 36 Ogilvy AJ, Smith G. The gastrointestinal tract after anaesthesia. Eur J Anaesthesiol Suppl 1995; 10: 35-42. [Medline Link] [Context Link] 37 Pedersen FM, Wilken-Jensen C, Knudsen F, Lindekaer AL, Svare EI. The influence of nitrous oxide on recovery of bowel function after abdominal hysterectomy. Acta Anaesthesiol Scand 1993; 37: 692-6. [Medline Link] [Context Link]

38 Krogh B, Jorn Jensen P, Henneberg SW, Hole P, Kronborg O. Nitrous oxide does not influence operating conditions or postoperative course in colonic surgery. Br J Anaesth 1994; 72: 55-7. [Medline Link] [Context Link] 39 Jensen AG, Kalman SH, Nystrom PO, Eintrei C. Anaesthetic technique does not influence postoperative bowel function: a comparison of propofol, nitrous oxide and isoflurane. Can J Anaesth 1992; 39: 938-43. [Medline Link] [Context Link] 40 Kehlet H. Modification of responses to surgery by neural blockade: clinical implications. In: Cousins MJ, Bridenbaugh PO, eds. Neural Blockade in Clinical Anesthesia and Management of Pain. 3rd ed. Philadelphia, Pennsylvania: Lippincott-Raven, 1998: 129-75. [Context Link] 41 Kehlet H. Multimodal approach to control postoperative pathophysiology and rehabilitation. Br J Anaesth 1997; 78: 606-17. [Medline Link] [Context Link] 42 Liu S, Carpenter RL, Neal JM. Epidural anesthesia and analgesia. Their role in postoperative outcome. Anesthesiology 1995; 82: 1474-506. [Fulltext Link] [Medline Link] [Context Link] 43 Steinbrook RA. Epidural anesthesia and gastrointestinal motility. Anesth Analg 1998; 86: 837-44. [Fulltext Link] [Medline Link] [Context Link] 44 Wallin G, Cassuto J, Hogstrom S, Rimback G, Faxen A, Tollesson PO. Failure of epidural anesthesia to prevent postoperative paralytic ileus. Anesthesiology 1986; 65: 292-7. [Medline Link] [Context Link] 45 Scheinin B, Asantila R, Orko R. The effect of bupivacaine and morphine on pain and bowel function after colonic surgery. Acta Anaesthesiol Scand 1987; 31: 161-4. [Medline Link] [Context Link] 46 Ahn H, Bronge A, Johansson K, Ygge H, Lindhagen J. Effect of continuous postoperative epidural analgesia on intestinal motility. Br J Surg 1988; 75: 1176-8. [Medline Link] [Context Link] 47 Wattwil M, Thoren T, Hennerdal S, Garvill JE. Epidural analgesia with bupivacaine reduces postoperative paralytic ileus after hysterectomy. Anesth Analg 1989; 68: 353-8. [Medline Link] [Context Link] 48 Bredtmann RD, Herden HN, Teichmann W, Moecke HP, Kniesel B, Baetgen R et al. Epidural analgesia in colonic surgery: results of a randomized prospective study. Br J Surg 1990; 77: 638-42. [Medline Link] [Context Link] 49 Riwar A, Schar B, Grotzinger U. Effect of continuous postoperative analgesia with peridural bupivacaine on intestinal motility following colorectal resection. Helv Chir Acta 1991; 58: 729-33. [Context Link] 50 Liu SS, Carpenter RL, Mackey DC, Thirlby RC, Rupp SM, Shine TS et al. Effects of perioperative analgesic technique on rate of recovery after colon surgery. Anesthesiology 1995; 83: 757-65. [Fulltext Link] [Medline Link] [Context Link] 51 Neudecker J, Schwenk W, Junghans T, Pietsch S, Bohm B, Muller JM. Randomized controlled trial to examine the influence of thoracic epidural analgesia on postoperative ileus after laparoscopic sigmoid resection. Br J Surg 1999; 86: 1292-5. [Fulltext Link] [Medline Link] [Context Link] 52 Thoren T, Sundberg A, Wattwil M, Garvill JE, Jurgensen U. Effects of epidural bupivacaine and epidural morphine on bowel function and pain after hysterectomy. Acta Anaesthesiol Scand 1989; 33: 181-5. [Medline Link] [Context Link] 53 Asantila R, Eklund P, Rosenberg PH. Continuous epidural infusion of bupivacaine and morphine for postoperative analgesia after hysterectomy. Acta Anaesthesiol Scand 1991; 35: 513-17. [Medline Link] [Context Link] 54 Bisgaard C, Mouridsen P, Dahl JB. Continuous lumbar epidural bupivacaine plus morphine versus epidural morphine after major abdominal surgery. Eur J Anaesthesiol 1990; 7: 219-25. [Context Link]

55 Hjortso NC, Neumann P, Frosig F, Andersen T, Lindhard A, Rogon E et al. A controlled study on the effect of epidural analgesia with local anaesthetics and morphine on morbidity after abdominal surgery. Acta Anaesthesiol Scand 1985; 29: 790-6. [Medline Link] [Context Link] 56 Seeling W, Bruckmooser KP, Hufner C, Kneitinger E, Rigg C, Rockemann M. No reduction in postoperative complications by the use of catheterized epidural analgesia following major abdominal surgery. Anaesthesist 1990; 39: 33-40. [Medline Link] [Context Link] 57 Jayr C, Thomas H, Rey A, Farhat F, Lasser P, Bourgain JL. Postoperative pulmonary complications. Epidural analgesia using bupivacaine and opioids versus parenteral opioids. Anesthesiology 1993; 78: 666-76. [Medline Link] [Context Link] 58 Mann C, Pouzeratte Y, Boccara G, Peccoux C, Vergne C, Brunat G et al. Comparison of intravenous or epidural patientcontrolled analgesia in the elderly after major abdominal surgery. Anesthesiology 2000; 92: 433-41. [Fulltext Link] [Medline Link] [Context Link] 59 Wulf H, Biscoping J, Beland B, Bachmann-Mennenga B, Motsch J. Ropivacaine epidural anesthesia and analgesia versus general anesthesia and intravenous patient-controlled analgesia with morphine in the perioperative management of hip replacement. Ropivacaine Hip Replacement Multicenter Study Group. Anesth Analg 1999; 89: 111-16. [Fulltext Link] [Medline Link] [Context Link] 60 Rimback G, Cassuto J, Wallin G, Westlander G. Inhibition of peritonitis by amide local anesthetics. Anesthesiology 1988; 69: 881-6. [Medline Link] [Context Link] 61 Udassin R, Eimerl D, Schiffman J, Haskel Y. Epidural anesthesia accelerates the recovery of postischemic bowel motility in the rat. Anesthesiology 1994; 80: 832-6. [Medline Link] [Context Link] 62 Groudine SB, Fisher HA, Kaufman RP Jr, Patel MK, Wilkins LJ, Mehta SA et al. Intravenous lidocaine speeds the return of bowel function, decreases postoperative pain, and shortens hospital stay in patients undergoing radical retropubic prostatectomy. Anesth Analg 1998; 86: 235-9. [Fulltext Link] [Medline Link] [Context Link] 63 Rimback G, Cassuto J, Faxen A, Hogstrom S, Wallin G, Tollesson PO. Effect of intra-abdominal bupivacaine instillation on postoperative colonic motility. Gut 1986; 27: 170-5. [Medline Link] [Context Link] 64 Delvaux M, Louvel D, Lagier E, Scherrer B, Abitbol JL, Frexinos J. The kappa agonist fedotozine relieves hypersensitivity to colonic distension in patients with irritable bowel syndrome. Gastroenterology 1999; 116: 38-45. [Medline Link] [Context Link] 65 Yuan CS, Foss JF, O'Connor M, Osinski J, Kamson T, Moss J et al. Methylnaltrexone for reversal of constipation due to chronic methadone use: a randomized controlled trial. JAMA 2000; 283: 367-72. [Fulltext Link] [Medline Link] [Context Link] 66 Kehlet H, Rung GW, Callesen T. Postoperative opioid analgesia: time for a reconsideration? J Clin Anesth 1996; 8: 441-5. [Medline Link] [Context Link] 67 Merry A, Power W. Perioperative NSAIDs: towards greater safety. Pain Rev 1995; 2: 268-91. [Context Link] 68 Kelley MC, Hocking MP, Marchand SD, Sninsky CA. Ketorolac prevents postoperative small intestinal ileus in rats. Am J Surg 1993; 165: 107-11. [Fulltext Link] [Medline Link] [Context Link] 69 Cheng G, Cassissi C, Drexler PG, Vogel SB, Sninsky CA, Hocking MP. Salsalate, morphine, and postoperative ileus. Am J Surg 1996; 171: 85-8. [Fulltext Link] [Medline Link] [Context Link] 70 Ferraz AA, Cowles VE, Condon RE, Carilli S, Ezberci F, Frantzides CT et al. Nonopioid analgesics shorten the duration of postoperative ileus. Am Surg 1995; 61: 1079-83. [Medline Link] [Context Link]

71 Cheatham ML, Chapman WC, Key SP, Sawyers JL. A meta-analysis of selective versus routine nasogastric decompression after elective laparotomy. Ann Surg 1995; 221: 469-76. [Medline Link] [Context Link] 72 Sagar PM, Kruegener G, MacFie J. Nasogastric intubation and elective abdominal surgery. Br J Surg 1992; 79: 1127-31. [Medline Link] [Context Link] 73 Rao SS, Beaty J, Chamberlain M, Lambert PG, Gisolfi C. Effects of acute graded exercise on human colonic motility. Am J Physiol 1999; 276: G1221-6. [Medline Link] [Context Link] 74 Waldhausen JH, Schirmer BD. The effect of ambulation on recovery from postoperative ileus. Ann Surg 1990; 212: 671-7. [Medline Link] [Context Link] 75 Allen C, Glasziou P, Del Mar C. Bed rest: a potentially harmful treatment needing more careful evaluation. Lancet 1999; 354: 1229-33. [Fulltext Link] [Medline Link] [Context Link] 76 Harper CM, Lyles YM. Physiology and complications of bed rest. J Am Geriatr Soc 1988; 36: 1047-54. [Medline Link] [Context Link] 77 Johnson LR. Physiology of the Gastrointestinal Tract. New York: Raven Press, 1994. [Context Link] 78 Moore FA, Feliciano DV, Andrassy RJ, McArdle AH, Booth FV, Morgenstein-Wagner TB et al. Early enteral feeding, compared with parenteral, reduces postoperative septic complications. The results of a meta-analysis. Ann Surg 1992; 216: 17283. [Medline Link] [Context Link] 79 Moss G. Maintenance of gastrointestinal function after bowel surgery and immediate enteral full nutrition. II. Clinical experience, with objective demonstration of intestinal absorption and motility. JPEN J Parenter Enteral Nutr 1981; 5: 215-20. [Medline Link] [Context Link] 80 Reissman P, Teoh TA, Cohen SM, Weiss EG, Nogueras JJ, Wexner SD. Is early oral feeding safe after elective colorectal surgery? A prospective randomized trial. Ann Surg 1995; 222: 73-7. [Medline Link] [Context Link] 81 Binderow SR, Cohen SM, Wexner SD, Nogueras JJ. Must early postoperative oral intake be limited to laparoscopy? Dis Colon Rectum 1994; 37: 584-9. [Medline Link] [Context Link] 82 Ortiz H, Armendariz P, Yarnoz C. Is early postoperative feeding feasible in elective colon and rectal surgery? Int J Colorectal Dis 1996; 11: 119-21. [Medline Link] [Context Link] 83 Bufo AJ, Feldman S, Daniels GA, Lieberman RC. Early postoperative feeding. Dis Colon Rectum 1994; 37: 1260-5. [Medline Link] [Context Link] 84 Hartsell PA, Frazee RC, Harrison JB, Smith RW. Early postoperative feeding after elective colorectal surgery. Arch Surg 1997; 132: 518-20. [Fulltext Link] [Medline Link] [Context Link] 85 Singh G, Ram RP, Khanna SK. Early postoperative enteral feeding in patients with nontraumatic intestinal perforation and peritonitis. J Am Coll Surg 1998; 187: 142-6. [Medline Link] [Context Link] 86 Di Fronzo LA, Cymerman J, O'Connell TX. Factors affecting early postoperative feeding following elective open colon resection. Arch Surg 1999; 134: 941-5. [Fulltext Link] [Medline Link] [Context Link] 87 Schilder JM, Hurteau JA, Look KY, Moore DH, Raff G, Stehman FB et al. A prospective controlled trial of early postoperative oral intake following major abdominal gynecologic surgery. Gynecol Oncol 1997; 67: 235-40. [Medline Link] [Context Link]

88 Stewart BT, Woods RJ, Collopy BT, Fink RJ, Mackay JR, Keck JO. Early feeding after elective open colorectal resections: a prospective randomized trial. Aust N Z J Surg 1998; 68: 125-8. [Medline Link] [Context Link] 89 Pearl ML, Valea FA, Fischer M, Mahler L, Chalas E. A randomized controlled trial of early postoperative feeding in gynecologic oncology patients undergoing intra-abdominal surgery. Obstet Gynecol 1998; 92: 94-7. [Medline Link] [Context Link] 90 Cutillo G, Maneschi F, Franchi M, Giannice R, Scambia G, Benedetti-Panici P. Early feeding compared with nasogastric decompression after major oncologic gynecologic surgery: a randomized study. Obstet Gynecol 1999; 93: 41-5. [Medline Link] [Context Link] 91 Heslin MJ, Latkany L, Leung D, Brooks AD, Hochwald SN, Pisters PW et al. A prospective, randomized trial of early enteral feeding after resection of upper gastrointestinal malignancy. Ann Surg 1997; 226: 567-77. [Medline Link] [Context Link] 92 Watters JM, Kirkpatrick SM, Norris SB, Shamji FM, Wells GA. Immediate postoperative enteral feeding results in impaired respiratory mechanics and decreased mobility. Ann Surg 1997; 226: 369-77. [Medline Link] [Context Link] 93 Hotokezaka M, Combs MJ, Mentis EP, Schirmer BD. Recovery of fasted and fed gastrointestinal motility after open versus laparoscopic cholecystectomy in dogs. Ann Surg 1996; 223: 413-19. [Medline Link] [Context Link] 94 Garcia-Caballero M, Vara-Thorbeck C. The evolution of postoperative ileus after laparoscopic cholecystectomy. A comparative study with conventional cholecystectomy and sympathetic blockade treatment. Surg Endosc 1993; 7: 416-19. [Medline Link] [Context Link] 95 Bohm B, Milsom JW, Fazio VW. Postoperative intestinal motility following conventional and laparoscopic intestinal surgery. Arch Surg 1995; 130: 415-19. [Fulltext Link] [Medline Link] [Context Link] 96 Davies W, Kollmorgen CF, Tu QM, Donohue JH, Thompson GB, Nelson H et al. Laparoscopic colectomy shortens postoperative ileus in a canine model. Surgery 1997; 121: 550-5. [Medline Link] [Context Link] 97 Lacy AM, Garcia-Valdecasas JC, Pique JM, Delgado S, Campo E, Bordas JM et al. Short-term outcome analysis of a randomized study comparing laparoscopic vs open colectomy for colon cancer. Surg Endosc 1995; 9: 1101-5. [Medline Link] [Context Link] 98 Schwenk W, Bohm B, Haase O, Junghans T, Muller JM. Laparoscopic versus conventional colorectal resection: a prospective randomised study of postoperative ileus and early postoperative feeding. Langenbecks Arch Surg 1998; 383: 49-55. [Medline Link] [Context Link] 99 Milsom JW, Bohm B, Hammerhofer KA, Fazio V, Steiger E, Elson P. A prospective, randomized trial comparing laparoscopic versus conventional techniques in colorectal cancer surgery: a preliminary report. J Am Coll Surg 1998; 187: 46-54. [Medline Link] [Context Link] 100 Barone JA, Jessen LM, Colaizzi JL, Bierman RH. Cisapride: a gastrointestinal prokinetic drug. Ann Pharmacother 1994; 28: 488-500. [Medline Link] [Context Link] 101 Tollesson PO, Cassuto J, Rimback G, Faxen A, Bergman L, Mattsson E. Treatment of postoperative paralytic ileus with cisapride. Scand J Gastroenterol 1991; 26: 477-82. [Medline Link] [Context Link] 102 Hallerback H, Bergman B, Bong H, Ekstrom P, Glise H, Lundgren K et al. Cisapride in the treatment of post-operative ileus. Aliment Pharmacol Ther 1991; 5: 503-11. [Medline Link] [Context Link]

103 Benson MJ, Roberts JP, Wingate DL, Rogers J, Deeks JJ, Castillo FD et al. Small bowel motility following major intraabdominal surgery: the effects of opiates and rectal cisapride. Gastroenterology 1994; 106: 924-36. [Medline Link] [Context Link] 104 Roberts JP, Benson MJ, Rogers J, Deeks JJ, Wingate DL, Williams NS. Effect of cisapride on distal colonic motility in the early postoperative period following left colonic anastomosis. Dis Colon Rectum 1995; 38: 139-45. [Medline Link] [Context Link] 105 Brown TA, McDonald J, Williard W. A prospective, randomized, double-blinded, placebo-controlled trial of cisapride after colorectal surgery. Am J Surg 1999; 177: 399-401. [Fulltext Link] [Medline Link] [Context Link] 106 von Ritter C, Hunter S, Hinder RA. Cisapride does not reduce postoperative paralytic ileus. S Afr J Surg 1987; 25: 19-21. [Medline Link] [Context Link] 107 Boghaert A, Haesaert G, Mourisse P, Verlinden M. Placebo-controlled trial of cisapride in postoperative ileus. Acta Anaesthesiol Belg 1987; 38: 195-9. [Medline Link] [Context Link] 108 Verlinden M, Michiels G, Boghaert A, de Coster M, Dehertog P. Treatment of postoperative gastrointestinal atony. Br J Surg 1987; 74: 614-17. [Medline Link] [Context Link] 109 Lander A, Redkar R, Nicholls G, Lawson A, Choudhury SR, Corkery JJ et al. Cisapride reduces neonatal postoperative ileus: randomised placebo controlled trial. Arch Dis Child Fetal Neonatal Ed 1997; 77: F119-22. [Medline Link] [Context Link] 110 Tonini M, De Ponti F, Di Nucci A, Crema F. Review article: cardiac adverse effects of gastrointestinal prokinetics. Aliment Pharmacol Ther 1999; 13: 1585-91. [Medline Link] [Context Link] 111 Sadek SA, Cranford C, Eriksen C, Walker M, Campbell C, Baker PR et al. Pharmacological manipulation of adynamic ileus: controlled randomized double-blind study of ceruletide on intestinal motor activity after elective abdominal surgery. Aliment Pharmacol Ther 1988; 2: 47-54. [Medline Link] [Context Link] 112 Madsen PV, Lykkegaard-Nielsen M, Nielsen OV. Ceruletide reduces postoperative intestinal paralysis. A double-blind placebo-controlled trial. Dis Colon Rectum 1983; 26: 159-60. [Medline Link] [Context Link] 113 Bonacini M, Quiason S, Reynolds M, Gaddis M, Pemberton B, Smith O. Effect of intravenous erythromycin on postoperative ileus. Am J Gastroenterol 1993; 88: 208-11. [Medline Link] [Context Link] 114 Jepsen S, Klaerke A, Nielsen PH, Simonsen O. Negative effect of metoclopramide in postoperative adynamic ileus. A prospective, randomized, double blind study. Br J Surg 1986; 73: 290-1. [Medline Link] [Context Link] 115 Cheape JD, Wexner SD, James K, Jagelman DG. Does metoclopramide reduce the length of ileus after colorectal surgery? A prospective randomized trial. Dis Colon Rectum 1991; 34: 437-41. [Medline Link] [Context Link] 116 Tollesson PO, Cassuto J, Faxen A, Rimback G, Mattsson E, Rosen S. Lack of effect of metoclopramide on colonic motility after cholecystectomy. Eur J Surg 1991; 157: 355-8. [Medline Link] [Context Link] 117 Breivik H, Lind B. Anti-emetic and propulsive peristaltic properties of metoclopramide. Br J Anaesth 1971; 43: 400-3. [Medline Link] [Context Link] 118 Davidson ED, Hersh T, Brinner RA, Barnett SM, Boyle LP. The effects of metoclopramide on postoperative ileus. A randomized double-blind study. Ann Surg 1979; 190: 27-30. [Medline Link] [Context Link] 119 Kivalo I, Miettinen K. The effects of metoclopramide on postoperative nausea and bowel function. Ann Chir Gynaecol Fenn 1970; 59: 155-8. [Medline Link] [Context Link]

120 Cullen JJ, Eagon JC, Dozois EJ, Kelly KA. Treatment of acute postoperative ileus with octreotide. Am J Surg 1993; 165: 113-19. [Fulltext Link] [Medline Link] [Context Link] 121 Hallerback B, Carlsen E, Carlsson K, Enkvist C, Glise H, Haffner J et al. Beta-adrenoceptor blockade in the treatment of postoperative adynamic ileus. Scand J Gastroenterol 1987; 22: 149-55. [Context Link] 122 Hallerback B, Ander S, Glise H. Effect of combined blockade of beta-adrenoceptors and acetylcholinesterase in the treatment of postoperative ileus after cholecystectomy. Scand J Gastroenterol 1987; 22: 420-4. [Medline Link] [Context Link] 123 Fanning J, Yu-Brekke S. Prospective trial of aggressive postoperative bowel stimulation following radical hysterectomy. Gynecol Oncol 1999; 73: 412-14. [Medline Link] [Context Link] 124 Basse L, Jakobsen DH, Billesblle P, Werner M, Kehlet H. A clinical pathway to accelerate recovery after colonic resection. Ann Surg 2000; 232: 51-7. [Medline Link] [Context Link] 125 Kehlet H, Mogensen T. Hospital stay of 2 days after open sigmoidectomy with a multimodal rehabilitation programme. Br J Surg 1999; 86: 227-30. [Fulltext Link] [Medline Link] [Context Link] 126 Moiniche S, Bulow S, Hesselfeldt P, Hestbaek A, Kehlet H. Convalescence and hospital stay after colonic surgery with balanced analgesia, early oral feeding, and enforced mobilisation. Eur J Surg 1995; 161: 283-8. [Medline Link] [Context Link] 127 Bardram L, Funch-Jensen P, Jensen P, Crawford ME, Kehlet H. Recovery after laparoscopic colonic surgery with epidural analgesia, and early oral nutrition and mobilisation. Lancet 1995; 345: 763-4. [Fulltext Link] [Medline Link] [Context Link] 128 Bardram L, Funch-Jensen P, Kehlet H. Rapid rehabilitation in elderly patients after laparoscopic colonic surgery. Br J Surg 2000 (in press). [Context Link] 129 Bradshaw BG, Liu SS, Thirlby RC. Standardized perioperative care protocols and reduced length of stay after colon surgery. J Am Coll Surg 1998; 186: 501-6. [Medline Link] [Context Link]

Accession Number: 00002413-200011000-00006 Copyright (c) 2000-2001 Ovid Technologies, Inc. Version: rel4.3.0, SourceID: 1.5031.1.149

Regional Anesthesia for Shoulder Surgery


Learning Objectives: 1. Describe the indications and contraindications for the various approaches to the brachial plexus appropriate for shoulder surgery 2. Describe the relevant anatomy of the brachial plexus 3. Describe the techniques, local anesthetics and adjuvants used for performing a block of the brachial plexus for shoulder surgery

Editorial

Interscalene Block: The Truth About Twitches

nethesia of the brachial plexus was first accomplished more than one century

ago when surgeons (the first regional anesthetists) dissected, then directly and separately injected the roots of the brachial plexus in the neck. Subsequently, percutaneous techniques for brachial plexus block that involved multiple separate injections of local anesthetic were described and practiced. Perhaps the most significant conceptual advance in brachial plexus anesthesia, both in terms of our scientific understanding as well as practical clinical application, can be attributed to Alon Winnie. Winnie promoted the concept of the brachial plexus as being enveloped by a fascial sheath. This conceptual sheath serves to isolate the plexus in a functional anatomical compartment. Based on this concept, he proposed that a single injection of local anesthetic anywhere along the tubular plexus sheath would result in successful brachial plexus anesthesia. The sheath concept paved the way for Winnies original description of the interscalene technique of brachial plexus block in 1970. 1 Brachial plexus block thus became analogous to epidural anesthesia; i.e., once the compartment is entered, a single injection of an adequate volume of local anesthetic results in successful anesthesia in the vast majority of cases. The consistent objective for the anesthetist performing the block reduced simply to reliably ascertaining that the needle tip is within the confines of the brachial plexus sheath. Whereas this problem appears to be straightforward, traditional teaching has included misconceptions and myths with regard to which paresthesias or twitches are acceptable when performing interscalene blocks. Only recently have some misconceptions been disproved by careful clinical research. Although often misreferenced to include the use of a nerve stimulator, Winnies original technique of interscalene block was in fact a single paresthesia technique. In his writings, Winnie emphasized that only a paresthesia below the level of the shoulder is acceptable, since a paresthesia to the shoulder could result from stimulation of the suprascapular nerve inside or outside the sheath. 2 Regional anesthesia textbooks also recommended that shoulder paresthesias be discarded and that more distal paresthesias be sought.2-5 Winnies technique has since been modified following the advent of the use of a peripheral nerve stimulator to incorporate this useful aid. Nigel Sharrock first promoted the acceptance of a shoulder paresthesia in clinical practice when performing interscalene block. He emphasized that nerve roots supplying sensory innervation to the shoulder were more superficial and, upon advancing the needle, were therefore more characteristically encountered earlier than were more distal paresthesias. Indeed, work in his department by Roch et al.6 in 1992 reinforced his clinical experience; a shoulder paresthesia was the first paresthesia encountered in 45% of patients in this study of 45 patients who underwent interscalene block. These results indicated that a shoulder paresthesia was as effective as a more distal paresthesia. The investigators discussed that further needle probing in search of a distal paresthesia may increase patient discomfort and theoretically increase block-related neuropraxia.
Accepted for publication October 26, 1999. doi:10.1053/rapm.2000.4402 See Silverstein et al. page 356 Regional Anesthesia and Pain Medicine, Vol 25, No 4 (JulyAugust), 2000: pp 340342 340.In

this issue of Regional

Anesthesia and Pain Medicine, Silverstein et al.7 report on interscalene block with a peripheral nerve stimulator. In a prospective study of 160 patients, they found that a deltoid twitch was as successful an indicator of correct

needle placement within the brachial plexus as was a biceps twitch. They found that an isolated deltoid twitch was elicited in 34% of patients. These investigators also emphasized that acceptance of a deltoid twitch may result in less needle trauma and less patient discomfort. The importance of careful clinical studies in regional anesthesia such as that performed by Silverstein et al. cannot be overemphasized. The practice of regional anesthesia has always been considered an art. However, it is crucial that the science of regional anesthesia is not ignored or sacrificed by the artist. A thorough understanding of the relevant anatomy as well as what actually occurs when we perform regional anesthesia despite long-held biases, improves our success and protects our patients. By accepting a more proximal deltoid twitch or shoulder paresthesia, we are doing far more than reducing extra needle probing. The search for a more distal evidence of brachial plexus innervation by paresthesia or twitch may require the needle to pass through the more proximal nerve roots of C5 or C6 to reach C7, C8, or T1 (Fig 1). The nerve roots at the level of the cricoid cartilage (where interscalene block is typically performed) are narrowly stacked in a cephalocaudad orientation. Once this anatomy is understood, it can be readily appreciated that acceptance of evidence of more proximal innervation when performing intersca-lene block may in fact be preferable. This is especially true if the patient is sedated, as was the case in the study by Silverstein et al., where patients received up to 3 mg intravenous midazolam and up to 50 g fentanyl. In a study by Urmey et al.8 on 20 unsedated patients undergoing interscalene block, a simple paresthesia technique was used. However, immediately after elicitation of the paresthesia, an attached peripheral nerve stimulator was turned
Fig 1. The search for a more distal evidence of brachial plexus innervation by paresthesia or twitch may require the needle to pass through the more proximal nerve roots of C5 or C6 to reach C7, C8, or T1. Interscalene Block: The Truth About Twitches c William F. Urmey 341.on and the amperage slowly

increased in 0.1-mA increments to 1 mA. Despite a 100% block success rate, only 25% of patients showed evidence of a twitch (2 deltoid, 2 arm, and 1 combined biceps/finger). Therefore, evidence of sensory nerve contact (paresthesia) and presence of the needle tip within the brachial plexus sheath (inferred from successful block) translated to ability to elicit a twitch in only 25% of patients. The first elicited paresthesia was to the shoulder in 85% of patients in this study. The authors concluded that placement of blocks in heavily sedated patients may be an unsafe practice. When a nerve-stimulating technique is used on a sedated patient, the patient may be unable to report such paresthesias that are not associated with evidence of motor response. It can also be stated that, if the shoulder paresthesias/deltoid twitches were discarded in this study, further advancement of the needle could have theoretically pierced the C5 or C6 nerve roots in up to 85% of the patients. An anesthesiologist must have a thorough knowledge of the anatomy for every block in his or her repertoire. Decisions to accept or discard any twitch or paresthesia must be made with attention to this anatomy. Careful attention to anatomy serves not only to increase success but to limit complications and discomfort as well. Importantly, use of a peripheral nerve stimulator should be considered as technical support when performing regional blocks. The nerve stimulator should never be regarded as a substitute for detailed knowledge of the anatomy or proper and careful technique. Rigorous clinical research is needed to eliminate subjective bias and preconceived notions that have been perpetuated far too long without scientific basis or adequate study. Thanks to Silverstein et al. for telling us the truth about twitches and interscalene block. William F. Urmey, M.D. Department of Anesthesiology

Weill Medical College of Cornell University New York, New York

References
1. Winnie AP. Interscalene brachial plexus block. Anesth Analg 1970;49:455-466. 2. Winnie AP. Plexus Anesthesia, Perivascular Techniques of Brachial Plexus Block. Philadelphia, PA: Saunders; 1990:176. 3. Raj PP. Practical Management of Pain. Chicago, IL: Year Book; 1986:610. 4. Scott DB. Techniques of Regional Anaesthesia. East Norwalk, CT: Appleton & Lange; 1989:92. 5. Wildsmith JA, Armitage E. Principles and Practice of Regional Anesthesia. New York, NY: Churchill Livingstone; 1987:144. 6. Roch JJ, Sharrock NE, Neudachin L. Interscalene brachial plexus block for shoulder surgery: A proximal paresthesia is effective. Anesth Analg 1992;75:386-388. 7. Silverstein WB, Saiyed MU, Brown AR. Interscalene block with a nerve stimulator: A deltoid motor response is a satisfactory endpoint for successful block. Reg Anesth 2000;25:356-359. 8. Urmey WF, Stanton J, OBrien S, Tagariello V, Wickiewicz TL. Inability to consistently elicit a motor response following sensory paresthesia during interscalene block adminis-tration (abstract). Reg Anesth 1998;23:7. 342 Regional Anesthesia and Pain Medicine Vol. 25 No. 4 JulyAugust 2000

Regional Anesthesia for Arm, Wrist and Hand Surgery


Learning Objectives: 1. Describe the indications and contraindications for the various approaches to the brachial plexus appropriate for arm, wrist and hand surgery 2. Describe the relevant anatomy of the brachial plexus 3. Describe the techniques, local anesthetics and adjuvants used for performing a block of the brachial plexus for arm, wrist and hand surgery

e Infracla
pa010 Gl-0% MD, Renato Coluccia, MD, Albert0 cll,~ l-abio Gazzotti Tassi, MD, Vincenzo L. Indrizzi, MD,

The infraclavicular brachial plexus block is still an underused technique for regional anesthesia of the upper limb, but represents a reliable and safe approach for surgery of the hand, the forearm, the elbow, and the antecubital fossa, also involving the musculocutaneous nerve. This report intends to describe, as well as the anatomical evidence, an infraclavicular technique modified by Grossi, in which the arm is adducted or in a rest position. Vertical direction of the needle and electrical nerve stimulator with insulated needle is required. An historical review of infraclavicular blocks and their relation with other approaches are reported. Evident advantages are represented by compliance of the patient, tourniquet tolerance, usefulness of this approach to place a catheter, an alternative to the axillary approach in presence of joint stiffness or ankylosis, fractures of the limb, local infection or scars, and previous axillary lymphoadenectomy. The possible complications are related to pleural or vascular puncture. No impairment of the respiratory function or involvement of the phrenic nerve is reported. Copyright 0 1999 by W.B. Saunders Company

nfraclavicular brachial plexus block (BPB) is an underused technique for regional anesthesia of the upper limb, but is an effective alternative to the axillary block when this is not suitable for specific purposes. This review describes a new technique as well as relevant anatomy of the brachial plexus and illustrates the safety and indications of this approach to the brachial plexus. Compliance of the patient, facility in catheter placement, and advantages over the axillary approach are discussed. The possible complications related to pleural or vascular puncture and the influence on respiratory function are also outlined. A lower incidence of pneumothorax and no impairment of respiratory function are reported.

Historical

Consideration

In this review we describe a modified (by Grossi) infraclavicular approach to BPB, which can produce excellent anesthesia of the whole arm up to the proximal third of the humerus, is safer than the supraclavicular approach, and more complete than the axillary approach. Its advantages inclucle the ability to perform the block with the patients arm in a resting position, minimization of the risk of pneumothorax, ease of securing a brachial plexus catheter
From Servizio di Anestesia Rianimazione e Terapia del Dolore-lstituto Ortopedico Gaetano Pini, Milano, Italy; Servizio di Anestesia e Rianimazione 2 Azienda Ospedaliera Poticlinico di Modena, Modena, Italy; and Scuola di Specializzazione in Anestesia e Rianimazione dell Universita di Modena e Reggio Emilia, Modena, Italy. Address reprint requests to Pa010 Grossi, MD, Servizio di Anestesia, lstituto Ortopedico Gaetano Pini, Piazza C. Ferrari 1, 20122 Milano, Italy.
Copyright o 1999 by W.B. Saunders Company 1084-208X/99/0304-00041$1 0.0010

and fixing it to the chest wall, and the inclusion of the musculocutaneous nerve using a single injection. The infraclavicular approach to the BPB is today an underused technique but it has been described already in the beginning of this century. In the technique described by Bazy, the needle was directed medially from the midpoint of the clavicle to the Chassaignacs tubercle. It was a compromise between Hirschels axillary and Kulenkampffs supraclavicular approaches,2,3 attempting to deposit the anesthetic solution in the same place, that is just above the first rib along the line of anesthesia at the level of the trunks of the plexus. Afterwards, many techniques with lightly different landmarks and points of introduction were then described by Balog,4 Babitzky,5 Kin1,6 and others,7 with an axillary or a clavicular direction of the needle. All these techniques have some advantages and many disadvantages over the supraclavicular or axillary approach. Clinically, the danger of pneumothorax remained unchanged, limiting the infraclavicular approach in popularity for several years. In 1973 Raj8 described a new approach to the brachial plexus in the infraclavicular region. In this technique the patient lies supine, his head is turned away from the arm to be blocked and his arm is abducted to 90. The landmarks are the whole length of the clavicle, the subclavian artery pulsation under the clavicle (this landmark usually is at the midpoint of the clavicle), the brachial artery in the arm, the C6 tubercle on the same side, and the line from the C6 tubercle to the brachial artery in the arm, passing through the midpoint of the clavicle. An 89 mm 22-gauge (G) unsheathed spinal needle connected to an electrical nerve stimulator (ENS) is introduced through the skin 2.5-cm under the midpoint of the clavicle (or where the subclavian artery pulsation is palpated before disappearing under the clavicle) and is directed laterally toward the brachial artery. The needle is advanced at an angle of 45 to the skin and when its tip is in the close proximity to nerve fibers of the brachial plexus we can observe arm movements, electricall> provoked. An injection of 30 to 40 mL of local anesthetic solution followed. Rajs technique differs from previous infraclavicular techniques with regard to needle introduction, which is more medial and directed laterally from the point of entry so that the needle is always outside of the thoracic cavity Thus there is no real danger of pneumothorax. Otherwise, there are also some disadvantages. First of all, the needle is advanced blindly with respect to the vessels and the possibility of vascular puncture is certainly greater \vith this technique -than with other approaches. Secondly, because a long needle is required to penetrate both the pectoralis major and minor muscles and because multiple attempts are frequently necessary, the patient acceptance of this procedure could be less; in obese patients the needle could be too short and it is common to puncture the neurovascular sheath anteriorly and posterinrl\r urith 1 nnccihlP
cinnifir.-ant lncc af +ho In,--I n-oc-*Ln*;r

inside the bundle. The arm is abducted to 9O, as in the axillary approach, excluding all the cases in which it is not possible to do it. Finally, due to the needle being directed laterally, the spread of the injected solution is directed laterally, If the tip of the needle is lateral to the coracoid process, the majority of the injected solution may therefore miss the musculocutaneous and axillary nerves and this technique may simulate an axillary block that is performed higher in the axillary perivascular compartment. In I977 Sims suggested new landmarks in the effort to overcome some of the disadvantages of Rajs technique, making the infraclavicular BPB easier (also in obese patients) and using a shorter needle. The landmarks are the inferior border of the clavicle and the coracoid process of the scapula. The index finger should be placed in the groove between these two structures; advancing medially and inferiorly, it will fall into a depression within the superior portion of the major pectoralis muscle (inferiorly and medially), the coracoid process of the scapula (laterally), and the clavicle (superiorly). A 40-mm needle is introduced at this point and advanced inferiorly, laterally, and posteriorly toward the apex of the axilla. Usually, the plexus is located 2 to 3 cm beneath the skin. In 1981, Whiffler proposed the coracoid block in which the injection site is not far from that proposed by Sims, but the technique of injection is completely different. The patient lies supine with the head turned away from the arm to be blocked; the shoulder is depressed and the arm abducted to 45 from the chest wall. Once the coracoid process is identified, it is possible to estimate the depth of the injection by palpating with the index finger the axillary arterial pulse as high as possible in the axilla, and placing the thumb of the same hand on the anterior surface of the chest wall over the site at which the index finger palpates the artery This point usually lies in the deltopectoral groove. The needle is inserted with a right angle to the skin, on a line marked between the point in which the subclavian artery pulsation disappears under the clavicle and the projection on the anterior surface of the chest wall of the axillary arterial pulsation, just inferiorly and medially to the coracoid process, to the depth estimated as indicated above. After an initial injection of I2 mL of local anesthetic, the needle is withdrawn 1 cm and a second similar injection is made (in muscular individuals a third injection of 12 mL is required after withdrawing the needle 1 more centimeter). This technique does not require the use of an ENS because the objective is not to make an injection inside the neurovascular sheath but to lay down a wall of anesthesia through which the plexus must pass.
injected

Anatomy
The brachial plexus is formed by the union of the anterior primary divisions of 0, C6, C7, C8, and Tl spinal nerves with frequent contribution of C4 and T2.12,13 It starts from the vertebral column, runs in the groove between the anterior and the middle scalene muscles, passes between the clavicle and the first rib where it is joined by the subclavian artery, Lvhich runs deep to the anterior scalene muscle, and proceedingunder the pectoralis minor muscle insertion on the coracoid process enters the upper limb in the axilla. In its course from the intervertebral foramina to the arm the plexus is composed consecutively of roots, trunks, divisions, cords, and terminal nerves, formed through a complex process of combining and dividing.

After leaving the intervertebral foramina, the roots of the fifth, sixth, seventh, and eighth cervical nerves pass behind the vertebral artery and travel laterally in the gutters formed by the superior surfaces of the anterior and posterior tubercles of the corresponding cervical transverse processes. At the distal end of the transverse processes the roots descend in front of the middle scalene muscle toward the first rib, above which they fuse with the root of the first thoracic nerve, which passes upward and laterally in front of the neck of the first rib and behind the pleura over the apex of the lung, to form the three trunks of the plexus. In its passage from the cervical transverse processes to the first rib, the plexus, first as roots and then as trunks, is sandwiched between the anterior and middle scalene muscles and so invested by the fascia of those muscles that limit the interscalene space. It is really the fascia covering the scalene muscles, derived from the prevertebral fascia, which constitutes the sheath of the brachial plexus. As the three trunks, named superior (formed by the union of CS, C6 roots), middle (C7), and inferior (C8, Tl), cross the first rib they are arranged one on top of the other vertically, as the name implies, and are joined inside the sheath by the subclavian artery to form the subclavian perivascular space. Not infrequently, the inferior trunk of the brachial plexus gets trapped behind, and even under the artery, which could make a barrier to the diffusion of local anesthetic solution injected higher in the interscalene space. After the trunks have passed over the first rib and under the clavicle, at about the upper border of the clavicle, each trunk divides into an anterior and posterior division. As the plexus emerges from under the midpoint of the clavicle in the infraclavicular region of the axilla, the fibers of the six divisions recombine to form the three cords of the plexus: medial, lateral, and posterior. They are surrounded by the four walls limiting the axilla: pectoralis major and minor muscles form the anterior; the subscapularis, teres major, and latissimus dorsi muscles complete the posterior wall; the medial wail is made by the first four ribs of the chest wall; and the lateral wall is formed by the medial side of humerus head and by glenoid process of the scapula. In passing under the clavicle, the subclavian artery becomes the axillary artery and lies central to the three cords that are not really medial, lateral, and posterior to the axillary artery until they pass behind the pectoralis minor muscle. Within this space the cords gradually rotate around the artery until, in the second portion of the artery, their position become truly medial, lateral, and posterior to the artery. As it passes over the first rib and under the clavicle, the subclavian Irein, in becoming the axillary vein. joins the neurovascular bundle that takes name of axillar) fascia, (extension of the prevertebral fascia). l4 It is approximately at the lateral edge of the pectoralis minor muscle that the cords divide into the major terminal nerves The lateral and medial cords give off as their branches tha lateral and medial heads of the median nerve and then the medial cord continues as the ulnar nerve and the lateral cord a: the musculocutaneous nerve; the posterior cord gives off the axillary nerve as major branch and then continues as the radia nerve. Only the median, radial, ulnar, and medial antebrachia. cutaneous (a medial cord secondary branch) nerves with the brachial artery and vein lie within the axillary sheath at the level at which the axillaq block is performed. It is ver) important to emphasize that the musculocutaneous, the axil, lary, and the medial brachial cutaneous nerves are not stil

,Ircrlvarcli the musculocutaneous nerve enters the coracobra,,,,.11,, ,,,,,~c~e and descends the arm between the biceps and I,,,,~ t,,l.tlls muscle. The intercostobrachial nerve (T2) travil<. 1, I,.,r.,llel to but always outside the axillary sheath. 1 he brachial plexus and the neurovascular bundle terminate n the upper arm sandwiched bet\veen the coracobrachialis, the h(>rt head of biceps, the long head of triceps muscles, and the ,t>;lc/, the neck, and shaft of the humerus. l./lc pcrincural compartment of brachial plexus can be ,,tcxred at any level from the interscalene to the axillar) I~rl\zcular space, along the line of anesthesia, and the extent f anesthesia will depend on the level and the volume of ncsthetic solution injected over the roots, trunks, cords, or eriphcral branches.

3. 4.

5.

Technical Aspects of the Infraclavicular Approach to the Brachial Plexus Block


a

6.

he patient lies supine with his head turned away from the arm I be blocked. Whereas according to Rajs original technique le arm should be abducted 90 from the chest wall and the hysician should stand on the opposite side from the arm to be locked, in our approach (technique modified by Grossi) the -m is adducted or even abducted but not over 45 from the lest wall. The anesthetist (right-handed) stands beside the ltient between the head and the right shoulder if the arm to 5 operated upon is the right, or behind the head of the patient tr a left-sided block; otherwise in the opposite position if the aerator is left-handed. &mnt-hs mdmarks (all the landmarks should be traced with a skin pen 1 the skin, to stress the line of anesthesia): (I) the ChassaigKS tubercle, corresponding to anterior tubercle of sixth :rvical vertebra, determined by extending a line laterally from !e cricoid cartilage, is identified and marked behind the teral border of the sternocleidomastoid muscle, on the same rck side; (2) the whole length of the clavicle and particularly e midpoint is marked after palpation; (3) the coracoid ocess of the scapula and the lateral border of the chest wall e identified and marked in the infraclavicular region. A point 1 the skin about 2 cm medial and about 2 cm caudad to the tip the coracoid process identifies the brachial plexus below; ) the axillary artery pulse in the axilla is marked, ideally the )int where the pulse disappears under the pectoralis major uscle; and (5) the line of anesthesia, passing through the 2vicle midpoint and running between the coracoid process Id the lateral border of the chest wall, is drawn from the lassaignacs tubercle to the axillary arterial pulse point. -ocedur-e The ground electrode of an ENS is attached to the opposite shoulder. :. The exploring hand moves along the line of anesthesia to the point between the coracoid process and the chest wall. Advancing the fingers inferiorly and medially from the coracoid process, they will fall into a depression bordered
FRACLAVICULAR BRACHIAL PLEXUS BLOCK

7.

8.

9.

minor muscle and superficial to the thinnest part of the pectoralis major lnusc~c, ahout z cm mediall\r I alld ilbOUt 2 cm inferiorly to the tip of the coracoid process of the scapula. .After skin cleaning, a skin Lvhcal is raised just inferior and medial to the coracoid process along the line of anesthesia. AfterLvards a 22-G 12O-mm sheathed needle connected to the ENS and to the ground electrode on the shoulder is introduced through the skin bvheal perpendicular to the line of anesthesia and advanced posteriorly> inferiorly, and laterally at an angle of 0 to 15 toward the axilla (Fig 1). At no time should the needle be directed medially or to\vnrcl the chest wall, the lung, or attempt to reach the pcriostium of the osseous structures. The average depth of needle insertion required to reach the brachial plexus is 5.1 cm (2.25 to 7.75 cm in men and 2.25 to 6.5 cm in women) with the arm abducted not over 45. Greater abduction of the arm will make the performance of the technique easier because the brachial plexus depth is reduced and the cords are more spread, but this implies a normal joint functionality. Our modified technique is particularly useful and advantageous ivhen the abduction of the arm is limited or painful (.joint stiffness or ankylosis, fractures, etc). As the sheathed needle connected to the ENS (delivering a 0.5 mA and 2 Hz stimulus) approaches the cords of the brachial plexus, movements (twitch) of the muscles supplied by those fibers will occur. Flexion or extension of the elbow, wrist, or fingers and external or internal rotation of the forearm confirms that the needle point is in close proximity to nerve fibers of the brachial plexus. The needle should be advanced slowly until the maximum muscle movements are observed. T\vo different twitches should be elicited to improve the success rate of this approach. I6 The twitch of the muscles supplied by the cord, the sensitive zone of which is interested by the surgical treatment, represents the best t\vitch. Holding the needle in that position, 2 mL of anesthetic solution are injected. If the needle is located correctly on
pmrdiS

Fig 1. Vertical introduction of the 120-mm needle, along the anesthetic line, 2 cm below the coracoid process; the arm of the patient is adducted to the body.

219

the nerve fibers there will be an immediate loss of prcyiously observed muscle movements. If not, the needle may have been pushed through the nerve and it should be withdrawn slightly. If acute pain is elicited by the injection, the needle could have been advanced inside the nerve and it should be withdrawn immediately. 1 0 . ,4fter careful aspiration, the volume of local anesthetic solution is injected at that site. The needle may then be removed from the patient. 11. After the deposition of the anesthetic solution in the infraclavicular region, anesthesia develops from the top down in 10 to 30 minutes (onset time depending on the anesthetic solution). The extent of anesthesia is from the hand to the medium part of the humerus, including the sensitive area supplied by the musculocutaneaus nerve. There is not involvement of the phrenic nerve. As for the axillary approach to the brachial plexus, the shoulder and the clavicle are not involved in the block.

Fig 3. This approach is ideal for catheterization, especially when a long period of analgesia is required, as for rehabilitation of posttraumatic ankylosis of the elbow in children.

Drugs Volume and Dosage


The neurovascular bundle in the infraclavicular region is very compliant, thus it is possible to inject a large volume of anesthetic solution: 30 to 40 IIIL (depending on the patient body weight). A contrast medium injected inside the brachial plexus sheath in the infraclavicular region is observed sprcading from the axilla to the inferior border of the clavicle (Fig 2). The anesthetic solutions used are: mepivacaine 1.5S (30 to 40 mL, max 600 mg). bupivacaine 0.25% to 0.5% (30 to 40 ml., max 150 mg), or ropivacainc 0.5% lo 0.75% (30 to 40 mL, max 225 mg).

Indications
The infraclavicular brachial plexus block is indicated for the surgery of the hand, forearm, and elbow. This approach is also very effective for postoperative or emergency analgesia and ideal for long-term catheter placement (Fig 3), more cffcctive than the axillary approach in which the movement and dislocation of the catheter are easier and the infectious risk increased. It is also indicated for all the conditions in which the axillary block is difficult to perform: shoulder ankylosis or stiffness, upper limb fractures, previous lymphadcnectomy of the axilla, and scars or local infection, because it is possible to perform it with the arm adducted.

Another important advantage of the infraclavicular approach is the large analgesia extent, comprehending also the area supplied by the musculocutaneous and axillary nerves, so the patient can better tolerate the placement of a tourniquet at the proximal extremity of the arm. The performance of the block is easy and safe because the risk of pneumothorax is just theoretical and it can not provoke paralysis of the phrenic nerve.* Moreover, patient comfort is optimal, even in the case of humeral fracture and in young people. Psycl~ologically, patients tolerate this kind of approach better than the axillary or the intcrscalene. Furthermore the risk of infection is very love. In obese patients, because of the difficulty to identify the landmarks, localization of the correct site of injection may be more difficult.

Complications and Limitations


Our modified technique appears safe and rarely complicated. Nevertheless, an anesthetist experienced in regional anesthesia technique is necessary If the site of puncture and the direction of the needle are correct the worst complication, pneumothorax, will be avoided. Moreover, there is no involvement of the phrenic ncrvve, thus no danger of respiratory function impairment. Vascular puncture (Fig 4) and, eventually, hematoma

Fig 4. Vascular puncture is a possible complication of the


infrarlauirlllar annrnnrh ri~w to the lame size o f v e s s e l s .

.,,mI,C,rcCf lvith o t h e r t e c h n i q u e s (inttrscalene or axillary ti,p~~~,,c~,~ ;\s \vith other approaches, intravascular injection is ,, ,,;/)I(. .~(lii ~1 I t 1s w-) important to perform an aspiration i ,[ /)L~,C,rt Injecting. The ust of ENS may prevent neural I.~,,,,~~ Lifter traumatic injuries determined by the needle or ntr.m~urL31 injection. 1 ,,nltrltlc,ns are the same as with the other approaches:i(~~~,~~,pcr~~tI\t patient, coagulopathies, preexisting neural pa/,,,l~,~l~~s, previous adverse reaction to a local anesthetic, local infection, and uncorrected metabolic derangeII ,\.>tetnIc llcnts.

Conclusions
.he infraclavicular approach to the brachial plexus is similar to thcr types of upper-extremity sheath blocks, such as the uiilar): the supraclavicular and the interscalcne, in which an mount of local anesthetic agent sufficient to fill the fascial leath is in.jected to bathe all the nerves contained. Blocking re plexus through the infraclavicular approach has advantages ,hen compared with other approaches.- The tip of the ccdle is pointed away from the lung, avoiding the complicaons of supraclavicular (pneumothorax) and interscalcne rachial plexus block (injection into the carotid or vertebral -teries, the jugular vein, into subarachnoid or epidural space, 4renic nerve, vagus or stellate ganglion). The injecLion Ipears to be less distressing for the patient than the supraclacular approach. This approach offers all the advantages of the tpraclavicular approach. Hobvever, the level of the block is wer than that obtained from a supraclavicular block, but the vel of anesthesia is higher than that obtained from the :illary block. It provides a complete, effective, and safe lesthesia of the whole arm and permits surgery of the hand, rearm, elbow, and antecubital fossa. Moreover patients requirg prolonged analgesia or sympathectomy of the upper .tremity for postoperative analgesia or complex regional pain ndromes may be especially eligible for infraclavicular cathzr placement because this region is ideal for securing the theter to the anterior chest wall. The advantages of the infraclavicular block techniques ajs, Sims, Whifflerh, and Grossis), that make these apoaches a safe alternative to the axillary brachial plexus block I: aseptic field, ability to block the musculocutaneous nerve the brachial plexus using a single injection, better comfort _ the patient, possibility to apply a constrictive tourniquet thout discomfort, ease of securing a continuous brachial zxus catheter to the chest wall, minimization of the pneumo3rax risk, and no impairment of respiratory function. The Grossi technique differs from previous infraclavicular :hniques in the arm position, which is adducted or abducted 0/45, the needle is introduced medially and inferiorly to 2 coracoid process, perpendicular to the line of anesthesia, d advanced posteriorly and slightly inferiorly and laterally at angle of 0/15 towards the axilla so that the needle is always tside of the thoracic cavity and there is no real danger of eumothorax. Moreover the almost direct posterior insertion the needle will make a perpendicular contact with the cords the brachial plexus in which they surround the second part the axillary artery. This avoids spread of the injected solution orally to the neurovascular bundle, thus missing the muscuutaneous and axillary nerves, as it happens when an axillary
RACLAVICULAR BRACHIAL PLEXUS BLOCK

0ur description of the vertical approach to infra&vicu[ar ~\CXUS block tnay provide advantages over the existing techniques (ie, Rajs approach): the ability to perform the block Lvith the arm in a rest position, Lvhich allo\vs f o r patient comfort. This tnakes it an effective and safe alternative lvhen asillary block is impossible because of joint stiffness or ankylosis, local scars or infection. upper limb fractures, or asillary lirnphoadencctom): Furthermore because of the presence of a consistent, palpable bon)~ landmark is likely to be easily, understood, taught, and performed. 011 the other hand, Lvith this approach the needle is a&mcd blindly uith respect to the vessels and vascular puncture with possible hematoma formation. A possible disadvantage may be the inability to externally compress the source of hematoma. Proper technique and careful avoidance of the block in patients with coagulopathics uxmlcl limit the occurrence of this problem.
hr:~~hial

References
1. Bazy L, Pauchet V, Sourdat P, et al (eds): LAnesthesie Regionale. P a r i s , G Doin e t C i e , 1 9 1 7 , p p 2 2 2 - 2 2 5 2 . Hirschel G: Anesthesia of the brachial plexus for operations on the upper extremity. Mijnchen Med Wochenschr 58:1555-1556,191l 3 . Kulenkampff D: Anesthesia of the brachial plexus. Zentralbl Chir 38:1337-1350, 1911 4 . Balog A: Conduction anesthesia of the infraclavicular portion of the brachial plexus. Zentralbl Chir 51:1563-1564, 1924 5 . Babitzki P: A new way of anesthetizing the brachial plexus. Zentralbl Chir45:215-217, 1918 6 . Kim MH: Anesthesia of the Brachial Plexus via the lnfraclavicular Groove. Zentralbl Chir 55:1423, 1928 7 . Labat G ( e d ) : R e g i o n a l A n e s t h e s i a . P h i l a d e l p h i a , P A , S a u n d e r s , 1 9 2 8 8 . Raj PP, Montgomery SJ, Nettles D, et al: lnfraclavicular brachial plexus block: A new approach. Anesth Analg 52:897-904, 1973 9 . Montgomery SJ, Raj PP, Nettles D, et al: The use of the nerve stimulator with standard unsheathed needles in nerve blockade. A n e s t h A n a l g 52:827-831, 1973 10. Sims JK: A modification of landmarks for infraclavicular approach to b r a c h i a l p l e x u s b l o c k . A n e s t h A n a l g 56:554-555, 1977 11. Whiffler K: Coracoid block: A safe and easy technique. Br J Anaesth 53:845-848, 1981 12. Winnie AP (ed): Perivascular techniques of brachial plexus block, in P l e x u s A n e s t h e s i a , vol 1 . P h i l a d e l p h i a , P A , S a u n d e r s , 1 9 9 0 13. P e r n k o p f E ( e d ) : A t l a s o f T o p o g r a p h i c a l a n d A p p l i e d H u m a n A n a t o m y . Philadelphia, PA, Saunders, 1980 14. Thompson GE, Rorie DK: Functional anatomy of the brachial plexus sheaths. Anesthesiology 59:117-l 22, 1983 15. Wilson JL, Brown DL, Wong GY, et al: lnfraclavicular brachial plexus block: Parasaggital anatomy important to the coracoid technique. A n e s t h A n a l g 87:870-873, 1998 16. Fitzgibbon DR, Debs AD, Erjavec MK: Selective musculocutaneous nerve block and infraclavicular brachial plexus anesthesia. Case report. Reg Anesth 20:239-241, 1995 17. Mehrkens HH, Geiger PK: Continuous brachial plexus blockade via the vertical infraclavicular approach. Anesthesia 53:19-20, 1998 (SUPPI 2) 18. Grossi P: Respiratory effects of the infraclavicular brachial plexus block. Abstract handbook, ISRAAuckland NZ, 187-188, 1996 19. N e u b u r g e r M , K a i s e r H , R e m b o l d - S c h u s t e r I , e t a l : V e r t i c a l infraclavicular brachial-plexus blockade. A clinical study of reliability of a new method for plexus anesthesia of the upper extremity. Anesthesist 47:595-599, 1998 20. Kilka HG, Geiger P, Mehrkens HH: lnfraclavicular vertical brachial plexus blockade: A new technique of regional anesthesia. Anesthesist 44:339-344, 1995 21. Brown DL: Brachial plexus anesthesia: An analysis of options. Yale J B i o l M e d 66:415-431, 1993

221

REGIONAL ANESTHESIA
SECTION EDITOR DENISE J. WEDEL

AND

PAIN MEDICINE

A Magnetic Resonance Imaging Study of Modifications to the Infraclavicular Brachial Plexus Block
ivind Klaastad, MD*, Finn G. Lilles, MD, Jan S. Rtnes, Harald Breivik, MD, PhD, and Erik Fosse, MD, PhD
MD, PhD, *Department of Anesthesiology, The National Hospital Orthopedic Centre; and The Interventional Centre and Department of Anesthesiology, The National Hospital, Oslo, Norway

A previously described infraclavicular brachial plexus block may be modified by using a more lateral needle insertion point, while the patient abducts the arm 45 or 90. In performing the modified block on patients abducting 45, we often had problems finding the cords of the brachial plexus. Therefore, we designed an anatomic study to describe the ability of the recommended needle direction to consistently reach the cords. Additionally, we assessed the risk of penetrating the pleura by the needle. Magnetic resonance images were obtained in 10 volunteers. From these images, a virtual reality model of each volunteer was created, allowing precise positioning of a simulated needle according to

the modified block, without exposing the volunteers to actual needle placement. In both arm positions, the recommended needle angle of 45 to the skin was too shallow to reach a defined target on the cords. Comparing the two arm positions, target precision and risk of contacting the pleura were more favorable with the greater arm abduction. We conclude that when the arm is abducted to 90, a 65-needle angle to the skin appears optimal for contacting the cords, still with a minimal risk of penetrating the pleura. However, this needs to be confirmed by a clinical study. (Anesth Analg 2000;91:929 33)

nfraclavicular brachial plexus blocks aim at the cords of the brachial plexus and have been designed to obtain complete nerve block of the upper extremity while minimizing the risk of pneumothorax (1 6). In a previous magnetic resonance (MR) study of the infraclavicular block described by Raj et al. (Rajs block) (1), we proposed a more lateral needle insertion point (7). This would bring the needle closer to the cords and farther away from the pleura. Our group had been introduced to such a modification of Rajs block during a workshop at a meeting in 1993. Regrettably, this method has not been published. We will refer to it as the lateral approach. According to our recollection, it was performed as follows (Figs. 1, 2B, 3B): The supine patient abducts the arm to 45. Two arterial points are palpated and marked where the subclavian artery dips under the superior border of the clavicle (alternatively at the base of the interscalene cleft) and approximately where the brachial artery crosses the lateral border of the pectoralis major. A line between these points is
Supported by the National Hospital Orthopedic Centre. Accepted for publication April 10, 2000. Address correspondence and reprint requests to Dr. . Klaastad, The National Hospital Orthopedic Centre, Department of Anesthesiology, Trondheimsveien 132, 0570 Oslo, Norway.
2000 by the International Anesthesia Research Society 0003-2999/00

drawn. The needle insertion point is on this line, at a radial distance of 2.5 cm from the lines intersection with the inferior border of the clavicle. The needle is directed laterally along the line while kept at an angle of 45 to the skin. A nerve stimulator aids in exact positioning of the needle. The lateral approach may also be performed with the patient abducting the arm 90 (workshop 1996, San Diego). This position brings the needle insertion point more cephalad and the needle course more lateral than with lesser degrees of abduction. Theoretically, this should reduce the risk of pneumothorax. This variant has not been published. The lateral approach differs from Rajs block only regarding the point of needle insertion (Fig. 1). By using Rajs block, the patient abducts the arm, preferably to 90. The needle insertion point is 2.5 cm below the inferior border of the clavicle, on a paramedian line through the point at which the subclavian artery is palpated dipping under the clavicle, or on a paramedian line through the midpoint of the clavicle. During the first 28 months after the 1993 meeting, we have tried the lateral approach in 161 patients with the arm abducted to 45. Frequently, we had to redirect the needle to find the nerves, and in 18 patients (11%), we finally discontinued the method. No patients demonstrated clinical signs of pneumothorax.
Anesth Analg 2000;91:92933

929

930

REGIONAL ANESTHESIA AND PAIN MEDICINE KLAASTAD ET AL. MRI AND INFRACLAVICULAR BRACHIAL PLEXUS BLOCKS

ANESTH ANALG 2000;91:929 33

Figure 1. Front view of infraclavicular anatomy, right side, arm 90 abducted (Volunteer 1). Merging segmented magnetic resonance images with a correlated surface picture of the identical person created the picture. The hand-held white and black needles demonstrate point of needle insertion and needle direction by using the lateral approach for infraclavicular brachial plexus block and the infraclavicular brachial plexus block described by Raj et al. (1), respectively. The subclavian/axillary artery (white) is not marked, but is located cephalad to the corresponding vein (V). The cords (gray), not marked, surround the artery longitudinally. Only parts of the pectoralis major muscle (pma, dark) and the pectoralis minor muscle (pmi, gray) are presented, caudad to the axillary vein. cl clavicle, C1 first costa, C2 second costa, cp coracoid process, c.hu caput humeri.

Because of the difficulties with the technique, we questioned if the recommended needle angle to the skin guides the needle close enough to the cords and therefore initiated the present study. The primary aim of this anatomical study was to investigate the ability of the lateral approach to reach the brachial plexus by using a variety of needle angles to the skin. Additionally, we wanted to confirm the clinical impression of a decreased risk of pneumothorax. We were also interested in comparing the results of the method with the arm abducted 45 and 90. For the study, we used MR imaging because it easily demonstrates the brachial plexus (5,79).

brachial artery (its midaxis) and the lateral edge of the pectoralis major muscle. The plane between these anteroposterior lines was the first of two planes defining the recommended needle direction (the needle trajectory). A third anteroposterior line was determined in this plane, 2.5 cm from the inferior border of the clavicle. The point at which this line hit the chest surface defined the needle insertion point. The target was determined as in our first study: Through the point at which the perpendicular line from the most anterocaudad point of the coracoid process hit the first plane defining the needle trajectory, a sagittal (paramedian) plane was constructed. The middle point of all nerve structures around the artery in this plane defined the target. The second plane, defining the needle trajectory, was perpendicular to the axial (transverse) plane (Fig. 3B), went through the needle insertion point, and had a 45 medial angle to the coronal plane. The final position of the needle point was defined as where the needle trajectory hit the sagittal plane through the target. The needle trajectorys distance from the target was measured in coronal and axial planes. From these measurements, the true distance between the trajectory and the target was calculated and could be controlled by direct measurements in the sagittal plane through the target. The needle angle to the skin contacting the target was measured in the axial plane. Needle depths to the target and the final position of the needle point were calculated after measurements in axial and coronal planes. The trajectorys relation to the pleura was analyzed in the axial plane through the simulated needle insertion site, measuring the needle angle to touch the pleura. The results are presented as mean sd or mean (range). Students paired t-test was used to assess differences of the lateral approach by 90 and 45 arm abduction. Probability values 0.05 were considered significant.

Results Methods
The protocol, approved by the regional ethical committee, was similar to our first study and used the same 10 healthy volunteers (7). A needle was never inserted in the volunteers. MR images were taken with arm abduction at 45 and 90, and a virtual reality model of each volunteers infraclavicular region was created. In the model, the needle insertion point was determined after marking the position of two anteroposterior lines in the coronal (frontal) plane (Fig. 2B). The first line abutted the superior border of the clavicle and went through the midaxis of the subclavian artery. The second line went through the junction of the The volunteers, five women and five men, were 30 9 (24 51) yr old with a height of 175 11 (160 193) cm and a weight of 68 16 (50 102) kg. The deviation of the simulated needle trajectory from the target was great, approximately 2 cm, with both 45 and 90 abduction of the arm (Table 1). In both arm positions, the needle angle to the skin (in the axial plane) necessary to contact the target was practically identical (mean, 68 and 67) and considerably greater than the 45-needle angle recommended. The precision was much better in the coronal than in the axial plane, particularly with 90 abduction, having a near 100% precision.

ANESTH ANALG 2000;91:929 33

REGIONAL ANESTHESIA AND PAIN MEDICINE KLAASTAD ET AL. MRI AND INFRACLAVICULAR BRACHIAL PLEXUS BLOCKS

931

Figure 2. A, Coronal magnetic resonance image, right side, arm 90 abducted (Volunteer 1). B, Drawing based on the coronal image in A. Right infraclavicular region, arm 90 abducted (Volunteer 1). Illustrations of the lateral approach for infraclavicular brachial plexus block in the coronal plane. Added to the figure are the complete projection of the clavicle (cl) and a large part of the lateral border of the pectoralis major muscle (pma). In the lower half of the picture is part of the right lung (pu). The artery (A) is depicted in its complete length, whereas only a shorter distal part of the corresponding vein (V) is seen, caudad to the artery. Part of the brachial plexus (pl) is marked black cephalad to the artery. The recommended needle direction (the needle trajectory) is defined by two planes, of which the first goes through A1 and A2, perpendicular on the coronal plane. In this volunteer, the needle trajectory did not deviate from the target in the coronal plane. Therefore, the final position of the needle point is identical to the position of the target, and the corrective angle to bring the needle point to the target is zero, in the coronal plane. The target on the cords is defined periarterially in the sagittal plane (not marked) through the point at which the perpendicular line from the most anterocaudad point of the coracoid process hit the first plane defining the needle trajectory. The posterior projection of the perpendicular line is indicated. cp coracoid process, ac acromion, c.hu caput humeri, I* the posterior projection of the needle insertion point, T* the anteroposterior projection of the target, A1 an anteroposterior line through the point at which the subclavian artery would have been palpated at the superior edge of the clavicle, A2 an anteroposterior line through the point at which the brachial artery would have been palpated at the lateral border of the pectoralis major muscle.

The needle depths were similar in both arm positions, approximately 4 cm when the simulated needle contacted the target. The needle angle to the skin necessary to touch the pleura was great in both arm positions, never 80, and distinctly greater with 90 arm abduction than with 45 abduction. With 90 arm abduction, also when applying the optimal needle angle to contact the target, the sector between the needle and the pleura was considerable, 39 (2854).

Discussion
The present noninvasive MR study of the lateral approach for infraclavicular brachial plexus block demonstrates that the deviation of the recommended needle direction from the target on the cords was great with both 45 and 90 arm abduction, mostly because the needle angle to the skin was too shallow. This

probably explains the difficulties we had locating the brachial plexus in patients. In both arm positions, the risk of contacting the pleura appeared minimal, confirming our clinical impression. An approximate doubling of the recommended 45 needle angle to the skin was required for the simulated needle to touch the pleura. We consider the lateral approach more favorable with 90 arm abduction than 45 because the needle trajectory was more precise in reaching the cords and had a larger gap to the pleura in the former position. With 90 arm abduction, the precision can be enhanced by increasing the needle angle to the skin from 45 to 65. One might prefer starting with an angle of 40 and, when necessary, increasing it in steps of 10 to a maximum of 80. The risk of penetrating the pleura would remain small within this angle range, provided

932

REGIONAL ANESTHESIA AND PAIN MEDICINE KLAASTAD ET AL. MRI AND INFRACLAVICULAR BRACHIAL PLEXUS BLOCKS

ANESTH ANALG 2000;91:929 33

Figure 3. A, Axial magnetic resonance image, right side, arm 90 abducted. Image through the point of needle insertion (Volunteer 1). B, Drawing based on the axial image in A, through the point of needle insertion (I). Right infraclavicular area, arm 90 abducted. Illustrations of the lateral approach for infraclavicular brachial plexus block in the axial plane. Through I, four lines are drawn, marked by the numbers 1, 2, 3, and 4. 1 the tangent to the skin, 2 the recommended 45 medial needle angle to the skin, 3 the needle angle to the skin necessary to contact the target; in this volunteer 74 and 4 the needle angle to the skin necessary to touch the pleura, 102 in this volunteer. The recommended needle direction (the needle trajectory) is defined by two planes, of which the second goes through Line 2, perpendicular to the axial plane. The angle difference between Lines 3 and 2 represents the deviation of the needle trajectory from the target, 74 45 29 in this volunteer. The angle between Lines 4 and 3 is the medial deviation of the simulated needle contacting the target necessary to touch the pleura, in this volunteer 102 74 28. Ant anterior, Post posterior, T* cephalad projection of the study defined target at the level of the cords, I-T* is the projection of the recommended needle direction (the needle trajectory) to the axial plane through I, F* cephalad projection of the final position of the needle point, pl cords of the brachial plexus, A and V cross sections of the artery and vein, respectively, pma the pectoralis major muscle, pmi the pectoralis minor muscle, pu the right lung, bro the right main bronchus, sc scapula, cl clavicle, C1 cross section of first costa, C2 cross section of second costa.

Table 1. Proximity of the Needle Trajectory to the Target and to the Pleura Target deviation (mm) Coronala Mean 45 abduction Mean 90 abduction Range 45 abduction Range 90 abduction P 83 33 313 07 0.033* Axialb 21 8 18 6 1036 529 0.239 Sagittal 24 7 19 6 1338 529 0.042* Target angle Axial 68 7 67 10 5577 4177 0.814 Needle depth (mm) Before 24 8 25 8 1342 1843 0.731 After 41 8 38 5 3053 3351 0.076 Pleura angle Axial 91 6 106 5 8198 95114 0.001* Pleura-target angle Axial 23 5 39 7 1630 2854 0.001*

Target deviation the distance in mm between the needle trajectory and the target, as seen in three different planes; the true deviation/distance is found in the sagittal plane through the target, Target angle the needle angle to the skin to touch the target (the optimal angle to the skin), in the axial plane through the point of needle insertion, Needle depth (mm)/Before and After the needle depth before and after redefining the optimal angle to the skin, Pleura angle the needle angle to the skin to touch the pleura, in the axial plane through the point of needle insertion, Pleura-Target angle the angle difference between the described pleura and target angles. a With 45 abduction of the arm, the simulated needle trajectory was caudad to the target in all volunteers. With 90 abduction, the trajectory was cephalad to or corresponding to the target in all volunteers except for Volunteer 3. He had a 1-mm caudad deviation of the simulated needle from the target. b With 45 abduction of the arm, the needle trajectory was anterior to the target in all volunteers. With 90 abduction, the trajectory was anterior to the target in all volunteers except for Volunteer 3. He had a 5-mm posterior deviation of the needle from the target. * Significant difference between the results in the two arm positions by using Students paired t-test, when P 0.05.

that the other details of the technique and the configuration of the thoracic cage are recognized. The selection of our target on the cords may be controversial. With a more proximal target on the cords, the needle angle to the skin necessary to hit this target would increase, increasing the risk of contacting the

pleura. A more distal target would reduce this angle, but could end in an area more easily reached by an axillary approach (10). Taken together, we think that our chosen target is appropriate. In conclusion, our MR study demonstrates that the 45-needle angle to the skin recommended by the

ANESTH ANALG 2000;91:929 33

REGIONAL ANESTHESIA AND PAIN MEDICINE KLAASTAD ET AL. MRI AND INFRACLAVICULAR BRACHIAL PLEXUS BLOCKS

933

lateral approach of the infraclavicular brachial plexus block is often too shallow to contact the cords of the brachial plexus, with both 45 and 90 arm abduction. By using 90 abduction, a 65-needle angle to the skin appears optimal, still with a minimal risk of penetrating the pleura. However, this needs to be confirmed by a clinical study.
For better understanding of the brachial plexus anatomy, parallel to our MRI studies, we performed human cadaver dissections on the brachial plexus. We thank Professor Per Brodal, at the Department of Anatomy, University for Oslo for encouraging discussions and cooperation by the dissections. We thank Terje Tillung (The Interventional Center) for processing the images and Per yvind Hvidsten (The Norwegian Defense Research Establishment) for developing the three-dimensional visualization software.

References
1. Raj PP, Montgomery SJ, Nettles D, Jenkins MT. Infraclavicular brachial plexus block: a new approach. Anesth Analg 1973;52: 897903.

2. Sims JK. A modification of landmarks for infraclavicular approach to brachial plexus block. Anesth Analg 1977;56:554 5. 3. Whiffler K. Coracoid block: a safe and easy technique. Br J Anaesth. 1981;53:845 8. 4. Kilka HG, Geiger P, Mehrkens HH. Die vertikale infraklavikulare blockade des plexus brachialis. Anaesthesist 1995;44: 339 44. 5. Wilson JL, Brown DL, Wong GY, et al. Infraclavicular brachial plexus block: parasagittal anatomy important to the coracoid technique. Anesth Analg 1998;87:870 3. 6. Salazar CH, Espinosa W. Infraclavicular brachial plexus block: variation in approach and results in 360 cases. Reg Anesth Pain Med 1999;24:411 6. 7. Klaastad , Lilles FG, Rtnes JS, et al. Magnetic resonance imaging demonstrates lack of precision in needle placement by the infraclavicular brachial plexus block described by Raj et al. [letter]. Anesth Analg 1999;88:593 8. 8. Posniak HV, Olson MC, Dudiak CM, et al. MR imaging of the brachial plexus. AJR Am J Roentgenol 1993;161:3739. 9. Brown DL, Cahill DR, Bridenbaugh LD. Supraclavicular nerve block: anatomic analysis of a method to prevent pneumothorax. Anesth Analg 1993;76:530 4. 10. Winnie AP. Guest discussion. Anesth Analg 1973;52:903 4.

1998 by International Anesthesia Research Society. Volume 87(4) October 1998 pp 870-873

Infraclavicular Brachial Plexus Block: Parasagittal Anatomy Important to the Coracoid Technique
[Regional Anesthesia And Pain Management] Wilson, Jack L. MD; Brown, David L. MD; Wong, Gilbert Y. MD; Ehman, Richard L. MD; Cahill, Donald R. PhD
Departments of (Wilson, Brown, Wong) Anesthesiology, (Ehman) Radiology, and (Cahill) Anatomy, Mayo Clinic, Rochester, Minnesota. Accepted for publication July 15, 1998. Address correspondence and reprint requests to J. L. Wilson, MD, Department of Anesthesiology, Mayo Clinic, 200 First St. SW, Rochester, MN 55906.

Abstract
Infraclavicular brachial plexus block is a technique well suited to prolonged continuous catheter use.We used a coracoid approach to this block to create an easily understood technique. We reviewed the magnetic resonance images of the brachial plexus from 20 male and 20 female patients. Using scout films, the parasagittal section 2 cm medial to the coracoid process was identified. Along this oblique section, we located a point approximately 2 cm caudad to the coracoid process on the skin of the anterior chest wall. From this point, we determined simulated needle direction to contact the neurovascular bundle and measured depth. At the skin entry site, the direct posterior insertion of a needle will make contact with the cords of the brachial plexus where they surround the second part of the axillary artery in all images. The mean (range) distance (depth along the needle shaft) from the skin to the anterior wall of the axillary artery was 4.24 +/- 1.49 cm (2.25-7.75 cm) in men and 4.01 +/- 1.29 cm (2.25-6.5 cm) in women. Hopefully, this study will facilitate the use of this block. Implications: We sought a consistent, palpable landmark for facilitation of the infraclavicular brachial plexus block. We used magnetic resonance images of the brachial plexus to determine the depth and needle orientation needed to contact the brachial plexus. Hopefully, this study will facilitate the use of this block. (Anesth Analg 1998;87:870-3)

Section Editor: Denise J. Wedel. The infraclavicular approach to brachial plexus block is an underused but effective technique. Anesthesiologists may opt for more familiar techniques of brachial plexus anesthesia, such as the axillary approach, given the common lack of experience with this technique and significant variation in infraclavicular anatomy among patients. Nevertheless, advantages of the infraclavicular approach include the ability to perform the block with the patient's arm in any position, avoidance of the neurovascular structures of the neck, minimization of the risk of pneumothorax, and ease of securing a continuous brachial plexus catheter to the chest wall at this site [1-4]. Magnetic resonance imaging (MRI) has emerged as the preferred radiological modality for studying the brachial plexus and the corresponding anatomy [5]. The purpose of this study was to use MRI and cadaver sections to define the anatomic measurements and variation relevant to the infraclavicular block to establish the orientation and depth of simulated needle placement required to reach the brachial plexus by using an infraclavicular/coracoid approach.

Methods
After obtaining institutional review board approval, we reviewed the oblique parasagittal T1-weighted magnetic resonance images of the brachial plexus from patients undergoing imaging for other reasons. The oblique parasagittal view is used routinely in our institution to obtain optimal intersection (90[degree sign]) with the brachial plexus. Patients with distorted brachial plexus anatomy from ma ss effect or postprocedural changes were not included. Included in the review were 20 male and 20 female patients imaged in the supine position with the arms adducted, simulating the usual position for infraclavicular/coracoid block. The mean (range) age of the patients was 53.6 +/- 15.6 yr (26-82 yr). Using scout films, we identified the parasagittal section 2 cm medial to the tip of the coracoid process. Along this oblique imaging section, we located a point approximately 2 cm caudad to the coracoid process on the skin of the anterior chest wall. From this point, we determined the simulated needle direction required to contact the anterior aspect of the axillary artery (neurovascular bundle) and measured the depth for each subject (Figure 1). A representative image used in the study with a corresponding line drawing is depicted in Figure 2. In addition to MRI studies, parasagittal cadaver sections were prepared to further identify the anatomy relevant to this infraclavicular/coracoid block (Figure 3).

Figure 1. Magnetic resonance imaging measurements for localization of the brachial plexus on oblique parasagittal sections.

Figure 2. Representative magnetic resonance image used for study measurements.

Figure 3. Sagittal cadaver section through the brachial plexus at the coracoid process level.

Results
At the point 2 cm medial and 2 cm caudad to the tip of the coracoid process, the direct posterior placement of a needle would contact the cords of the brachial plexus where they surround the second part of the axillary artery in all images (Figure 4). The distance from the skin to the anterior wall of the axillary artery was 4.24 +/- 1.49 cm (2.25-7.75 cm) in men and 4.01 +/- 1.29 cm (2.25-6.5 cm) in women. (Table 1)

Figure 4. Anatomic landmarks for the infraclavicular/coracoid block.

Table 1. Demographic Data

Discussion
Our description of the coracoid approach to infraclavicular brachial plexus block may provide advantages over existing techniques. Raj et al. [4] described an approach to infraclavicular block using lateral needle orientation to overcome the risk of pneumothorax inherent with blocks performed under the clavicle with the needle directed medially. Other techniques using lateral needle angulation or different landmarks for infraclavicular blocks have been described. The technique described by Sims [3] has a more medial and cephalad needle entry site with a inferior and lateral needle angulation. Whiffler's technique [6] uses a needle entry site that is most often inferior and medial to the coracoid process determined by palpation of vascular landmarks with the affected arm abducted and the relevant shoulder depressed. The needle direction, such as that we describe, is directly posterior. The depth of needle insertion required to reach the brachial plexus often requires the entire length of the needle (51 mm). The risk of penetrating the

thoracic cavity, as noted in the preliminary cadaver study, was zero with this method. Kilka et al. [7] studied 175 patients undergoing surgery of the upper limb and anesthetized them using an infraclavicular approach based on previous anatomic studies. They divided the distance between the fossa jugularis and the ventral process of the acromion into equal parts and inserted the needle under the clavicle at the midpoint. The needle was passed directly posterior. A nerve stimulator was used to obtain muscle contractions in the area to be operated on with a current <or=to0.3 mA. The success rate of the block (surgical anesthesia) was 94.8%. The remaining patients underwent general anesthesia but had complete blocks after sur gery. Venous puncture occurred in 18 patients (10.3%), and Horner's syndrome was noted in 12 patients (6.8%). No arterial or pleural injury was noted. The advantages of the coracoid/infraclavicular block are the ability to perform the block with the ipsilateral arm in any position and, more importantly, the presence of a consistent, palpable bony landmark. Additionally, the block is likely to be easily understood and taught because the needle insertion is directly posterior from the skin entry site. Other theoretic advantages common to other infraclavicular blocks include the ability to block the musculocutaneous nerve of the brachial plexus using a single injection, minimization of the risk of pneumothorax, and avoidance of neurovascular structures of the neck. Our study is limited with respect to measurements of distance of the needle tip from pulmonary tissue. This is because the magnetic resonance images are directed lateral to medial inferiorly (oblique). Nonetheless, our clinical experience has not resulted in a pneumothorax while using this approach. In some cases, fluoroscopy was used to facilitate placement of a continuous catheter. Using our coracoid landmarks, no needle was close to the lung in the anteroposterior views. Vascular puncture with hematoma formation is possible with an infraclavicular approach to the brachial plexus, similar to other blocks along the plexus. A disadvantage with our technique is the inability to externally compress the source of hematoma. Proper technique and careful avoidance of the block in patients with coagulopathies would limit the occurrence of this problem. Indications for this block are identical to those for an axillary approach and include surgery of the forearm and hand. Patients requiring prolonged analgesia or sympathectomy of the upper extremity for postoperative analgesia or complex regional pain syndromes may be especially suited for infraclavicular catheter placement because this site is ideal for effectively securing the catheter to the anterior chest wall [1,2]. In conclusion, using the infraclavicular/coracoid brachial plexus block in simulation, the brachial plexus would have been contacted in all subjects. The required depth of insertion varies with body habitus. This description of our infraclavicular/coracoid block and measurements related to the block may make anesthesiologists more comfortable in performing this technique.

REFERENCES
1. Wedel DJ. Peripheral nerve blocks. In: Wedel DJ, ed. Orthopedic anesthesia. New York, Churchill Livingstone, 1993:256-71. [Context Link] 2. Brown DL. Brachial plexus anesthesia: an analysis of options. Yale J Biol Med 1993:66:415-31. [Medline Link]
[Context Link]

3. Sims JK. A modification of landmarks for infraclavicular approach to brachial plexus block. Anesth Analg 1977;56:554-5. [Context Link] 4. Raj PP, Montgomery SJ, Nettles D, Jenkins MT. Infraclavicular brachial plexus block: a new approach. Anesth Analg 1973;52:897-904. [Medline Link] [Context Link] 5. Blair DN, Rapoport S, Sostman HD, Blair OC. Normal brachial plexus: MR imaging. Radiology 1987;165:763-7. [Medline Link] [Context Link] 6. Whiffler K. Coracoid block: a safe and easy technique. Br J Anaesth 1981;53:845-8. [Medline Link] [Context Link] 7. Kilka HG, Geiger P, Mehrkens HH. Infraclavicular vertical brachial plexus blockade: a new technique of regional anaesthesia. Anaesthesist 1995;44:339-44. [Medline Link] [Context Link]

Accession Number: 00000539-199810000-00023

Copyright (c) 2000-2001 Ovid Technologies, Inc. Version: rel4.3.0, SourceID: 1.5031.1.149

Supraclavicular Approaches to Brachial Anesthesi


Alon P Winnie, MD and Carlo D. France, MD

The concept of a continuous perineural and perivascular space surrounding the brachial plexus from roots to terminal nerves simplifies conduction anesthesia of the upper extremity. As with peridural techniques, the space may be entered at any level, and following the injection of a local anesthetic, the extent of anesthesia will depend on the volume of anesthetic and the level at which it is injected. The single injection techniques based on this concept have resulted in both an increased incidence of success and a decreased incidence of serious complications as compared to the multiple injection techniques used previously. The space above the clavicle, the interscalene space, is potentially just such a space; and it is only after the injection of a local anesthetic that it expands to become a true, fluid-filled compartment. Furthermore, this space is quite extensive in its vertical and horizontal axes but very narrow in its anteroposterior axis. It was for this reason that the original subclavian perivascular and interscalene techniques both called for the needle to be introduced in one of the long axes of the space, whereas the more recently introduced parascalene techniques all call for the needle to be inserted in the very narrow anteroposterior axis. As a result, the chance of the needle leaving the space during the performance of the block is minimized with the subclavian perivascular and interscalene techniques, whereas with the parascalene techniques the slightest movement of even the properly placed needle may cause it to leave the space, resulting in the injection of local anesthetic outside of the fascial compartment. In short, the subclavian perivascular and interscalene techniques of brachial plexus block can provide simple, safe, and effective anesthesia for all types of surgery on the upper extremity and shoulder. Copyright o 1997 by W.B. Saunders Company

lower- si.Y cervical vcrttbrae, ciescends parallel to the middle SC8lCliC niwxlc, and inserts on the scalenc tubcrcle on the first
r i b , i t s insertion being separated from t h a t o f t h e nIiddle

scalenc muscle by the suhclavian groove, through which rhe subclaviaii artery passes (Fig 2). As the prcvertebral fascia extends laterally it becomes the scalcnc fascia by splitting 10 in\mt the scaltnc muscles, nerves, and artery between them
(Fig 3).

little over thirty years ago the supraclavicular of brachial plexus ancsthcsia were siiiiplificd the introduction of single inijcction techniques. all have 3s their basis the f&t that the hrachial plexus is
b y a n exrcnsion o f the Ixwcrtebral fascia from

techniques greatly by of which


en~~lopcd

The roots of the brachial plexus emerge from the intcrvertebra1 foramina and travel laterally in the trough on the superior aspect of the ~ransvcrsc proccsscs of the cervical vertebrae. Thus, at [his level they travel in a space bo~~ndcd anteriorly and posteriorly by the intertransverse muscles and superiorly and inferiorly by the bony transverse processes themselves. After reaching the tips of the transverse processes, the roots pass between the anterior and posterior tuberclcs of the transverse processes and then descend toward the first rib between the two walls of fascia covering the anterior and middle scalene muscles that form the interscalene space (Fig 4). As the roots pass clown through this space, they converge atop the first rib to form the three trunks of the plexus (Fig 5), which together with the subclavian artery invaginate the scalene fascia to form a subclavian perivascular space, bvhich in turn becomes the axillary perivascular space as the ncurovascular bundle continues into the axilla. The existence of this perineum1 and perivascular space renders brachial plexus anesthesia very simple. Just as with the epidural space, this space can also be entered at any level, and the volun~e of local anesthetic injecicd at that level will determine the extent of the resultant anesthesia. Furthermore, as with epidural anesthesia. the technique to be used in any given case should be determined by the surgical site and the dcsircd level of anesthesia, not by the training or bias of the anesthesiologisl.

the ccrvicnl

vertebrae to the distal anilia. The middle scalcnc nl~rscle third, fourth, fifth,

arises

from the ~OSICI ior tubcrclcs of the trans~crse proccss~s of the

The Subclavian Perivascular Technique


In 1964 \Iinnie described his subclavian perivasculaitwhnique that allowed for the first time accurate percutaneous location of rhe plcx~~s by using landmarks that mere constant and eas) to find. Further-more, this was the fil-st supraclavicular Icchnique that allom~~.l a single injection to pro\+clc anesthesia of the entire plexus. The block is carried out at the level Lvhere the three trunks cross the first rib; a point Lvliere 111~ plcx~~s is reduced to its fe\vest component parts, so at this level a smaller volume of local anesthetic is required to fill rhc space and block all of the components contained therein than with any other technique, \\,hich is the reason for the popularity of this technique and for its high success rate.

and

sixth cervical

vcrtchrac. s;calcnc

and cfcsccnds

t o w a r d s a n d inserts o n rhc first rib (Fig 1). 1~1st hehind

the
al-iscs of the

subclavian nrrcr);
froin the <itltct

\vhcrcas rhc anterior

iiiusclc

ior tuhcrclcs

o f rhe Irani\vsc

1~i-occsscs

F r o m t h e Department o f A n e s t h e s i o l o g y a n d P a i n M a n a g e m e n t , C o o k County Hospital. Chicago, IL; and the Department of Anesthesiology,


Rush-Presbyterian St. Lukes Medical Center, Chicago, IL. A d d r e s s r e p r i n t r e q u e s t s t o Aion P . W i n n i e , M D , D e p a r t m e n t o f A n e s t h e siology and Pain Management, Cook County Hospital, 1835 West Harris o n S t , S u i t e 7 3 0 8 . C h i c a g o , lL60612. Copyright c 1 9 9 7 b y W . B . S a u n d e r s C o m p a n y 1084-208X/97/01 04-0002$5.00/O

Fig 1. The posterior wall of the interscalene space is provided by the middle scalene muscle. 7iichniquc

Fig 3. The prevertebral fascia, as it extends laterally, splits to invest the scalene muscles.

The patient is placed in the dorsal recumbent position with the head turned somewhat to the opposite side. The rotation should not be so great that it stretches the II~L~SC~CS of the neck because this makes palpation difficult, The ipsilateral shoulder is lowered by asking the patient to reach for his knee. This maneuver, by lowering the clavicle, improves the access to the

trunks. After the procedure has been explained to the patient, the lateral border of the clavicular (lateral) head of the sternocleiclomastoid muscle is located by asking the patient to lift his head slightly off the table. The index finger of the palpating hand is placed behind this muscle at the level of C6 (the level of the cricoid cartilage). When the patient replaces

Fig 2. The anterior wall is provided by the anterior scalene muscle. The rootsof the plexus leave the cervical trans: verse processes, descend within the interscalene space toward the first rib, above which they combine to form the three trunks of the plexus.

Fig 4. Thus, the posterior fa:;cia of the anterior seal ene muscle and the anterior fascia of the middle scalene muscle form a fa scial sheatl n around the plexus.

his head on the table, the muscle relaxes and the palpating finger moves medially behind it, and comes to lie on the belly of the anterior scalene muscle. The palpating finger is now rolled laterally across the belly of the anterior scalene muscle until it encounters the interscalene groove, the space between the anterior and middle scalene muscles. Once the groove is identified, the index finger is moved as far down toward the clavicle as possible. In some patients movement of Ihe palpating finger inferiorly is obstructed by the omohyoid muscle that crosses the interscalene groove at this level; and if this is the case, the groove can usually be picked up 1 cm inferiorly. If not, the block is carried out at this level. If the groove can be followed all the way to the clavicle, the pulsation of the

subclavian artery can usually be felt where it emerges from between the scalenes, although palpation of the artery is not essential. At this point a 22 gauge 1 l/2 inch, short bevel, immobile needle* is inserted just superior to the palpating finger; and once it is through the skin, it is advanced directly caudad deep to the finger. A paresthesia below the shoulder or an appropriate twitch in response to a nerve stimulator confirms that the needle is properly placed within the subclavian perivascular space. As may be seen in Figure 6, the needle is inserted slightly closer to the middle scalene than to the anterior scalene muscle; and as it is advanced, because the three trunks are stacked vertically one on top of the other, there are three opportunities for a paresthesia: if the needle does not

Fig 5. Once formed, the three trunks of the plexus cross the first rib in the subciavian perivascular space, a lateral extension of the interscalene space. Note that the three trunks are stacked on top of one another at this level, and they lie closer to the middle than to the anterior scalene muscle.

single injections3 than after the traditional supraclavicular technique- and although most healthy patients are asymptomatic following any technique above the clavicle, phrenic block should probably be avoided in patients with respiratory insufficiency. Even in healthy patients, bilateral blocks are clearly unwise. Horners syndrome is a side effect rather than a complication of supraclavicular blocks. However, any technique that results in a sympathetic block should be avoided in severe asthmatics because of the risk of bronchospasm.

The Interscalene Technique


The roots enter the interscalene space and descend towards the first rib in the fascia-invested space between the anterior and middle scalene muscles after they leave the transverse processes, (Fig 4). With the interscalene technique7 it is within this space at the level of the roots that this block is carried out.

Fig 6. The Subclavian Perivascular Technique: the needle is inserted into the interscalene space directly caudad and therefore in one of the long axes of the space. Note that the needle is inserted slightly closer to the middle than to the anterior scalene muscle and that the advancing needle has three chances to contact one of the three trunks.

make contact \vith the superior trunk, it will hit the middle trunk. If it misses the middle trunk, it will hit the inferior trunk. And if the needle misses all three of the trunks, it will insert on the first rib along with the scalene muscles. Although contact with the rib is not deliberately sought, it is there to stop the needle if it should penetrate that far. \Vhen the desired response is obtained, the appropriate volume of local anesthetic is injected in 3 to 4 mL increments, with repcatcd aspiration for blood between increments. If a tourniquet is to be used, the intercostobrachial and medial brachial cutaneous nerves must be blocked by the subcutaneous injection of 3 to 4 mL of local anesthetic superficial to the axillary artery in the axilla. The easily identifiable landmarks, the direction of the needle insertion, the use of a short needle, and the need for only a single injection minimize the likelihood of pneumothorax with this technique, and make it simple and effective. Furthermore, as stated earlier, a smaller volume can be used to block the plexus with this technique than with any other, minimizing the possibility of systemic toxicity in poor risk, cachectic patitntc;.

Intravascular injcclion, phrenic ncrvt block, pncumoiliorax, hematoma, infection, and ncrvc injury arc potential complications of any supraclavicular technique, but most of these caii be avoided by a complctc understanding of the anatomy and careful technique. Although pneumothorax is the most dreaded complication of tcchniqucs above the clavicle, thcrc have been no reports of pneumothorax rcportcd follo\ving the suhclavian ;Tcrivascular technique. Phrenic block is less frcqutnt after

The patient is placed in the dorsal recumbent position, as is done for the subclavian perivascular technique, with the same slight rotation of the head to rhe opposite side. The block is performed al the level of C-6, which is determined by projecting a line laterally from the cricoid cartilage. After the procedure has been explained to the patient, he is asked to lift his head slightly to bring the clavicular (lateral) head of the sternocleidomastoid muscle into prominence. Both the middle and index fingers are placed behind the lateral margin of the sternocleidomastoid muscle, and the patient is asked to relax. As he does, the fingers move medially behind the sternocleidomastoid muscle and come to lie on the belly of the anterior scalenc muscle. The palpating fingers are rolled laterally across the anterior scalene muscle until the interscalene groove is perceived. The index and middle fingers are separated and pressed into the interscalene groove, depressing the skin to minimize the distance between skin and transverse process. At this point a 22 G 1 l/2 inch, short bevel, immobile needle is inserted perpendicular to the skin in all planes. Thus, as it is advanced, the needle direction is mostly mesiad, though it is also slightly dorsad and slightly caudad (Fig 7). The slight caudad direction is important from the point of view of safety because the slight caudad direction of the needle will cause it to strike bone if the plexus is not contacted. If the needle direction is too horizontal and the plexus is missed, the needle is free to advance into a vertebral vessel, the epidural space, the spinal canal, or even the spinal cord. The needle is ad\Fanced slowly until a paresthesia below the shoulder or an appropriate response to a nerve stimulator is elicited. As stated previously, if the plexus is missed, a transverse process will be contacted by the proptrly directed needle. In such a case, the U-shaped end of the transxerse process is walked millimeter b) millimeter until the desired response is evoked. Follo\ving aspirations, the local anesthetic is then slowly injected in 3 to 4 mL increments \vith repeated aspiration for blood bet\veen them.

Although pneumothorax has never been reported writh this technique, the incidence of phrenic block approaches 100%.H Thus, as suggested by Urme): rhis should not really be

was vertical, ie, perpendicular to the long axis of the body (and to the direction of needle insertion in the subclavian perivascular technique)

Fig 7. The lnterscalene Technique: the needle is inserted into the interscalene space mostly mesiad, but slightly dorsal and slightly caudad, again, in one of the long axes of the space, and closer to the middle than to the anterior scalene muscle.

considered a complication, but rather an expected sequel of interscalene block.s Nonetheless, even though phrenic block is asymptomatic in healthy patients, all of the supraclavicular techniques should be considered to be contraindicated in patients with compromised respiratory function, because acute respiratory failure has been described secondary to diaphragmatic paralysis in one such patient following an interscalene block. Furthermore, bilateral interscalene block should be staggered or avoided altogether, even in healthy patients. Horners syndrome is also very common after an interscalenc block, but this is more of a side effect than a complication and usually causes no adverse sequelat. There is one case report of severe bronchospasm, however, following an interscalene block in a severe asthmatic, presumably secondary to the sympathetic block. In addition, the interscalene technique carries with it the risk of an intravascular injection (into the vertebral artery or veins) as well as epidural and/or intrathecal injection, although all of these complications can be avoided or minimized by careful attention to proper needle direction. These two supraclavicular techniques, the subclavian perivascular and interscalenc techniques, together with the axillary perivascular technique, have allowed the present authors to provide safe and effective anesthesia for surgery anywhere on the entire upper extremity and shoulder girdle with few, if any, complications. Nonetheless, new techniques continue to be described, all of which involve penetration of and injection into the interscalene space, but from a different direction. Interestingly, each of these techniques was designed to decrease the incidence of pneutnothorax and not to increase the incidence of SLICCCSS.

The patient lies in the dorsal recumbent position with a pillow under his head and with the head turned to the side opposite that to be blocked. The patient is asked to raise his head to bring the stcrnoclcidomastoid muscle into prominence, the lateral edge of that muscle is marked, and the patient is told to relax. The anesthesiologist now places his index finger immediately lateral to the sternocleidomastoid muscle just above the clavicle. The finger is now on the belly of the anterior scalene muscle and is rolled laterally into the interscalene groove. An X is marked at a point immediately lateral to the edge of the anterior scalene muscle 1.5 to 2 cm above the clavicle (Fig 8). A 22 G, 4 cm needle attached to a filled syringe is now inserted vertically in an anteropostcrior direction (perpendicular to the table) and is advanced until a paresthesia is elicited, at which point the local anesthetic is injected after careful aspiration. If the plexus is missed by the advancing needle, the authors state that the first rib will be contacted. If, after several attempts, no paresthesias have been elicited, the authors simply inject the anesthetic solution along the lateral edge of the [anterior scalene] muscle a few millimeters above the first rib in a fanlike manner.] In their first 100 patients the authors reported an 89% SLICCCSS rate after the first injection, but they were able to increase the success rate to 97% by a second, supplemental injection. They reported no serious complications. The need to supplement an incomplete block increases the possibility 01 producing nerve damage, however, if and when the exploring needle encounters an unresponsive nerve.

Parascalene Approach of Dalens et al in Children


In 1987 the technique described by Vongvises and Panijaya nond was modified for use in children by Dalens et al,15 wht

______ ANTERIOR SCALENE MUSCL ____----------MIDDLE SCALENE MUSCLE

; BRACHIAL PLEXUS SUBCLAVlAN I ARTERY

Parascalene Technique of Vongvises and Panijayanond


In 1979 the first of scvcral parascaltne techniques \vas described by Vongviscs and Panija~anond.12,13 Their injection site was almost identical to that used in the subclavian pcrivascular technique, but the direction of needle insertion
Fig 8. The brachial plexus and scalene muscles and the poi (+) at which the needle is inserted for the parascaler technique of Vongvises. (Reprinted with permission.13)

dctcrniined in anatomic studies on pediatric cadavers that if 11~. injection site of Vongvises and Pani,jayanond were utilized
in
i11falltS injLlrCC1

in

a n d children, the cupola of the lung would he lllorc than 50% of the cases.

Technique
Tire child is placed in the dorsal recumbent position with a rolled towel under the shoulders. The head is turned toward the opposite side, and a line is drawn from the midpoint of the ciavicle t o Chassaignacs tubercle, that is located either by palpation or by the plane projected laterally from the cricoid cartilage to the lateral border of the sternoicleidomastoic muscle. This line is trisected, and the puncture site corrcsponds to the junction between the lower and middle thirds of the line. 4 22 G, 3 cm insulated needle connected to a nerve stimulator is advanced through this point directly posteriorly ipcrpendicular to the table) until twitch rcsponscs are elicited in the upper limb, at which point the anesthetic solution is injected. Using this modification, Dalcns et al reported a 98% success rate in their first 58 patients, with no major complications. Certainly, this is an impressive SLKCCSS rate, but the technique is a more complicated cookbook technique in terms of the landmarks, so more data by other investigators are necessary before its efficacy can lx assessed.

inserted at this point in a direction that is directly posterior (perpendicular to the table). The needle is adLanced until a paresthesia is elicited, at which point the injection is carried out. If a paresthesia at or distal to the ellmw is not obtained during the initial insertion, or if the first rib is not contacted, the needle is redircctcd cephalad in small steps until a paresthesia is obtained or until it is angled approximately 30 degrees cephalad. If the brachial plexus has still not been contacted, the needle is redirected caudad in small steps until a paresthesia is obtained or until an angle of 30 degrees caudad is reached. While the report states that the technique has heen used in more than 110 patients without a pneumothorax, the SUCCESS rate is not mentioned.

L F

Discussion
As stated earlier, the present authors have continued to use the original subclavian perivascular and intcrscalene techniques because they are easy to master, they are extremely effective, and they al-e also cxtrtmely safe. From an anatomical point of view these original techniques would appear to he more logical than the newer parascalene techniques, because with both the subclavian perivascular and intcrscalenc techniques the needle is advanced towards and enters the very narrow interscalene space in its long axes (Figs 5, 6, and 10A). As a result, the chance of the needle leaving the space as it is advanced or as the injection is taking place is minimized. With all three parascalcne techniques, the needle enters the space in its shortest axis (Fig lOB), so the slightest movement of even the properly placed needle may cause it to leave the space, resulting in the injection being made, in full or in part, outside of the fascial compartment. This possibility is enhanced by the fact that at the time of needle insertion the interscalene space is only a potential one, with the scalene fascia sandwiching the nerves between its two layers. It is only after the injection of the local anesthetic that the space becomes the true, fluid-filled compartment depicted in Figures I-6. From a clinical point of view these two techniques provide the anesthesiologist with the capability of providing regional anesthesia on any part of the hand, wrist, forearm, elbow, upper arm or shoulder. Again, with the subclavian perivascular

The Plumb-Bob Technique of Brown et al


The technique of Brown et alIi from 1993 is really describing yet another parascalenc technique, but the site of the itrjection is lower than that used in the two previously described techniques. The authors refer briefly to both the technique of Vbngvises and to that of Dalens, but they dismissed them because they require more complex measurements or equip ment than their Plumb-Bob technique.

Technique
The patient is placed supine with the head turned toward the opposite side (Fig 9). The point at which the lateral border of the sternocleidomastoid muscle joins the superior aspect of the clavicle is marked, and a 22 G, 5 to 6 cm, blunt needle is

Fig 9. Plumb-bob Technique of Brachial Plexus Block. (A) Needle entry site (black dot on dotted line) immediately lateral to insertion of clavicular head of sternocleidomastoid muscle onto clavicle. (B) Parasagittal plane showing vertical path (black dot on dotted line) from a position immediately superior to clavicle. (C) Right oblique view of the plumb bob technique, with a plane of glass outlining the plane in which the simulated needle-syringe assembly is moved through 30-degree arcs to seek parasthesia. (Reprinted with permission.ll)

unable to find a single report of pneumothorax following subclavian perivascular or interscalcne block in the literatut and in the two Departments chaired by one of the authors ov the past thirty years, several thousand subclavian perivascul blocks have been performed by residents and faculty withoul single pneumothorax (unless asymptomatic pneumothorac occurred). Interestingly, in an even larger series of interscalelk blocks at the same institutions, one experienced member a faculty had a pneumothorax in an elderly lady with adv bullous emphysema who turned out to have a single large on the apex of her lung which was encountered by the n Nonetheless, the safety record of these two supraclav techniques is impressive, and does not seem to indicate a net for the development of newer ones.

Conclusion
In summary, the subclavian perivascular and interscale] techniques of brachial plexus block provide the anesthesiol gist totally familiar with the anatomy two techniques that a technically easy to master and clinically effective for all types surgery on the upper extremity and shoulder. And mc importantly, both techniques have a very low incidence serious complications.

Fig 10. Schematic representation of the path of the needle, which is in the long axes of the interscalene space, with the lnterscalene and Subclavian Perivascular Technique (A) and the path of the needle, which crosses the shortest axis of the space, with the parascalene techniques. Clearly, there is much more leeway for movement of needle (A) than needle (B) during performance of the block.

References
Winnie AP, Collins VJ: The subclavian perivascular technique brachial plexus anesthesia. Anesthesiology 25:353-363,1964 Winnie AP: An immobile needle for nerve blocks. Anesthesiolo 31577-578, 1969 Farrar MD, Scheybani M, Nolte H: Upper extremity block effectiv ness and complications. Reg Anesth 6:133-134, 1981 Hartel F, Kepler W: Experience with Kulenkampffs technic of brach plexus anesthesia with special regard to complications during al after the block [in German]. Arch f Klin Chir 103: l-43, 1913 Shaw WM: Paralysis of the phrenic nerve during brachial plexl anesthesia. Anesthesiology 10:627-628, 1949 Knoblanche GE: The incidence and aetiology of phrenic ner blockade associated with supraclavicular brachial plexus block Anaesth Intensive Care 7:346-349, 1979 Winnie AP: lnterscalene brachial plexus block. Anesth Analg 49:45 466,197O Urmey WF, Talts KH, Sharrock NE: One hundred percent incidence hemidiaphragmatic paresis associated with interscalene brach plexus anesthesia as diagnosed by ultrasonography. Anesth Ana 72:498-503, 1991 Hood J, Knoblanche G: Respiratory failure following brachial plexl block. Anesth Intensive Care 7:346-349, 1979 Lim EK: interscalene brachial plexus block in the asthmatic patier Anaesthesia 34:370, 1979 (letter) Brown DL, Cahill DR, Bridenbaugh LD: Supraclavicular nerve bloc Anatomic analysis of a method to prevent pneumothorax. Anes A n a l g 76:530-534, 1 9 9 3 Vongvises P, Panijayanond T: A parascalene technique of brachi plexus anesthesia. Anesth Analg 58:267-273, 1979 Vongvises P, Beokhaimook, M: Computed tomographic study parascalene block. Anesth Analg 84: 379-382, 1997 Selander D, Edshage S, Wolff T: Paresthesiae or no paresthesiat Nerve lesions after axillary blocks. Acta Anaesthiol Stand 23:27-3 1979 Dalens B, Vanneuville G, Tanguy A: A new parascalene approach the brachial plexus in children: Comparison with the supraclavicul approach. Anesth Analg 66:1264-l 271, 1987 Moore DC: Complications of regional anesthesia. Clin Anesth 721 251,1969 Brand L, Papper EM: A comparison of supraclavicular and axilla techniques for brachial plexus blocks. Anesthesiology 22:226-22 1961

technique the local anesthetic is injected at the level of the trunks, so the distribution of the anesthesia (or the lack of anesthesia if the block is incomplete) will be in the distribution of the trunks. Clearly, therefore a successful subclavian perivascular block will provide anesthesia of the entire upper extremity. With the interscalene technique, on the other hand, the injection of local anesthetic is made at the level of the upper roots of the plexus, so the resultant anesthesia (or lack of anesthesia if the block is incomplete) will be in the distribution of the roots. This is in contrast to the distribution of anesthesia with an axillary perivascular block, which is in the distribution of the peripheral nerves that are bathed by the injected local anesthetic. Furthermore, since the injection is made at the level of the upper roots of the brachial plexus and lower roots of the cervical plexus, the interscalene technique is ideal for surgery on the upper arm or shoulder. Either technique can be used for surgery on the arm or shoulder, however, if necessary, by simply increasing the volume of the local anesthetic injected. When it is necessary to use an interscalene block for hand and forearm surgery, it has been our experience that firm digital pressure above the needle during the injection will favor the caudad spread of the solution and facilitate blockade of the elusive eighth cervical and first thoracic roots. From the point of view of safety, all of the newer techniques were introduced not to improve the success rate but to lessen the incidence of pneumothorax, the incidence of which, all three articles state, ranges from 0.5% to 6%. These often quoted statistics, however, represent the incidence of pneumothorax reported for the supraclavicular techniques that were used prior to the introduction of the subclavian perivascular and interscalene techniques. Ih, Yet, the present authors have been

5. 6.

7. 8.

9. 10. 11.

12. 13. 14.

15.

16. 17.

Regional Anesthesia for Hip Surgery


Learning Objectives: 1. Describe the indications and contraindications for the various regional anesthesia and analgesia techniques for surgery on the hip 2. Describe the relevant neuro anatomy of the lumbar plexus, lumbosacral plexus and innervation to the hip 3. Describe the techniques, local anesthetics and adjuvants used for performing regional anesthesia and analgesia for surgery on the hip 4. Compare and contrast the benefits of regional anesthesia and analgesia versus general anesthesia for both the intra- and postoperative period.

2000 American Society of Anesthesiologists, Inc. Volume 93(1) July 2000 pp 115-121

Lumbar Plexus Block Reduces Pain and Blood Loss Associated with Total Hip Arthroplasty [Clinical Investigations] Stevens, Robert D. M.D.*; Van Gessel, Elisabeth M.D.; Flory, Nicolas M.D.; Fournier, Roxane M.D.; Gamulin, Zdravko M.D. *Research Fellow. Staff Anesthesiologist. Chief Resident. Received from the Department of Anesthesiology, Pharmacology and Surgical Intensive Care, Hpitaux Universitaires de Genve, Switzerland. Address reprint requests to Dr. Stevens: Department of Anesthesiology, Pharmacology and Surgical Intensive Care, Hpitaux Universitaires de Genve, 1211 Geneva 14, Switzerland. Address electronic mail to: Robert.Stevens@hcuge.ch Abstract Background: The usefulness of peripheral nerve blockade in the anesthetic management of hip surgery has not been clearly established. Because sensory afferents from the hip include several branches of the lumbar plexus, the authors hypothesized that a lumbar plexus block could reduce pain from a major hip procedure. Methods: In a double-blind prospective trial, 60 patients undergoing total hip arthroplasty were randomized to receive general anesthesia with (plexus group, n = 30) or without (control group, n = 30) a posterior lumbar plexus block. The block was performed after induction using a nerve stimulator, and 0.4 ml/kg bupivacaine, 0.5%, with epinephrine was injected. General anesthesia was standardized, and supplemental fentanyl was administered per hemodynamic guidelines. Postoperative pain and patient-controlled intravenous morphine use were serially assessed for 48 h. Results: The proportion of patients receiving supplemental fentanyl intraoperatively was more than 3 times greater in the control group (20 of 30 vs. 6 of 29, P = 0.001). In the postanesthesia care unit, a greater than fourfold reduction in pain scores was observed in the plexus group (visual analogue scale [VAS] pain score at arrival 1.3 2 vs. 5.6 3, P < 0.001), and rescue morphine boluses (administered if VAS > 3) were administered 10 times less frequently (in 2 of 28 vs. in 22 of 29 patients, P < 0.0001). Pain scores and morphine consumption remained significantly lower in the plexus group until 6 h after randomization (VAS at 6 h, 1.4 1.3 vs. 2.4 1.4, P = 0.007; cumulative morphine at 6 h, 5.6 4.7 vs. 12.6 7.5 mg, P < 0.0001). Operative and postoperative (48 h) blood loss was modestly decreased in the treated group. Epidural-like distribution of anesthesia occurred in 3 of 28 plexus group patients, but no other side-effects were noted. Conclusions: Posterior lumbar plexus block provides effective analgesia for total hip arthroplasty, reducing intra- and postoperative opioid requirements. Moreover, blood loss during and after the procedure is diminished. Epidural anesthetic distribution should be anticipated in a minority of cases.

TOTAL hip arthroplasty, one of the most frequently performed surgical procedures, generates significant postoperative pain that may be treated using regional anesthesia. When compared with other regimens, regional anesthesia provides superior pain relief and may favorably influence outcomes such as blood loss and thromboembolic events. 1 In the case

of major knee surgery, recent evidence suggests that early rehabilitation may be improved by use of regional anesthetic techniques. 2,3 Peripheral nerve blocks of the lower extremities, which offer many of the advantages of other types of regional anesthesia, are also reported to circumvent some of the drawbacks associated with these other types. 24 Regional blocks have thus gained acceptance for perioperative management of procedures on the knee and below, both as a complement to general anesthesia and as an alternative to centroneuraxial analgesia. For surgery of the hip, however, the role of peripheral nerve blockade needs to be defined. Sensory innervation of the hip involves branches of the lumbar plexus (LP) and the sacral plexus. 5,6 Local anesthetic may be directed to the LP by an anterior approach called paravascular, or 3-in-1 block, 7 or by a posterior approach, of which several methods have been described. 810 The concept of the 3-in-1 technique as a plexus block is controversial because it does not consistently produce anesthesia of the obturator or lateral femoral cutaneous nerves. 11,12 Moreover, in a recent study of total hip arthroplasty, the benefit of 3-in-1 block in terms of reduced pain scores or opioid requirement was not clearly shown. 13 Studies show that posterior LP techniques are reliable in their ability to block major LP branches. 11,14 Based on the observation that the LP innervates a significant portion of the hip region, we tested the hypothesis that an LP block would provide effective analgesia for total hip arthroplasty. Ancillary end points, such as blood loss, block side-effects and complications, postoperative nausea and vomiting, and patient satisfaction, were also assessed. Materials and Methods After obtaining institutional review board approval and written informed patient consent, we conducted a randomized, controlled, double-blind trial of 60 consecutive patients undergoing elective total hip arthroplasty during general anesthesia. Exclusion criteria were contraindications to regional anesthesia, use of opioids during the preoperative period, and dementia preventing proper comprehension of the study. Patients were randomly allocated to receive general anesthesia combined with an LP block (plexus group, n = 30) or general anesthesia alone (control group, n = 30). All patients were scheduled for surgery at 8:00 am. After premedication with 7.5 mg midazolam orally, general anesthesia was induced with 46 mg/kg sodium thiopental and 2 g/kg fentanyl and maintained with 0.31% isoflurane and 6070% nitrous oxide. Tracheal intubation was facilitated with use of 0.2 mg/kg mivacurium and lungs were mechanically ventilated (end-tidal carbon dioxide, 3038 mmHg [4 to 5 kPa]). Patients were placed in the lateral decubitus position, operative extremity superior. After recovery from use of mivacurium (as assessed by train-of-four stimulation of the ulnar nerve or by the appearance of respiratory efforts interfering with mechanical ventilation), patients assigned to the plexus group were administered a single-injection posterior LP block using the approach described by Winnie et al.8 The LP was localized by inducing contractions of the quadriceps femoris with use of a nerve stimulator (DualStim; Life-Tech, Houston, TX), delivering 0.20.5 mA impulses of 50 s at 2 Hz linked to a 23-gauge, 100-mm sterile needle (Pole Needle; Top Corporation, Tokyo, Japan). After aspiration to ensure absence of blood or cerebrospinal fluid, 0.4-ml/kg bupivacaine, 0.5%, with epinephrine 1/200,000 was injected. In the control group, lumbar skin was perforated with use of a needle, but no placebo was administered. In conformity with the blinded study design, the anesthesiologist responsible for the patient was temporarily absent during treatment allocation, leaving the patient in the care of an attending anesthesiologist. The attending anesthesiologist administered the block when assigned, then transferred the patient to the initial staff. The surgical procedure was standardized and was performed with patients positioned as previously mentioned. Intraoperatively, increments of vecuronium (1 mg) or mivacurium (2 mg) were administered only if warranted to facilitate the surgical procedure or mechanical ventilation. Hemodynamic goals were to maintain mean arterial pressure (MAP) and heart rate within 70130% of preinduction or baseline levels. MAP increases to more than 130% of baseline were treated by raising end-tidal isoflurane to 1 vol% and administering boluses of fentanyl, 1 g/kg. A decrease in MAP to less than 70% of baseline was treated with ephedrine, 10 mg intravenously, followed by reduction of end-tidal isoflurane to 0.3%. In the postanesthesia care unit (PACU), a patient-controlled analgesia device was given to all patients and set to deliver intravenous morphine in 1-mg boluses, with a lockout at 6 min and a 4-h maximum of 40 mg. Patients in the PACU reporting pain greater than 3 on the VAS (0 = no pain, 10 = most severe pain) despite patientcontrolled analgesia were administered boluses of intravenous rescue morphine as needed. In addition, all patients

received propacetamol, 2 g intravenously every 8 h administered shortly before the end of surgery and continued for 24 h, and ibuprofen, 400 mg orally every 8 h started on the morning after surgery. Before selection for the study, a majority of study participants had been enrolled in an autologous blood transfusion program (plexus group, 24 of 30 patients; control group, 19 of 30 patients, P = 0.25). Each patient predonated 2 to 3 units of blood during the weeks preceding surgery. In accordance with common practice at the Hpitaux Universitaires de Genve, indications for perioperative autologous blood transfusions were not subject to specific guidelines. Pain scores at rest and morphine consumption were assessed every 30 min in the PACU and then 6, 12, 24, and 48 h after randomization by observers who were blind to treatment allocation. Other recorded variables included intraoperative MAP, heart rate, and end-tidal isoflurane concentration; intraoperative opioid and muscle relaxant use; loss of blood, intraoperative (volume of blood measured in suction canisters and estimated from operative dressings [40 ml blood/dressing]) and postoperative (volume of blood recovered in suction drains before removal at 48 h); bilateral distribution of anesthesia, suggesting epidural spread (tested in the PACU with use of a cold stimulus applied to lower thoracic and lumbar dermatomes contralateral to the operated hip); evidence for block-related neurologic damage at 48 h; postoperative nausea or vomiting; and patient satisfaction with anesthetic and pain management (rated using VAS, 10 = very satisfied, 0 = very unsatisfied). All observations were made by blinded assessors. Statistical Analysis Based on previous data, 13,15 it was computed that a sample of 30 patients/group would detect a difference between groups in mean morphine consumption of more than 20 mg and a difference in VAS pain scores of more than 1.7 cm, with a power of 90% and a two-tailed significance level of 5% ([beta] = 0.1, [alpha] = 0.05). In an analysis made after obtaining results and using other data, 16 it was calculated that a sample of this size had 90% power to detect a difference of 130 ml in mean operative blood loss. Where relevant, data are presented as the mean SD. Comparisons between groups of continuous variables (MAP, end-tidal isoflurane concentration, morphine use, pain scores, satisfaction scores, blood loss) were assessed using the unpaired Student t test, whereas comparisons of discontinuous data (proportions of patients) were evaluated using the chi-square test. P < 0.05 was considered to be significant. Some data were analyzed using StatMate and Prism software (GraphPad, San Diego, CA). Results Preoperative characteristics were similar in the two groups (table 1). Two patients, one in each group, were excluded from analysis of postoperative data because of delirium, impeding accurate evaluation. One patient in the plexus group received a dose of local anesthetic incompatible with the study requirement and was excluded from intra- and postoperative analysis. Bilateral distribution of anesthesia, suggesting an epidural extension of local anesthetic, occurred in 3 of 28 (10.7%) patients undergoing the block.

Table 1. Preoperative CharacteristicsValues are SD.ASA = American Society of Anesthesiologists physical status; VAS = visual analogue scale. Data related to anesthetic and perioperative management are summarized in table 2 and figures 1 and 2. The number of patients requiring supplemental fentanyl and the mean end-tidal isoflurane concentration were significantly increased in the control group, whereas intraoperative MAP was lower in the plexus group during most of the procedure. These differences in MAP remained significant even when patients with epidural-type distribution were excluded from analysis (P < 0.01 for the difference between groups in MAP from the twentieth to the sixtieth min of surgery, and P < 0.02 for the difference between groups at the eightieth and ninetieth min). A reduction in blood loss was observed in the plexus group, during the surgical procedure (22% reduction in loss) and 48 h after (45% reduction). When patients who had epidural distribution were excluded from analysis, the intraoperative blood-sparing effect was not statistically significant (plexus group mean blood loss, 434 196 ml vs. controls, 538 254 ml;P = 0.09), whereas the postoperative loss remained significantly lower (144 97 ml vs. 310 204 ml, P = 0.0006). Eight patients, four in each group, received a unit of autologous blood intraoperatively, and all remaining autologous units were transfused over the ensuing 48 h. No heterologous blood was administered.

Table 2. Anesthetic and Perioperative VariablesValues are mean SD.Supplemental = agents administered after the induction period (see text).

Fig. 1. Isoflurane administration is reduced in patients undergoing lumbar plexus block (plexus). Values are mean SD. Zero on the time scale marks the beginning of surgery. *P <= 0.001 for differences between groups.

Fig. 2. Intraoperative mean arterial pressure is decreased in patients who underwent a lumbar plexus block (plexus). Values are mean SD. On time scale, B refers to baseline mean arterial pressure recorded before induction of anesthesia, and 0 marks the beginning of surgery. *P <= 0.005 for differences between groups; #P = 0.01; and +P = 0.03. Pain scores were significantly reduced in the plexus group until 6 h after randomization (VAS at 6 h, 1.4 1.3 vs. 2.4 1.4, P = 0.007). Differences in pain scores were particularly evident in the immediate postoperative period (fig. 3). Therefore, at arrival to the PACU, 18 of 28 patients in the plexus group reported no pain (VAS = 0) versus 3 of 29 in the untreated control group (P < 0.0001). Conversely, at this same time point, 17 of 29 control patients had severe pain (VAS > 5), versus 2 of 28 of treated patients (P < 0.0001). Morphine administration, including patient-controlled analgesia and rescue doses, was diminished to a significant degree in the treated group until 6 h after randomization (cumulative morphine at 6 h, 5.6 4.7 vs. 12.6 7.5 mg, P < 0.0001), with a trend toward reduced consumption at 12 h (fig. 4). Only 2 of 28 patients required rescue morphine in the plexus group, versus 22 of 29 among controls (P < 0.0001).

Fig. 3. Postoperative pain scores. (A) Pain scores in the postanesthesia care unit (PACU). (B) Pain scores in the wards. Upper and lower sides of boxes represent twenty-fifth and seventy-fifth percentiles, horizontal lines inside boxes are medians, shaded squares are means, and error bars represent ranges. *P < 0.001 and #P = 0.007 for differences between group means.

Fig. 4. Patients who underwent lumbar plexus block (plexus) use significantly less morphine in the early postoperative period. Cumulative morphine includes patient-controlled analgesia and rescue doses. Upper and lower sides of boxes represent twenty-fifth and seventy-fifth percentiles, horizontal lines inside boxes are medians, shaded squares are means, and error bars represent ranges. *P < 0.0001 and #P = 0.05 for differences between group means. Postoperative nausea and vomiting occurred in a minority of patients, and the incidence was similar in both groups. No other side-effects or complications were noted. Patient satisfaction with anesthetic and pain management was not significantly different between the two groups. Discussion Our results show that posterior LP block successfully reduces pain associated with total hip arthroplasty, decreasing intra- and early postoperative opioid requirements. Another interesting finding was a block-associated reduction in hemorrhage. Peripheral nerve or plexus blocks, with or without perineural catheter placement, constitute an attractive option for anesthetic management of lower limb surgery, either substituting for or complementing other anesthetic techniques. Blocks are appreciated for the superlative and long-lasting analgesia they provide; moreover, when compared with other types of anesthesia, nerve blocks may enhance intraoperative hemodynamic stability, 4,17 decrease risk of urinary retention, 3,18 reduce risk of puncturing the dura mater, and diminish surgical blood loss. 16 The LP is formed by the ventral rami of the first four lumbar roots, with a contribution from the twelfth thoracic root. These elements converge in the posterior part of the psoas muscle and produce seven to eight branches, all of which contribute to a varying degree to the sensory innervation of the hip region. 5 The hip joint is innervated by branches of both the LP and the sacral plexus, including the femoral (L2L4) and obturator nerves (L2L4), which contain afferents from the anteromedial aspects of the joint, and the sciatic nerve (L4S3), the nerve to the quadratus femoris (L4S1) and

the superior gluteal nerve (L4S1), which cover the posterior aspect of the joint. 5,6 Although the relative contributions of lumbar and sacral components to sensory innervation are not known, our results suggest that local anesthetic blockade of the LP is sufficient to provide effective analgesia early after total hip arthroplasty. Although this finding is in itself of interest, we predict that its clinical relevance will be further shown by the application of continuous catheterbased techniques, as indicated in a recent preliminary report. 19 Lumbar plexus blocks were developed by Winnie et al., who described both an anterior 7 and a posterior 8 technique. The anterior, or 3-in-1, approach was based on the hypothesis that a large volume of local anesthetic placed in the femoral nerve sheath would spread proximally to the LP and block other branches, including the obturator and lateral femoral cutaneous nerves. Although the original account supported this postulate, subsequent reports failed to show reliable blockade of obturator or lateral femoral cutaneous nerves with use of this technique, and the concept of the 3-in1 technique as a plexus block is contested. 11,12 With the posterior LP approach, local anesthetic is deposited within or adjacent to the dorsal portion of the psoas muscle. Several methods of performing the posterior LP block have been described, 810 all of which consistently produce anesthesia of LP branches important to lower limb surgery. In addition, some authors using posterior techniques have noted simultaneous anesthesia of lumbar and sacral plexi, 8,20 whereas others have not. 11,14 Few studies have addressed the clinical interest of peripheral nerve or plexus blocks for surgery of the hip. Using a posterior LP technique, Chayen et al.9 reported successful anesthesia in 52 of 57 hip procedures, but recommended combining lumbar and sciatic blocks in this indication. In a trial that compared general anesthesia alone or combined, either with subarachnoid block or with posterior LP block for femoral neck fracture surgery, White and Chappell 17 noted greater cardiovascular stability in the group receiving the block; but provided no data regarding other outcomes, such as postoperative analgesia. In a study of hip replacement, Odoom et al.21 compared bupivacaine with or without epinephrine in a posterior LP block and noted lower peak bupivacaine concentrations and a prolongation of mean analgesia in the epinephrine group. Dalens et al., 20 reporting a pediatric population of which 9 of 50 children underwent hip arthrotomies, contrasted posterior LP block performed according to the techniques of Winnie et al.8 or Chayen et al., 9 noting lumbosacral anesthetic distribution in the former group and bilateral or epidural-like distribution in the latter, without providing information about pain or analgesia. Finally, Fournier et al.13 compared general anesthesia combined with 3-in-1 block versus general anesthesia alone for total hip arthroplasty and observed a longer opioid-free interval after surgery in the block group, although intraoperative use of opioids and isoflurane was not influenced by the block. Contrasting these results and ours suggests that the posterior LP block, by providing more extensive anesthetic coverage of the region, may be better suited to surgical procedures of the hip than is the 3-in-1 approach. The decreased blood loss we report, intraoperatively and postoperatively, confirm the findings of Twyman et al., 16 although only intraoperative loss was significantly reduced in that study. The clinical importance of this hemorrhagesparing effect in terms of transfusion requirements was not directly addressed in the current study because most patients predonated blood for which transfusion policy was not guideline-directed, prohibiting a proper statistical comparison between groups. Diminished hemorrhage has been documented with various regional anesthetic techniques, including spinal and epidural anesthesia, 1 and is thought to result from attenuated sympathetic tone in medium and small vessels, with concomitantly reduced arterial and venous pressure. Two distinct mechanisms may influence blood loss in patients undergoing peripheral nerve block: a direct effect on vasoconstrictive sympathetic fibers contained in peripheral nerves and an indirect effect mediated by antinociception and reduced systemic blood pressure. In this regard, we noted lower intraoperative MAP in the plexus group (fig. 2). Because this difference was observed in the presence of equivalent baseline MAP measurements and similar prevalence of treated arterial hypertension in the two groups, with lesser administration of isoflurane and fentanyl in the plexus group, we suggest that it is attributable to attenuated nociception and autonomic arousal in the block-treated patients. Finally, lower intraoperative blood loss in the plexus group may be related to epidural-like blockade appearing in some of these patients. In 3 of 28 patients in the plexus group, bilateral distribution of anesthesia developed, evoking spread of local anesthetic to the epidural space. Hypotension or extension of block to the upper thoracic dermatomes did not develop in any of these patients. Epidural extension of anesthesia is a known side effect of posterior LP block, occurring in 310% of adult patients without adverse consequences. 9,11,14 Notwithstanding, we believe that this result should be systematically checked after posterior LP block has been performed. Reported complications of posterior LP block, such

as total spinal anesthesia, 22 renal subcapsular hematoma, 23 and psoas hematoma with lumbar plexopathy, 24 were not observed in this trial. The strengths of this study include the fact that it was designed to detect targeted differences between groups in the two major end points (pain, morphine use). A post hoc analysis showed that the trial also was sufficiently powered to detect observed differences in an ancillary end point: blood loss. Among the limitations, we noted that, although care was taken to ensure the double-blind design, observations such as asymmetric thigh mobility, epidural blockade, or decreased opioid requirements might have introduced bias among patients and data collectors. In summary, LP block performed using a posterior approach is an effective complement to anesthetic and analgesic management of total hip arthroplasty, reducing opioid administration and blood loss during the surgical procedure and in the early postoperative period. Future studies should address the possibility of extending the duration of this benefit in the postoperative period, namely via continuous catheter-based techniques. Studies are also needed to confirm reduced bleeding in the setting of peripheral nerve blockade and to explore the clinical importance of this effect. References 1. Covert CR, Fox GS: Anaesthesia for hip surgery in the elderly. Can J Anaesth 1989; 36:3119 [Medline Link] [Context Link] 2. Singelyn FJ, Deyaert M, Joris D, Pendeville E, Gouverneur JM: Effects of intravenous patient-controlled analgesia with morphine, continuous epidural analgesia, and continuous three-in-one block on postoperative pain and knee rehabilitation after unilateral total knee arthroplasty. Anesth Analg 1998; 87:8892 [Fulltext Link] [Medline Link] [Context Link] 3. Capdevila X, Barthelet Y, Biboulet P, Ryckwaert Y, Rubenovitch J, dAthis F: Effects of perioperative analgesic technique on the surgical outcome and duration of rehabilitation after major knee surgery. A nesthesiology 1999; 91:8 15 [Fulltext Link] [Medline Link] [Context Link] 4. Fanelli G, Casati A, Aldegheri G: Cardiovascular effects of two different regional anaesthetic techniques for unilateral leg surgery. Acta Anaesthesiol Scand 1998; 42:804 [Medline Link] [Context Link] 5. Grays Anatomy, 38th edition. Nervous system. Edited by Williams PL, Bannister LH, Berry MM, Collins P, Dyson M, Dussek JE, Ferguson MWJ. New York, Churchill Livingstone, 1995, pp 127792 [Context Link] 6. Birnbaum K, Prescher A, Hepler S, Heller KD: The sensory innervation of the hip jointAn anatomical study. Surg Radiol Anat 1997; 19:3715 [Medline Link] [Context Link] 7. Winnie AP, Ramamurthy S, Durrani Z: The inguinal paravascular technic of lumbar plexus anesthesia: the 3-in-1 block. Anesth Analg 1973; 52:98996 [Medline Link] [Context Link] 8. Winnie AP, Ramamurthy S, Durrani Z, Radonjic R: Plexus blocks for lower extremity surgery: New answers to old problems. Anesth Review 1974; 1:116 [Context Link] 9. Chayen D, Nathan H, Chayen M: The psoas compartment block. A nesthesiology 1976; 45:959 [Medline Link] [Context Link] 10. Hanna MH, Peat SJ, DCosta F: Lumbar plexus block: An anatomical study. Anaesthesia 1993; 48:6758 [Medline Link] [Context Link] 11. Parkinson SK, Mueller JB, Little WL, Bailey SL: Extent of blockade with various approaches to the lumbar plexus. Anesth Analg 1989; 68:2438 [Medline Link] [Context Link]

12. Seeberger MD, Urwyler A: Paravascular lumbar plexus block: Block extension after femoral nerve stimulation and injection of 20 vs. 40 ml mepivacaine 10 mg/ml. Acta Anaesthesiol Scand 1995; 39:76973 [Medline Link] [Context Link] 13. Fournier R, Van Gessel E, Gaggero G, Boccovi S, Forster A, Gamulin Z: Postoperative analgesia with 3-in-1 femoral nerve block after prosthetic hip surgery. Can J Anaesth 1998; 45:348 [Medline Link] [Context Link] 14. Farny J, Girard M, Drolet P: Posterior approach to the lumbar plexus combined with a sciatic nerve block using lidocaine. Can J Anaesth 1994; 41:48691 [Medline Link] [Context Link] 15. Serpell MG, Millar FA, Thomson MF: Comparison of lumbar plexus block versus conventional opioid analgesia after total knee replacement. Anaesthesia 1991; 46:2757 [Medline Link] [Context Link] 16. Twyman R, Kirwan T, Fennelly MJ: Blood loss reduced during hip arthroplasty by lumbar plexus block. J Bone Joint Surg 1990; 72:7701 [Context Link] 17. White IWC, Chappell WA: Anaesthesia for correction of fractured femoral neck. A comparison of three techniques. Anaesthesia 1980; 35:110710 [Medline Link] [Context Link] 18. Matheny JM, Hanks GA, Rung GW, Blanda JB, Kalenak A: A comparison of patient-controlled analgesia and continuous lumbar plexus block after anterior cruciate ligament reconstruction. Arthroscopy 1993; 9:8790 [Medline Link] [Context Link] 19. Pandin P, Huybrechts I, Mathieu N, Vandensteene A, Engelman E, dHollander A: Psoas compartement catheter for hip replacement: Preliminary results about an alternative for analgesia (abstract). A nesthesiology 1998; 89:A843 [Context Link] 20. Dalens B, Tanguy A, Vanneuville G: Lumbar plexus block in children: A comparison of two procedures in 50 patients. Anesth Analg 1988; 67:7508 [Medline Link] [Context Link] 21. Odoom JA, Zuurmond WW, Sih IL, Bovill J, Osterlof G, Oosting HV: Plasma bupivacaine concentrations following psoas compartment block. Anaesthesia 1986; 41:1558 [Medline Link] [Context Link] 22. Gentili M, Aveline C, Bonnet F: Total spinal anesthesia after posterior lumbar plexus. Block. Ann Fr Anesth Reanim 1998; 17:7402 [Medline Link] [Context Link] 23. Aida S, Takahashi H, Shimoji K: Renal subcapsular hematoma after lumbar plexus block. A nesthesiology 1996; 84:4525 [Fulltext Link] [Medline Link] [Context Link] 24. Klein SM, DErcole F, Greengrass RA, Warner DS: Enoxaparin associated with psoas hematoma and lumbar plexopathy after lumbar plexus block. A nesthesiology 1997; 87:15769 [Fulltext Link] [Medline Link] [Context Link]

1999 American Society of Anesthesiologists, Inc. Volume 91(4) October 1999 p 926

Randomized Trial of Hypotensive Epidural Anesthesia in Older Adults [Clinical Investigations] Williams-Russo, Pamela M.D., M.P.H.*; Sharrock, Nigel E. M.B.Ch.B.; Mattis, Steven Ph.D.; Liguori, Gregory A. M.D.; Mancuso, Carol M.D.[//]; Peterson, Margaret G. Ph.D.#; Hollenberg, James M.D.*; Ranawat, Chitranjan M.D.**; Salvati, Eduardo M.D.; Sculco, Thomas M.D. *Associate Professor, Department of Medicine, Cornell Medical College. Senior Scientist and Assistant Clinical Professor, Department of Anesthesia, Cornell Medical College. Professor, Departments of Psychology and Psychiatry, Hillside Hospital. Assistant Clinical Professor, Department of Anesthesia, Cornell Medical College. [//]Assistant Professor, Department of Medicine, Cornell Medical College. #Statistician, Department of Medicine, Cornell Medical College. **Clinical Professor, Department of Orthopedic Surgery, Lenox Hill Hospital. Clinical Professor, Department of Orthopedic Surgery, Cornell Medical College. Received from the Hospital for Special Surgery, Cornell University Medical College, Lenox Hill Hospital, Hillside Hospital, Glen Oaks, and the Albert Einstein College of Medicine, New York, New York. Submitted for publication December 14, 1998. Accepted for publication June 1, 1999. Supported by a grant 2RO1 AGO8562 from the National Institute of Aging, Bethesda, Maryland, and in part by the Cornell Multipurpose Arthritis and Musculoskeletal Disease Center, Cornell University Medical College, New York, New York. Address reprint requests to Dr. Williams-Russo: Hospital for Special Surgery, 535 E. 70th Street, New York, New York 10021. Address electronic mail to: pgwruss@mail.med.cornell.edu This article is accompanied by an Editorial View. Please see: Raja SN, Haythornthwaite JA: Anesthetic management of the elderly: Measuring function beyond the immediate perioperative horizon. ANESTHESIOLOGY 1999; 91:90911 Abstract Background: Data are sparse on the incidence of postoperative cognitive, cardiac, and renal complications after deliberate hypotensive anesthesia in elderly patients. Methods: This randomized, controlled clinical trial included 235 older adults with comorbid medical illnesses undergoing elective primary total hip replacement with epidural anesthesia. The patients were randomly assigned to one of two levels of intraoperative mean arterial blood pressure management: either to a markedly hypotensive mean arterial blood pressure range of 4555 mmHg or to a less hypotensive range of 5570 mmHg. Cognitive outcome was assessed by within-patient change on 10 neuropsychologic tests assessing memory, psychomotor, and language skills from before surgery to 1 week and 4 months after surgery. Prospective standardized surveillance was performed for cardiovascular and renal outcomes, delirium, thromboembolism, and blood loss and replacement. Results: The two groups were similar at baseline in terms of age (mean, 72 yr), sex (50% women), comorbid conditions, and cognitive function. After operation, no significant differences in the incidence of early or long-term cognitive dysfunction were observed between the two blood pressure management groups. There were no significant differences in the rates of other adverse consequences, including cardiac, renal, and thromboembolic complications. In addition, no differences occurred in the duration of surgery, intraoperative estimated blood loss, or transfusion rates. Conclusions: Elderly patients can safely receive controlled hypotensive epidural anesthesia with this protocol. There was no evidence of greater risks, or early benefits, with the use of the more markedly hypotensive range.

DELIBERATE hypotensive anesthesia offers significant advantages for surgical procedures that require dry surgical fields or are associated with substantial blood loss. These advantages must be balanced against the risks of ischemic injury in nonsurgical regions, such as the brain and heart. The use of hypotensive anesthesia generally has been limited to healthy young patients and to surgical procedures that cannot be performed under normotensive conditions. Total hip replacement (THR) surgery is a clearly beneficial procedure performed most often in older adults. 1 Hypotensive anesthesia for THR offers several benefits: a drier surgical field, better visualization of anatomy, and reduced intraoperative blood loss. 2,3 However, the use of deliberate hypotension in elderly patients raises concerns of hypoperfusion and ischemic injury to the brain, heart, or other vital organs. 47 Cerebral injury could take the form of obvious stroke or more subtle long-term loss of cognitive function. 811 Previous controlled prospective studies of hypotensive anesthesia lacked rigorous surveillance for postoperative cognitive outcomes, had limited generalizability, or had low power because of small patient populations. 2,1215 A recently developed technique for induced hypotension with epidural anesthesia offers a way to preserve or augment cardiac output and systemic flow despite significant systemic hypotension. 16,17 The purpose of this randomized controlled trial was to assess the risks and benefits of hypotensive epidural anesthesia with this technique for elderly patients with comorbid cardiovascular disease undergoing THR. Patients were randomized to have their intraoperative mean arterial blood pressure (MAP) maintained within one of two levels during surgery: in the range of 4555 mmHg or 5570 mmHg. The surveillance protocol for detecting perioperative changes in neuropsychologic performance was previously developed and tested for a trial comparing the effects of epidural and general anesthesia on cognitive outcomes after total knee replacement. 18 Methods Assembly of Patients All patients older than 50 yr who were undergoing elective unilateral primary THR with participating surgeons at the Hospital for Special Surgery between March 1993 and August 1995 were screened. All patients gave signed informed consent according to the Institutional Human Rights Committees approved protocol. Inclusion Criteria. To be eligible, patients had to be 70 yr or older or 5069 yr and have at least one of the following. Cardiac Disease. Cardiac disease was defined as S/P myocardial infarction (patients hospitalized for chest pain who had either new Q waves in at least two leads, new ST segment depression, or new T-wave inversion, with elevation of creatine kinase [CK] or CKMB isoenzyme levels 19 ); history of angina as defined by the Rose criteria 20; S/P coronary artery bypass grafting; or a history of congestive heart failure including pulmonary edema, paroxysmal nocturnal dyspnea, or dyspnea on exertion (by Rose criteria) that required continued pharmacologic therapy. Hypertension. Hypertension was defined as systolic pressure of >160 mmHg, diastolic pressure >95 mmHg, or treatment with a medication used explicitly to treat the patients blood pressure. 21 Diabetes Mellitus.

Diabetes mellitus was marked by treatment with insulin or oral hypoglycemic agents or an elevated fasting glucose level on more than one occasion (plasma >140 mg/dl; whole blood >120 mg/dl). Exclusion Criteria. Exclusion criteria consisted of the following. Contraindications to Epidural Anesthesia. Contraindications to epidural anesthesia were ankylosing spondylitis, bleeding diathesis, and use of systemic anticoagulants. Contraindications to Hypotensive Anesthesia. Hemodynamically significant aortic valve or mitral valve stenosis (documented by Doppler echocardiography or cardiac catheterization) and severe carotid artery stenosis (>70% occlusion) were contraindications to hypotensive anesthesia. Conditions Seriously Affecting Cognitive Testing Performance. Deafness, blindness, severe hand deformity or dysfunction, psychosis, and nonfluency in English were conditions that would affect cognitive testing performance and were therefore included as exclusion criteria. Allocation of Interventions A blocked randomization schedule was prepared before the start of the trial and was known only by the study statistician. Opaque allocation assignment envelopes were opened by the treating anesthesiologist in the operating room just before surgery. Preoperative Evaluation The preoperative evaluation performed 17 days before surgery assessed demographic status, including education and occupational history; medical, psychiatric, and surgical history, including past perioperative complications; medication use and substance abuse; physical examination; and preoperative laboratory values. The Charlson comorbidity score, a weighted index accounting for the number and seriousness of comorbid medical conditions, was computed for all patients. 22 Cognitive Assessment The perioperative cognitive assessment protocol has been described in detail, including the definition of a minimally important clinical difference in score for each test (appendix 1). The battery includes 10 tests: the Boston Naming, Controlled Word Association, Digit Symbol, Trail Making A and B, Digit Span, Benton Visual Retention, Benton Visual Recognition, MattisKovner Verbal Recall, and MattisKovner Verbal Recognition. Neuropsychologic testing was repeated 1 week and 4 months after operation. The Ammons Quick Test 23 for verbal IQ was performed before operation. Anesthesia Protocol No premedication was given. All patients received oxygen supplementation via nasal cannula. Oxygen saturation was monitored using disposable fingertip sensor pulse oximeters. Cardiac rate, rhythm, and waveform were monitored continuously using a displayed anterior chest lead. Continuous arterial systolic, diastolic, and MAPs were monitored via radial arterial lines. All patients had central venous catheters placed, and central venous pressures were transduced and displayed. Patients with a history of congestive heart failure or severe renal insufficiency also had pulmonary artery catheter monitoring.

Bupivacaine (2025 ml), 0.75%, was administered viaepidural catheter using standardized techniques. 24,25 Adjunctive medications for sedation included midazolam, fentanyl, and thiopental sodium. All patients received a low-dose intravenous epinephrine infusion at an infusion range of 15 g/min to maintain circulatory stability, as previously described. 16,17 Mean arterial blood pressure was maintained within the range of 4555 mmHg or 5570 mmHg during the operative procedure. These levels of hypotension resulted from the epidural anesthetic alone in all patients; no additional agents were used to decrease blood pressure. The MAP was stabilized in the target range by adjusting the epinephrine infusion rate and by intravenous infusion of Ringers lactate solution to replace lost blood and to maintain a stable central venous pressure, limited to a maximum of approximately 1.5 l crystalloid. If the MAP decreased to less than the target range despite a maximal infusion rate of 5 g/min epinephrine, an infusion of phenylephrine, boluses of intravenous ephedrine, or both were used to increase the MAP. At the end of the surgical procedure, the MAP was increased to 7075 mmHg in both groups using intravenous boluses of ephedrine. In the postanesthesia care unit, both groups had MAP maintained at more than 7075 mmHg with fluid replacement and intravenous boluses of ephedrine. Intraoperative Assessment and Data Collection One of the investigators (J.H.) created a customized software data collection system to record and analyze intraoperative surgical, anesthetic, and hemodynamic THR data for this study. A research assistant in the operating room entered events on a laptop computer using the program and pop-up menus. The events included incision, cementing, relocation, and so forth, and also all medications and fluids administered during the procedure. Simultaneously, the digital form of the physiologic data collected and analyzed by the anesthesia monitor was downloaded every 20 s from the monitor to the study computer via an RS 232 connection to allow exact temporal correlation of events and hemodynamic parameters. Downloading of intraoperative hemodynamic data was successful in 233 of 235 patients; in the other two cases, copies of operating room monitor trends data yielded adequate intraoperative MAP data. Intraoperative blood loss was measured using a standardized nursing protocol for THR. Blood loss was calculated as the sum of sponges weighed as they were passed off the surgical field plus the difference between irrigation and suction volumes. Postoperative Surveillance The standardized surveillance for cognitive and cardiovascular complications included at least once-daily examinations by study personnel from the postoperative anesthesia care unit through the seventh postoperative day or discharge. The postoperative examination focused on cardiopulmonary and neurologic status. Electrocardiograms were performed within 6 h after surgery (in the postanesthesia care unit) and on postoperative days 1, 2, and 3. 26 If significant electrocardiographic changes (as described in Cardiovascular Outcomes) were observed, cardiac isoenzymes were measured. Definition of Outcomes Change in Cognitive Function. The primary outcome of change in cognitive function was within-subject change in score for each neuropsychologic test (score at 4 months after surgery minus the score before surgery). The mean within-subject change in the two blood pressure groups was compared for each test. Change from baseline to 1 week after operation was also assessed to provide a basis for attributing changes observed at 4 months to the acute perioperative period. Other Outcomes. Cardiac complications included the perioperative occurrence of definite or probable myocardial infarction, pulmonary edema, cardiopulmonary arrest, or death. Definite postoperative myocardial infarction required a CK-MB level of > 5% and one of the following significant electrocardiographic changes: (1) new Q waves lasting more than 0.03 s and more than 1 mm in depth in more than two leads in the absence of a new conduction abnormality or a marked change in the QRS axis; and (2) T-wave and ST segment changes lasting more than 48 h in more than two leads (in the absence of new electrolyte

abnormalities or the new use of digitalis): new inversions of previously upright T waves; new ST depression of more than 1 mm; an additional 1 mm or more of ST depression if ST depression existed previously; previously depressed ST segments that returned to normal; or unequivocal ST elevation of more than 2 mm (excluding J point elevation). Changes in T waves from biphasic to inverted or vice versa were not counted as significant T-wave changes. Probable myocardial infarction required a CK-MB level of >3% and one of the significant electrocardiographic changes. Pulmonary edema required a pulmonary capillary wedge pressure >25 cm water or rales heard over two thirds of the lung fields with a typical chest radiograph for pulmonary edema. Renal dysfunction was defined as an increase in the serum creatinine of >20% that persisted for >48 h beginning in the first 3 days after surgery. In a previous study, this change had a true-positive rate of 67% in identifying patients whose decrease in postoperative creatinine clearance was >50%. 21 The outcome definition of postoperative delirium was based on an algorithm developed and used in two previous studies of older adults undergoing elective joint replacement. 18,27 The diagnosis required the presence of cognitive impairment of acute onset and fluctuating course and evidence of a significantly impaired attention disorder, plus at least two of the following signs: disorientation, disorganized thought, altered level of consciousness, hyperactive or hypoactive psychomotor activity, perceptual disturbances, or memory impairment. Blinding The anesthesiologists, surgeons, and study personnel recording data in the operating room could not be blinded to the type of blood pressure management. Because the purpose of blinding is to prevent bias in the evaluation of outcomes, the physicians assessing delirium and cardiac and renal outcomes were blinded to the intraoperative blood pressure range. Statistical Analyses Comparison between the two blood pressure groups of within-subject change in score on each of the 10 neuropsychologic tests was performed using the Student t test for two samples, using PROC TTEST, which is available in the Statistical Analysis System (SAS Institute, Cary, NC). The customary alpha level of significance of 0.05 was adjusted for the 10 different comparisons to P < 0.005 (using Bonferroni correction for multiple outcomes). All Pvalues were two tailed. The one-sample paired comparisons ttest was used to test for a significant change from before surgery to 1 week and 4 months after surgery. Examination of simultaneous effects of blood pressure level and other covariates thought to be of potential significance was performed with a multiple linear regression model using PROC GLM in SAS. The dependent variable in this model was the change from baseline to 4 months after operation on the 10 neuropsychologic tests. The Fisher exact test was used to compare the incidence of cardiovascular and other categorical outcomes. Results Summary of Patient Screening and Enrollment During the enrollment period, from March 1993 to August 1995, 461 patients were eligible. The most common reasons for ineligibility were nonparticipating surgeon, revision procedure, and not meeting entry criteria. In all, 235 patients agreed to participate. With regard to the effects of the entry criteria (age > 70 yr or age 5069 yr plus hypertension, cardiac disease, or diabetes), there was no difference between the younger and older groups in the prevalence of cardiac disease (30% and 27%, respectively) or diabetes (10% and 8%), although more of the younger patients were hypertensive (75% vs. 42%; P < 0.001). Baseline Preoperative Characteristics and Comparability One hundred seventeen patients were randomized to the 4555 MAP range and 118 patients to the 5570 MAP range. There were no statistically significant differences in baseline demographic and medical characteristics between the two groups (table 1). The age range of patients was 50 to 88 yr, with a mean of 72 yr; one half were women; one third of patients were working, either full or part time; and none of the patients resided in a long-term care institution. Table 2 shows the baseline

test scores before surgery on the neuropsychology tests. There were no statistically significant differences between the baseline scores of the two intervention groups on any of the 10 tests.

Table 1. Baseline Preoperative Demographic and Clinical CharacteristicsNo significant differences between groups.MAP = mean arterial pressure.

Table 2. Neuropsychologic Test Results at Preoperative BaselineValues are given as mean score (SD). No significant differences between groups. A higher score signifies better performance on all tests except trail making A and B, in which a longer completion time signifies worse performance.MAP = mean arterial pressure. Intraoperative Management All patients received epidural anesthesia except two patients who had inadvertent injections of 0.75% bupivacaine into the subarachnoid space resulting in a total spinal, which was diagnosed by shallow ventilation and a decrease in oxygen saturation to the low 90s. One patient had tracheal intubation and the other received assisted ventilation for 30 min. Neither patient was eliminated from the study. The overall mean intraoperative MAP in the lower MAP range group was 50 3.6 (SD) mmHg, compared with 65 4.8 (SD) mmHg in the higher MAP group (P < 0.0001). Variability around the mean MAP was minimal, with a mean standard deviation around the mean MAP of 3.2 mmHg in the lower and 4.4 mmHg in the higher MAP group. In the 4555 MAP group, less than 14% of all MAP measurements exceeded the upper limit of 55 mmHg. In the 5570 MAP group, only 2.5% of all MAP measurements were less than the lower limit of 55 mmHg. The mean duration of surgery in both groups was 75 min.

As expected, the percentage of patients who received a phenylephrine infusion was significantly greater in the higher pressure range group at 50%, compared with 7% in the lower pressure range group. Patients in the higher pressure range group received significantly more intraoperative fluid: 1,750 ml Ringers lactate solution compared with 1,600 ml in the lower pressure range group (P = 0.022). Compliance with Follow-up In-hospital outcome surveillance was completed for all 235 patients. Long-term follow-up interviews were completed by 216 patients (92%). Nineteen patients were not available for follow-up, with equal attrition rates in the two MAP groups. The reasons for attrition included two deaths from cancer; one patient with severe intercurrent illness caused by cancer diagnosed at the time of THR; two patients had moved out of the area; and 14 patients refused. Cognitive Outcomes Comparison by Intraoperative Pressure Range. Table 3 shows the results for within-subject change in score at both postoperative time points for each of the 10 neuropsychological tests. There were no significant differences between the two intraoperative blood pressure groups from baseline to 4 months after operation, with a power of more than 99% (at a two-sided alpha level of 0.005) to detect a difference between the two arms of the trial greater than or equal to the minimally clinically important difference for each test. Similarly, analysis of within-patient change from baseline to 1 week after operation revealed no significant differences between the two groups.

Table 3. Cognitive Outcome: Within-subject Change from Preoperative Baseline ScoreValues are given as mean change score (SD). No significant differences in change scores between groups at either time point. A positive change represents improvement on all tests except the trail making A and B tests, for which a positive value of change represents poorer performance.MAP = mean arterial pressure. Multivariate Analysis. Increased age was not a significant predictor of decline in neuropsychologic performance 4 months after surgery on any of the 10 tests but was predictive of poorer performance on one test 1 week after surgery (Controlled Word Association, P < 0.005). However, age explained only 4% of the observed variance (r2 = 0.04). Multivariate analysis of age, blood pressure group, and an ageMAP group interaction term showed no significant interactive effect on any of the 10 neuropsychologic tests. Intraoperative variables such as duration of surgery, mean MAP during surgery, and the product of the duration of surgery and mean MAP were not predictive of long-term cognitive declines. Noncognitive postoperative complications were also not predictive of long-term cognitive declines.

Cognitive Outcome: Temporal Patterns. There was a generalized pattern of decline at 1 week after surgery, followed by a return to baseline levels or an improvement on most tests by 4 months after surgery. A significant transient deterioration in performance was observed at 1 week compared with baseline on two of the 10 tests: Digit Symbol, and MattisKovner verbal recognition (by paired t test, P < 0.001). Other Outcomes. No in-hospital deaths occurred. Table 4 shows the incidence of noncognitive outcomes by intervention group. The incidence of myocardial infarction or ischemia was similar in the two groups, at 6% in the 4555 MAP and 4% in the 5570 MAP groups. Only one of these events was a new Q-wave myocardial infarct; this occurred in the higher MAP group. There were no cerebrovascular accidents. The power of this study to detect a 5% difference in cardiovascular complications was 85%.

Table 4. Noncognitive OutcomesNo significant differences between groups.LOS = length of stay.*n = 111 for the 4555 MAP group and n = 108 for the 5570 MAP group. The incidence of postoperative renal dysfunction was not significantly different, at 1% in the 4555 MAP and 4% in the 55 70 MAP groups. No patient experienced persistent renal failure; one patient in the higher MAP group had renal dysfunction associated with transient acute renal failure. The onset of transient renal failure occurred late in the first week after surgery and was deemed to have been secondary to postoperative complications and management. The power of this study to detect a 5% difference in renal complications was 82%.

Overall, the incidence of delirium was 7%: 9% in the lower pressure arm compared with 4% in the higher pressure arm (P = 0.3). The significant positive predictors of delirium were male sex (11% of men compared with 3% of women) and poorer preoperative performance on 7 of the 10 neuropsychology tests. Age was not a significant predictor of delirium. No patient with delirium progressed to stupor, coma, or death, nor did they have an increased rate of other postoperative complications or long-term cognitive deterioration. The length of stay was greater in patients with delirium, at 10.9 days compared with 8.2 days (P < 0.03). The power of this study to detect a 6% difference in the incidence of delirium was 62%, compared with a 93% power to detect a 10% difference. As shown in table 4, no significant difference occurred in intraoperative blood loss, with mean blood losses of 199 ml and 212 ml in the lower and higher MAP groups, respectively. There were also no significant differences between the two groups in the mean total number of units transfused or the percentage of patients who received any autologous units, any allogeneic units, or no blood transfusion. There were also no significant differences between the groups in length of hospital stay. A postoperative lower extremity Doppler study or a venogram was performed in 219 of the 235 patients. The overall rate of distal lower extremity clot was 8%: 11% in the lower and 6% in the higher MAP groups (P = 0.12). No proximal clots or pulmonary emboli were found. Discussion In this study, we found no difference in long-term postoperative cognitive deterioration between patients randomized to a markedly hypotensive intraoperative MAP range of 4555 mmHg or a mildly to moderately hypotensive range of 5570 mmHg during epidural anesthesia for THR. The power of the study to have detected an important deterioration on any one of the neuropsychologic tests in the study battery was more than 99%. There were also no differences in the rates of early cognitive dysfunction, delirium, adverse cardiovascular outcomes, renal dysfunction, or thromboembolism. The overall incidence of major cardiovascular complications was low, despite the high prevalence of comorbid cardiovascular risk factors in our study population. The concern was that systemic hypotension could lead to brain hypoperfusion and subtle permanent cognitive dysfunction. The physiologic effect of the decrease in systemic pressure depends on the lower limit for cerebral autoregulation, which is higher in patients with hypertension. 6,11,28 The evidence provided by this study suggests that elderly patients, including those with documented cardiovascular disease, can safely tolerate a period of significant systemic hypotension under the controlled conditions of this anesthesia management protocol. This does not include patients with occlusive carotid disease (>70% occlusion) or hemodynamically significant aortic valve or mitral valve stenosis, who were specifically excluded from participating in the study. Complication rates after elective THR observed in this study of high-risk patients are similar to or lower than those of previous studies of elective hip and knee replacement in unselected patients receiving normotensive anesthesia. 18,2931 Several previous studies specifically addressed the safety of deliberate hypotension for THR, but they had limited generalizability or inadequate outcome surveillance. For example, one randomized study found no difference in complications between patients given hypotensive or normotensive anesthesia, but it excluded patients with a history of stroke, transient ischemic attack, myocardial infarct, renal disease, or poorly controlled hypertension. 2 Another study of hypotensive anesthesia for THR did not find any evidence of increased complications, but it considered only death or major stroke. 8 A large case series at our own institution found no evidence of an increase in adverse events, even in elderly hypertensive patients, but it was not a controlled trial with a strict complication surveillance protocol. 31 Reported intraoperative blood loss during THR with normotension varies from 500 to 1,800 ml. 2,17,3234 Reduced blood loss has been reported in patients receiving spinal and epidural anesthesia compared with general anesthesia and in patients with controlled hypotension. 2,35,36 Orthopedic surgeons strongly believe that the relatively bloodless field provided by hypotensive anesthesia facilitates their performance of the procedure because anatomic structures can be better seen. 2 This trial, however, did not find evidence of clinically or statistically significant differences between the mildly and markedly hypotensive groups in intraoperative blood loss, duration of surgery, or transfusions of either autologous or nonautologous blood. The lack of a difference in estimated blood loss may be due to imprecision in the measurement technique, particularly because blood loss in the higher MAP group was already reduced to an average of 212 ml. 37 It is important to note that both

levels of intraoperative MAP were relatively hypotensive compared to patients preoperative MAP. More than 95% of patients had a mean intraoperative MAP that was at least 20 mmHg less than their preoperative MAP. The benefits of controlled reduction of intraoperative MAP on these outcomes may reach a plateau beyond which further benefit does not occur. One potential benefit not tested in this study is that a drier surgical field may reduce the amount of blood at the cementbone interface, thus improving the quality of fixation of the prosthesis to the bone. 3841 Aseptic loosening of the prosthesis from the bone is the most common cause of failure of a cemented THR. 42 The surgeons in the study could discern differences in the wetness of the field as a function of the level of intraoperative blood pressure (P < 0.001), but prosthesis loosening cannot be assessed until follow-up 510 yr after surgery. Different techniques for inducing intraoperative hypotension may be associated with different risks of hypoperfusion injury and complication rates, 4,43 as described for patients receiving high levels of inspired isoflurane, 44 intravenous nitroprusside, 4548 autonomic blocking agents, 1214 or combinations of these agents. This study, in contrast, achieved hypotension via epidural anesthesia, which produces a chemical sympathectomy with profound peripheral venodilatation. The epinephrine infusion does not directly affect MAP, 16 but maintains normal stroke volumes, is associated with improved cardiac indices without increases in preload, and may prevent significant bradycardia. 16 Therefore, although diastolic perfusion pressure decreases, the work and metabolic demands of the heart are reduced. This method of deliberate hypotensive epidural anesthesia appears to have maintained blood flow and oxygen delivery at adequate levels despite low systemic MAPs. An important caveat of this study is that the demonstrated safety of hypotensive epidural anesthesia in older adults with comorbid diseases is limited to the protocol described in this study, including the use of continuous hemodynamic monitoring, supplemental oxygen, and avoidance of hypovolemia. The results are not necessarily generalizable to other techniques of hypotensive anesthesia. In conclusion, this randomized trial found no differences between two different levels of intraoperative blood pressure management during epidural anesthesia for THR, specifically MAP ranges of 4555 and 5570 mmHg, in early and longterm cognitive, cardiac, and renal complications in elderly patients. References 1. NIH Consensus Development Panel on Total Hip Replacement. Total hip replacement. JAMA 1995; 273: 19506 [Fulltext Link] [CINAHL Link] [Context Link] 2. Thompson GE, Miller RD, Stevens WC, Murray WR: Hypotensive anesthesia for total hip arthroplasty. A NESTHESIOLOGY 1978; 48: 916 [Medline Link] [Context Link] 3. Barbier-Bohm G, Desmonts JM, Couderc E, Moulin D, Prokocimer P, Oliver H: Comparative effects of induced hyotension and normovolaemic haemodilution on blood loss in total hip arthroplasty. Br J Anaesthesia 1980; 52: 103943 [Context Link] 4. Green DW: Cardiac and cerebral complications of deliberate hypotension, Hypotensive Anesthesia. Edited by Enderby GEH. Edinburgh, Churchill Livingstone, 1985, pp 23661q;3> [Context Link] 5. Lindop MJ: Complications and morbidity of controlled hypotension. Br J Anaesth 1975; 47: 799803 [Medline Link] [Context Link] 6. Jones MJT: The influence of anesthetic methods on mental function. Acta Chir Scand Suppl 1988; 550: 16976 [Medline Link] [Context Link] 7. Charlson ME, MacKenzie CR, Gold JP, Ales KL, Topkins M, Fairclough GP Jr, Shires GT: Intraoperative blood pressure. Ann Surg 1990; 212: 56780 [Medline Link] [Context Link]

8. Davis NH, Jennings JJ, Harris WH: Induced hypotensive anesthesia for total hip replacement. Clin Orthop Rel Res 1974; 101: 938 [Medline Link] [Context Link] 9. Rollason WN, Robertson GS, Cordiner CM, Hall DJ: A comparison of mental function in relation to hypotensive and normotensive anaesthesia in the elderly. Br J Anaesth 1971; 41: 56165 [Medline Link] [Context Link] 10. Roach GW, Kanchuger M, Mangano CM, Newman M, Nussmeier N, Wolman R, Aggarwal A, Marschall K, Graham SH, Ley C: Adverse cerebral outcomes after coronary bypass surgery. N Engl J Med 1996; 335: 185763 [Fulltext Link] [Medline Link] [Context Link] 11. Gold JP, Charlson ME, Williams-Russo P, Szatrowski TP, Peterson JC, Pirraglia PA, Hartman GS, Yao FSF, Hollenberg JP, Barbut D, Hayes JG, Thomas SJ, Purcell MH, Mattis S, Gorkin L, Post M, Krieger KH: Improvement of outcomes after coronary artery bypass: A randomized trial comparing intraoperative high versus low mean arterial pressure. J Thorac Cardiovasc Surg 1995; 110: 130214 [Medline Link] [Context Link] 12. Eckenhoff JE, Compton JR, Larson A, Davies RM: Assessment of cerebral effects of deliberate hypotension by psychological measurements. Lancet 1964; ii: 71114 [Context Link] 13. Townes BD, Dikmen SS, Bledsoe SW, Hornbein TF, Martin DC, Janesheski JA: Neuropsychological changes in a young, healthy population after controlled hypotensive anesthesia. Anesth Analg 1986; 65: 9559 [Medline Link] [Context Link] 14. Toivonen J, Kukka P, Kaukinen S: Effects of deliberate hypotension induced by labetalol with isoflurane on neuropsychological function. Acta Anaesth Scand 1993; 37: 711 [Medline Link] [Context Link] 15. Bembridge JL, Moss E, Grummitt RM, Noble J: Comparison of propofol with enflurane during hypotensive anaesthesia for middle ear surgery. Br J Anaesthesia 1993; 71: 8957 [Context Link] 16. Sharrock NE, Urquhart B, Mineo R: Hemodynamic response to low-dose epinephrine infusion during hypotensive epidural anesthesia for total hip replacement. Reg Anesth 1990; 15: 2959 [Medline Link] [Context Link] 17. Sharrock NE, Salvati EA: Hypotensive epidural anesthesia for total hip arthroplasty. Acta Orthop Scand 1996; 67: 91 107 [Medline Link] [Context Link] 18. Williams-Russo PG, Sharrock NE, Mattis S, Szatrowski TP, Charlson ME: Cognitive effects after epidural versus general anesthesia in older adults: A randomized trial. JAMA 1995; 274: 4450 [Fulltext Link] [Medline Link] [Context Link] 19. Goldman L, Caldera D, Nussbaum S, Southwick FS, Krogstad D, Murray B, Burke DS, OMalley TA, Goroll AH, Caplan CH, Nolan J, Carabello B, Slater EE: Multifactorial index of cardiac risk in non-cardiac surgical procedures. N Engl J Med 1977; 297: 84550 [Medline Link] [Context Link] 20. Rose GA, Blackburn H: Cardiovascular Survey Methods. Belgium, World Health Organization Monograph, 1968 [Context Link] 21. Charlson ME, MacKenzie CR, Gold JP, Ales KL, Shires GT: Postoperative renal dysfunction can be predicted. Surg Gynecol Obstet 1989; 169; 3039 [Medline Link] [Context Link] 22. Charlson ME, Pompei P, Ales KL, MacKenzie CR: A new method of classifying prognostic comorbidity in longitudinal studies: Development and validation. Journal of Chronic Diseases 1987; 40: 37383 [Medline Link] [Context Link] 23. Levine NR: Validation of the Quick Test for intelligence screening of the elderly. Psychol Rep 1972; 29: 16772 [Medline Link] [Context Link]

24. Sharrock NE: Epidural dose responses in patients aged 20-80. A NESTHESIOLOGY 1978; 49: 4258 [Medline Link] [Context Link] 25. Sharrock NE, Lesser ML, Gabel RA: Segmental levels of anaesthesia following the extra-dural injection of 0.75% bupivacaine at different lumbar spaces in elderly patients. Br J Anaesth 1984; 56: 2857 [Medline Link] [Context Link] 26. Charlson ME, MacKenzie CR, Ales KA, Gold JP, Fairclough GP Jr, Shires GT: Monitoring the postoperative electrocardiogram: Implications for detection and diagnosis of postoperative myocardial infarction in noncardiac surgery. J Clin Epidemiol 1989; 42: 2534 [Medline Link] [Context Link] 27. Williams-Russo PG, Urquhart B, Sharrock N, Charlson ME: Post-operative delirium in an elderly orthopedic population: Predictors and prognosis. J Am Geriatr Soc 1992; 40: 75967 [Medline Link] [Context Link] 28. Lassen NA, Astrup J: Cerebral blood flow: Normal regulation and ischemic thresholds, Protection of the Brain from Ischemia. Edited by Weinstein PR, Faden AI. Baltimore, Williams & Wilkins, 1990 [Context Link] 29. Baron JA, Barrett J, Katz JN, Liang MH: Total hip arthroplasty: Use and select complications in the US Medicare Population. Am J Public Health 1996; 86: 702 [Fulltext Link] [Medline Link] [Context Link] 30. Lynch NM, Trousdale RT, Ilstrup DM: Complications after concomitant bilateral total knee arthroplasty in elderly patients. Mayo Clin Proc 1997; 72: 799805 [Fulltext Link] [Medline Link] [Context Link] 31. Sharrock NE, Cazan MG, Hargett JL, Williams-Russo P, Wilson PD: Changes in mortality after total hip and knee arthroplasty over a ten-year period. Anesth Analg 1995; 80: 2428 [Fulltext Link] [Medline Link] [Context Link] 32. Covert CR, Fox GS: Anaesthesia for hip surgery in the elderly. Can J Anaesth 1989; 36: 31119 [Medline Link] [Context Link] 33. Keith I: Anaesthesia and blood loss in total hip replacement. Anaesthesia 1977; 32: 44450 [Medline Link] [Context Link] 34. Sharrock NE, Mineo R, Urquhart B: Haemodynamic effects and outcome analysis of hypotensive extradural anaesthesia in controlled hypertensive patients undergoing total hip arthroplasty. Br J Anesth 1991; 67: 1725 [Medline Link] [Context Link] 35. Modig J, Borg T, Karlstrom G, Maripuu E, Sahlstedt B: Thromboembolism after total hip replacement: Role of epidural and general anesthesia. Anesth Analg 1983; 62: 17480 [Medline Link] [Context Link] 36. Chin SP, Abou-Madi MN, Eurin B, Witvoet J, Montagne J: Blood loss in total hip replacement: Extradural v phenoperidine analgesia. Br J Anaesth 1982; 54: 4914 [Medline Link] [Context Link] 37. Sharrock NE, Mineo R, Urquhart B, Salvati EA: The effect of two levels of hypotension on intraoperative blood loss during total hip arthroplasty performed under lumbar epidural anesthesia. Anesth Analg 1993; 76: 5804 [Medline Link] [Context Link] 38. Ranawat CS, Beaver WB, Sharrock NE, Maynard MJ, Urquhart B, Schneider R: Effect of hypotensive epidural anaesthesia on acetabular cement-bone fixation in total hip arthroplasty. J Bone Joint Surg (Br) 1991; 73-B: 77982 [Medline Link] [Context Link] 39. Benjamin JB, Gie GA, Lee AJ, Ling RS, Volz RG: Cementing technique and the effects of bleeding. J Bone Joint Surg (Br) 1987; 69-B: 6204 [Medline Link] [Context Link]

40. Eftekhar NS, Nercessian O: Incidence and mechanism of failure of cemented acetabular component in total hip arthroplasty. Orthop Clin North Am 1988; 19: 55766 [Medline Link] [Context Link] 41. Bannister GC, Young SK, Baker AS, Mackinnon JG, Magnusson PA: Control of bleeding in cemented arthroplasty. J Bone Joint Surg (Br) 1990; 72-B: 4446 [Medline Link] [Context Link] 42. Kavanagh BF, Dewitz MA, Ilstrup DM, Stauffer RN, Coventry MB: Charnley total hip arthroplasty with cement: Fifteen year results. J Bone Joint Surg (Am) 1989; 71-A: 1496503 [Medline Link] [Context Link] 43. Moss E: Cerebral blood flow during induced hypotension. Br J Anaesth 1995; 74: 6357 [Medline Link] [Context Link] 44. Madsen JB, Cold GE, Hansen ES, Bardrum B, Kruse-Larsen C: Cerebral blood flow and metabolism during isoflurane induced hypotension in patients subjected to surgery for cerebral aneurysms. Br J Anaesth 1987; 59: 12047 [Medline Link] [Context Link] 45. Bernard JM, Pinaud M, Francois T, Babin M, Macquin-Mavier I, Letenneur J: Deliberate hypotension with nicardipine or nitroprusside during total hip arthroplasty. Anesth Analg 1991; 73: 3415 [Medline Link] [Context Link] 46. Vazeery AK, Lunde O: Controlled hypotension in hip joint surgery. Acta Orthop Scand 1979; 50: 433441 [Medline Link] [Context Link] 47. Bunemann L, Jensen K, Thomsen L, Riisager S: Cerebral blood flow and metabolism during controlled hypotension with sodium-nitroprusside and general anaesthesia for total hip replacement a.m. Charnley. Acta Anaesthesiol Scand 1987; 31: 48790 [Medline Link] [Context Link] 48. Vesey CJ, Cole PV, Simpson PJ: Cyanide and thiocyanate concentrations following sodium nitroprusside infusion in man. Br J Anaesth 1976; 48: 65160 [Medline Link] [Context Link] 49. Kaplan EF, Goodglass H, Weintraub S: The Boston Naming Test. Boston, E Kaplan and H Goodglass, 1978 [Context Link] 50. Benton AL, Hamsher K: Multilingual Aphasia Examination. Iowa City, University of Iowa Hospitals, 1976 [Context Link] 51. Wechsler D: Instruction Manual for the Wechsler Memory ScaleRevised. New York, Psychological Corp., 1987 [Context Link] 52. Army Individual Test Battery: Manual of Directions and Scoring. Washington, DC, War Department, Adjutant Generals Office, 1944 [Context Link] 53. Benton AL: The Revised Visual Retention Test, 4th ed. New York, Psychological Corp., 1974 [Context Link] 54. Mattis S, Kovner R, Goldmeier E: Different patterns of mnemonic deficit in two organic amnestic syndromes. Brain Lang 1987; 6: 17991 [Context Link] 55. Swets JA, ed: Signal Detection and Recognition by Human Observers. New York, John Wiley & Sons, 1964 [Context Link] Appendix 1. The Neuropsychologic Battery Linguistic Domain Boston Naming Test. 49

Patients are asked to name a sequence of pictured objects. The 60-picture version was split into odd and even items. Possible scores would range from 0 to 30. The minimal clinically important difference (CID) in score on repeated administrations is 4 or more. Controlled Word Association. 50 This is a test of oral word fluency consisting of three 1-min trials of production of spoken words beginning with a specific letter. Possible scores range from 0 to 70 (CID >= 13). PsychomotorAttention Domain Digit Symbol. 51 A subtest of the Wechsler Adult Intelligence Scale, Revised (WAIS-R), Digit Symbol tests psychomotor performance and speed and requires symbol substitution and copying. Possible scores range from 0 to 93 (CID >= 12). Trail Making Tests A and B. 52 These tests measure visuomotor speed, conceptual tracking, and attention. Part A requires patients to connect circles containing numbers in ascending numeric order. Part B requires patients to alternate between numbered circles and lettered circles. Possible scores range from 0 to 300 s (Part A, CID >= 22 s; Part B, CID >= 85 s). Digit Span. 51 This subtest of the WAIS-R combines Digits Forward and Digits Backward. The tester reads aloud progressively longer series of digits, and the patient repeats each series from memory. After reaching their maximum series forward, the task is changed to repeat the digit series back in reverse sequence. Possible scores range from 0 to 28 (CID >= 4). Memory Domain Benton Visual Retention. 53 This measures immediate visual recall memory and visuospatial ability. Patients are presented with a series of 10 progressively more complex geometric designs and asked to draw each design from memory. Possible scores range from 0 to 10 (CID >= 2). Benton Visual Recognition. This is a test of visual recognition memory performed in tandem with the visual retention test. After each recall drawing, patients are shown four designs and asked to select the correct original. Possible scores range from 0 to 10 (CID >= 2). MattisKovner Verbal Recall and Verbal Recognition. 54 This measures verbal recall and verbal recognition memory. The recall test consists of repeated trials to learn a 20-word list of words in a common category. The recall score is the best percentage of words correctly remembered. A probe list of 40 words containing the 20 words on the recall learning list plus an additional 20 words is then given for recognition. The recognition score is derived from the percentage of true positives and the percentage of false positives, which are used to compute the signal detection theory measure of accuracy, d'. 55 Possible recall scores range from 0 to 100 (CID >= 15). Possible recognition scores range from 0 to 3.93 (CID >= 0.7). Key words:: Computerized anesthesia record; geriatrics; memory; neuropsychologic; postoperative.

The Efficacy and Safety of Epidural Infusions of Levobupivacaine With and Without Clonidine for Postoperative Pain Relief in Patients Undergoing Total Hip Replacement
Kevin R. Milligan, PhD, Philip N. Convery, and Denis Connolly, FFARCSI
FRCA,

Paul Weir,

FFARCSI,

Patricia Quinn,

RGN,

Department of Anaesthesia, Musgrave Park Orthopaedic Hospital, Belfast, Northern Ireland

We assessed the efficacy and tolerability of epidural infusions of levobupivacaine, levobupivacaine plus clonidine, and clonidine for postoperative analgesia in 86 patients undergoing total hip replacement. For each group, an epidural cannula was inserted before surgery and 15 mL of 0.75% plain levobupivacaine was administered. Three hours later, an epidural infusion (6 mL/h) of levobupivacaine 0.125% (L), levobupivacaine 0.125% plus clonidine 8.3 g/mL (LC) or clonidine alone (8.3 g/mL) (C) was initiated. Morphine consumption was recorded for the following 24 h as were visual analog pain scores and the degree of sensory and motor blockade. The mean (median) morphine consumption was lowest in the combination

group (LC),14 (7) mg; higher in the clonidine group (C), 23 (21) mg; and highest in the levobupivacaine group (L), 37 (36) mg (P 0.022). The median times until the first request for analgesia which were 2.9, 5.9, and 12.5 h for Groups L, C, and LC, respectively (P 0.01). There were no statistical differences among the groups regarding the maximum degree of postoperative motor blockade. On average, the systolic blood pressure in the two clonidine groups was slightly lower than in those from the levobupivacaine group. We conclude that the epidural administration of a combination of levobupivacaine plus clonidine is well tolerated and gives better analgesia than either drug used alone. (Anesth Analg 2000;91:3937)

upivacaine is a long-acting, effective local anesthetic that is commonly administered by the epidural route for the relief of postoperative pain. Despite its undoubted efficacy, bupivacaine is associated with cardio- and neurotoxicity, and this has occasionally resulted in death (1). The bupivacaine molecule is a racemate, and in the proprietary solution, two enantiomers, levo-(S[-]) and dex- (R[]) bupivacaine are present in a 50:50 ratio. Data indicate that, whereas levobupivacaine and racemic bupivacaine have a similar clinical efficacy, levobupivacaine has a reduced potential for producing toxicity (2,3). Clonidine is an 2-adrenergic agonist which, when administered by the epidural route, has analgesic properties and potentiates

the effects of local anesthetics (4). The use of clonidinelevobupivacaine mixtures has not been previously described. The aim of this study was to investigate the safety and the analgesic properties of levobupivacaine 0.125% alone and in combination with clonidine when administered as an epidural infusion for the treatment of postoperative pain in patients undergoing total hip replacement.

Methods
The study was approved by The Queens University of Belfast Research Ethics Committee, and written, informed consent was obtained from all patients. Ninety patients, ASA physical status IIII, between 18 and 80 yr old, weighing 50 110 kg, and presenting for elective primary total hip replacement were recruited. Patients taking medications with adrenergic or psychotropic activity; receiving chronic analgesic therapy other than simple analgesics (acetaminophen/codeine, nonsteroidal antiinflammatories); with a history of neurological or neuromuscular disorders, drug or alcohol
Anesth Analg 2000;91:3937

Drugs and financial support for this study were supplied by Chiroscience Ltd., Cambridge, UK. Accepted for publication April 24, 2000. Address correspondence and requests for reprints to Dr. K. R. Milligan, Department of Anaesthetics, Musgrave Park Hospital, Stockmans Lane, Belfast BT9 7JB, N. Ireland.
2000 by the International Anesthesia Research Society 0003-2999/00

393

394

REGIONAL ANESTHESIA AND PAIN MEDICINE MILLIGAN ET AL. EPIDURALS, LEVOBUPIVACAINE, AND CLONIDINE

ANESTH ANALG 2000;91:3937

abuse; or in whom there was a contraindication to regional anesthesia were excluded from the study, as were women of child-bearing potential. During the preoperative visit, patients were tutored in the use of a 100-mm visual analog scale (VAS) marked no pain at one end and worst pain imaginable at the other. The use of a patient-controlled analgesia (PCA) system to provide supplementary postoperative analgesia was also explained. Patients were premedicated orally with temazepam 20 mg and ranitidine 150 mg 11.5 h preoperatively. On arrival in the operating room, IV access was established and an infusion of dextran 70 initiated as prophylaxis against deep venous thrombosis. The patients were then placed in the lateral decubitus position. The skin over the proposed epidural site (L2-3 or L3-4) was infiltrated with 3 mL of lidocaine, 1% plain. A 16-gauge epidural catheter was placed 35 cm into the epidural space, and, after a negative aspiration, a 3-mL test dose of lidocaine 2% with epinephrine 1:200,000 was injected. Provided there was no evidence of subarachnoid or intravascular injection, three 5-mL increments of levobupivacaine 0.75% were administered at 5-min intervals. At the completion of the final 5-mL injection, sensory levels were assessed every 5 min by using loss of sensation to cold until the sensory block reached T10, at which point surgery proceeded. If this level was not achieved after 30 min, the patient received further 1-mL increments of levobupivacaine 0.75% up to a maximum of 5 mL. If an adequate block was not achieved after a further 15-min period, the patient was withdrawn from the study. In this situation, the patients study number was reallocated to the next study patient. Motor block was assessed by using a modified Bromage scale (see below) at 0, 10, 20, and 30 min after the completion of the epidural injection until full motor block was achieved or surgery commenced. Patients received IM droperidol 2 mg and cefamandole 2 g IV before surgery as antiemetic and antibiotic prophylaxis, respectively. Routine monitoring consisted of electrocardiogram and noninvasive blood pressure monitoring. Respiratory rate and pulse oximetry was done on all patients. Intraoperative sedation was provided by a propofol infusion, titrated to effect, the patients being drowsy but arousable. Supplemental oxygen was provided via a face mask. Blood and fluids were replaced as clinically indicated, and hypotension, defined as a decrease of more than 30% of baseline systolic pressure, was treated with IV fluids and ephedrine as necessary. Postoperatively patients were connected to a PCA programmed to deliver morphine 1 mg IV on demand, with a 5-min lockout. The time to the first demand for analgesia, the total morphine consumption, and the number of requests for analgesia were recorded. Three hours after the completion of the final epidural injection, the patients were randomly allocated

Table 1. Patient Characteristics


Levobupivacaine Sex (M/F) Age (yr) Weight (kg) Height (cm) 13/17 65 10 75 12 164 8 Levobupivacaine/ clonidine Clonidine 6/24 65 9 73 14 160 8 13/17 65 9 73 12 164 9

Values are mean sd.

to one of three study groups, all of whom received an epidural infusion which ran at 6 mL/h for 24 h. These were: levobupivacaine 0.125% (Group L); levobupivacaine 0.125% plus clonidine 8.3 g/mL (Group LC); or clonidine alone 8.3 g/mL (Group C). The solutions were prepared in the pharmacy department, and the patient and the medical/nursing attendants were not aware of which solution the patient was receiving. VAS scores were recorded hourly for the first 12 h after the start of the infusion and every 2 h thereafter up to 24 h. Recordings were taken both at rest and on passive movement of the operated leg. Sensory and motor block were assessed hourly for 24 h. Motor block was assessed on the nonoperated leg by using a modified Bromage scale where 0 no paralysis, full flexion of the knee and ankle; 1 inability to raise extended leg, able to move knee; 2 inability to flex knee, able to flex ankle; and 3 inability to move lower limb. All patients received routine postoperative clinical monitoring and all adverse events were recorded. Statistical analysis of the data was carried out by using the SAS-PC-Windows package (Version 6.10; Chiroscience, Cambridge, UK). The time until the first request for analgesia was analyzed by using the Wilcoxon test using survival analysis techniques, and the number of requests for analgesia was analyzed by using the Wilcoxon two-sample test on the basis that data were nonnormal. Motor block was analyzed by using logistical regression. A P value 0.05 was considered significant. In a previous study (5), the mean dose of morphine delivered to patients receiving an infusion of 50 g/mL clonidine was 10.5 mg with a standard deviation of 6.0 mg. Based on these estimates, 0.017 (i.e., adjusting for multiple comparisons) and 0.2, the number of patients required to detect a difference of 50% from the clonidine mean was 30 patients per group.

Results
Surgery and follow up were successfully completed for 86 of the 90 patients: 27 from Group L, 30 from Group LC and 29 from Group C. The groups were similar in terms of the patient characteristics except for

ANESTH ANALG 2000;91:3937

REGIONAL ANESTHESIA AND PAIN MEDICINE MILLIGAN ET AL. EPIDURALS, LEVOBUPIVACAINE, AND CLONIDINE

395

Table 2. Time to First Request for Morphine, Total Morphine Consumption, and Total Morphine Requests Levobupivacaine (n 27) Time to first morphine (h) Total morphine consumption (mg) Total no. of morphine requests (median/range)
Values are mean sd. * P 0.05 (levobupivacaine vs clonidine). P 0.01 (levobupivacaine vs levobupivacaine/clonidine). P 0.01 (vs other two groups).

Levobupivacaine/ clonidine (n 30) 13.0 8.3 13.9 17.3 29 41 (9/0170)

Clonidine (n 29) 7.2 5.9 21.8 12.3 46 46 (28/0183)

5.0 5.6* 34.9 22.7* 68 56 (55/5185)

Table 3. Mean Postoperative Visual Analog Pain Scores (mm) on Passive Movement Time (h) 0 4 8 12 16 20 24 h h h h h h h Levobupivacaine 4.1 16.4 (080) 22.3 24.9 (072) 8.5 13.9 (048) 13.8 23.2 (078) 8.0 11.3 (030) 12.8 19.7 (070) 7.7 12.3 (045) Levobupivacaine/ clonidine 3.6 11.0 (050) 2.5 6.0 (020) 1.3 5.4 (026) 1.3 5.0 (020) 5.4 10.9 (040) 12.6 21.4 (084) 9.4 13.8 (041) Clonidine 1.0 5.5 (030) 7.4 16.6 (065) 12.4 19.9 (070) 12.5 18.2 (050) 25.2 25.7 (080) 19.4 20.6 (055) 16.2 26.4 (0100)

Values are mean sd (range).

Group LC, which contained proportionately more women than the other two groups (Table 1). The onset of sensory blockade was rapid, with most patients achieving a sensory block above the level of T10 by the time the final 5 mL of epidural levobupivacaine had been administered. In six patients, the T10 level was not reached until the second assessment, 5 min after the completion of the epidural injection. In four cases (two from Group L and two from C), the patients received an additional 5 mL of epidural levobupivacaine as the block had not reached T10 after 30 min. This level was achieved within 5 min of the final 5 mL in all four patients. The onset of motor block was rapid, and assessment immediately after the final 5-mL epidural injection showed that one patient (Group C) already had full (grade 3) motor block, while the remainder had grade 0 (n 39), grade 1 (n 36), or grade 2 (n 14) blocks. There were no differences among the groups in terms of their intraoperative course: the duration of surgery was 68 11, 70 12, and 68 13 (mean sd) min, and the intraoperative blood losses were 502 327, 532 249, and 539 400 (mean sd) mL for Groups L, C, and LC, respectively. The time until the first request for analgesia and the total postoperative morphine requirements were significantly different among the three groups (Table 2), Group LC having the lowest requirements and the longest interval before analgesia was requested. Morphine consumption by the patients in Group C was significantly higher than in Group LC patients but significantly lower than that of patients in Group L,

who also had the shortest period until analgesia was requested. The VAS scores recorded on passive movement of the operated limb are summarized in Table 3. These were generally satisfactory for all groups, the mean score being below 30 mm and the median score 0 mm at almost every assessment. Group LC had the lowest scores at most of the assessments, but this difference did not achieve statistical significance. In all groups, the patients appeared to have had effective pain relief. The pattern of postoperative sensory and motor blockade was different among the three groups (Figs. 1 and 2). The motor and sensory block in Group C regressed rapidly. The motor block in Group LC was much more intense than that of the other two groups. However, the motor block was not profound in the majority of patients in the LC group and was comparable to that of the other two groups by the conclusion of the epidural infusion (Fig. 2). Blood pressures were similar in the clonidine groups (C and LC) and were lower than in Group L. The difference between the mean systolic pressures amounted to approximately 10 mm Hg, and although statistically significant (P 0.05), this was not thought to be of major clinical importance (Fig. 3). There was one postoperative death, in a patient from Group LC. This patient, who had no significant cardiac risk factors, had had an uneventful hospital stay but collapsed and died at home 9 days after discharge. The cause of death was not established but was thought to be cardiac in origin and unrelated to

396

REGIONAL ANESTHESIA AND PAIN MEDICINE MILLIGAN ET AL. EPIDURALS, LEVOBUPIVACAINE, AND CLONIDINE

ANESTH ANALG 2000;91:3937

Figure 1. Postoperative sensory block (median, interquartile range). L levobupivacaine, C clonidine.

his anesthesia. There were no differences between the groups in terms of other adverse events, such as hypotension and cardiac arrhythmias, but there was a higher incidence of nausea in Group L (14 patients) compared with 6 patients in Group LC and 4 patients in Group C.

Figure 2. Patients with minimal postoperative motor block (Bromage score 0 or 1). L levobupivacaine, C clonidine.

Discussion
Bupivacaine possesses an asymmetric carbon atom and can therefore take the form of two enantiomers, levoand dex-bupivacaine. These have identical physical properties, but their chemical groups occupy different positions and therefore form different three-dimensional relationships in the asymmetric environment of receptors and enzymes. This can result in differences in both receptor affinity and intrinsic activity of the enantiomers (6), leading to differences in their toxicities, distribution, protein binding, metabolism, and elimination (7,8). Clonidine, a partial 2-adrenergic agonist, has a variety of different actions, including antihypertensive properties and the ability to potentiate the effects of local anesthetics. This has been demonstrated in a variety of clinical settings (4,9) and has been shown to result in the prolongation of the sensory blockade and a reduction in the amount or the concentration of local anesthetic required to produce postoperative analgesia. Previous work with epidural infusions has shown that 150 g of clonidine, when added to bupivacaine 0.25% approximately doubled the duration of the analgesia produced (4). In a study of epidural infusions of levobupivacaine in patients undergoing hip or knee surgery, significantly longer analgesia was achieved with levobupivacaine 0.25% than 0.125% or 0.0625% (10). The incidence of motor block was similar in the 0.125% and the 0.25% groups, and the latter provided the most effective pain relief as assessed by using VAS. The results of the present study demonstrate that epidural infusions of levobupivacaine are potentiated

Figure 3. Mean sd postoperative systolic blood pressure (mm Hg). L levobupivacaine, C clonidine. *Difference between groups: P 0.05.

by the addition of clonidine. The increase in efficacy of the combination of clonidine and levobupivacaine compared with levobupivacaine alone was demonstrated by the increase in the time to first request for analgesia (from 5 to 13 h) and an accompanying decrease in the total morphine consumption. VAS scores were also lower in the combination group, but all three groups appear to have had reasonable pain relief, as might be expected given that they all had free access to a PCA system. The clinical relevance of the small differences in pain scores is therefore open to question, but the improved analgesia may have contributed to the reduction in blood pressure in the clonidine groups. Another effect of the clonidine was that the motor blockade produced by the levobupivacaine was also increased, particularly in the early part of the study. This is a well recognized property of the drug (5) and was not a particular problem, as the patients were not

ANESTH ANALG 2000;91:3937

REGIONAL ANESTHESIA AND PAIN MEDICINE MILLIGAN ET AL. EPIDURALS, LEVOBUPIVACAINE, AND CLONIDINE

397

required to mobilize. It might, however, be a drawback in other settings. The sensory blockade was also slower to regress in the combination group, but this was again not considered to be a major problem. Postoperative nausea was less common in the clonidine groups postoperatively, and this probably reflects the lower consumption of morphine by these patients. In conclusion, although clonidine lowered blood pressure and increased the degree of motor and sensory block, the combination of clonidine and levobupivacaine, administered by the epidural route, was well tolerated and produced significantly improved postoperative pain management compared with either drug used alone.

References
1. Albright GA. Cardiac arrest following regional anesthesia with etidocaine and bupivacaine. Anesthesiology 1979;51:2857. 2. Huang YF, Pryor ME, Mather LE, Veering BT. Cardiovascular and central nervous system effects of intravenous levobupivacaine and bupivacaine in sheep. Anesth Analg 1998;86:797 804.

3. Bardsley H, Gristwood R, Watson N, Nimmo W. The local anaesthetic activity of levobupivacaine does not differ from racemic bupivacaine (Marcain): first clinical evidence. Exp Opin Invest Drugs 1997;6:18835. 4. Carabine UA, Milligan KR, Moore J. Extradural clonidine and bupivacaine for postoperative analgesia. Br J Anaesth 1992;68: 1325. 5. Carabine UA, Milligan KR, Mulholland D, Moore J. Extradural clonidine infusions for analgesia after total hip replacement. Br J Anaesth 1992;68:338 43. 6. Burm AGL, Van Der Meer AD, Van Kleef JW, et al. Pharmacokinetics of the enantiomers of bupivacaine following intravenous administration of the racemate. Br J Clin Pharmacol 1994; 38:1259. 7. Tucker GT, Lennard MS. Enantiomer specific pharmacokinetics. Pharmacol Therap 1990;45:309 29. 8. Vanhoutte F, Vereecke J, Verbeke N, Carmeliet E. Stereoselective effects of the enantiomers of bupivacaine on the electrophysiological properties of the guinea-pig papillary muscle. Br J Pharmacol 1991;103:1275 81. 9. Racle JP, Benkhadra A, Poy JY, Gleizal B. Prolongation of isobaric bupivacaine spinal anesthesia with epinephrine and clonidine for hip surgery in the elderly. Anesth Analg 1987;66: 442 6. 10. Murdoch J, Dickson U, Wilson P, et al. Levobupivacaine administered as a continuous epidural infusion for postoperative pain in patients undergoing elective surgery. Internat Mon Reg Anaesth 1998;10:9.

Regional Anesthesia for Knee Surgery


Learning Objectives: 1. Describe the indications and contraindications for the various regional anesthesia and analgesia techniques for surgery on the knee 2. Describe the relevant neuroanatomy of the lumbar plexus, lumbosacral plexus and innervation to the knee 3. Describe the techniques, local anesthetics and adjuvants used for performing regional anesthesia and analgesia for surgery on the knee 4. Compare and contrast the benefits of regional anesthesia and analgesia versus general anesthesia for both the intra- and postoperative period. Include both outpatient and inpatient procedures.

Regional Anesthesia for Knee Surgery


Peter G. Atanassoff, MD, and Maximilian WE3 Hartmannsgruber, MD

Knee surgery may be performed under general anesthesia, preferably, however, under regional anesthesia because of the profound preemptive analgesic effect that is provided by regional anesthetic techniques. Both centroneuraxis (spinal, epidural) as well as peripheral nerve blocks may be used for knee surgery. Although the former may be used less frequently in the near future in knee surgery because of increasing administration of low molecular heparins with resultant epidural hematoma formation, the latter enjoy more and more popularity. Peripheral nerve blocks used mostly for knee surgery include femoral/sciatic nerve blocks and intra-articular injection of local anesthetics and/or opioids into the knee joint. They are devoid of extensive sympathectomy and provide sufficient surgical analgesia and motor block. Copyright 0 1999 by W.B. Saunders Company

Spinal Anesthesia
Spinal anesthesia provides a temporary interruption of nerve transmission achieved by the injection of a relatively small amount of a local anesthetic solution into the subarachnoid space. This space is separated from the epidural space by the dura and contains cerebrospinal fluid and nerve fibers. In most adults, the spinal cord ends at Ll-L2, therefore spinal anesthesia can be safely administered between L2 and L5. Generally, spinal anesthesia is achieved via single injection of a local anesthetic agent. Continuous spinal anesthesia can be achieved with the use of extremely thin intrathecal catheters (28 and 32 gauge IG]) that were briefly commercially available in the United States. This technique has been widely abandoned in the United States because of fear of cauda equina syndrome (paresthesia, motor weakness, and paralysis of the lower extremities, bladder and bowel dysfunction) that were associated with the microcatheter technique.E However, large cumulative doses of local anesthetic agents, not the microcatheter technique, may have been responsible for the development of neurotoxicityg Microcatheters continue to be used in Europe with great success. The most common agent used is 0.5% bupivacaine administered in doses of 5 to 7.5 mg.rO Continuous spinal anesthesia is still performed here with macrocatheters (18 to 20 G) placed through an epidural needle. This technique is particularly attractive if an epidural needle is inserted accidentally into the subarachnoid space. Rather than remove the needle and reinserte it at a different interspace, the as.&w~ xecwxinvd. ivsdkw2l i<$&Q\ of tke XX& ant%thetic through the epidural needle and/or insertion of an intrathecal catheter. Surgical anesthesia may then be managed with repeat injections of 0.5% bupivacaine in 0.5 mL increments. Spinal anesthesia compared with epidural anesthesia offers the advantage of a dense sensory and motor block with a short onset time. Administration of a hypobaric local anesthetic such as plain bupivacaine 0.5% enables the patient to perceive a temperature increase in the lower extremities about 30 to 60 seconds after injection, hinting towards correct subarachnoid needle placement. If a unilateral block is preferred, a hyperbaric local anesthetic solution should be used. Complex procedures on the knee frequently require the use of an upper thigh tourniquet. Tourniquet pain develops despite adequate sensory or motor block at various times, depending on the local anesthetic used. Conception reported a statistically significant difference in the incidence of tourniquet pain when bupivacaine (15 mg) was compared to tetracaine (15 mg) in patients undergoing lower extremity procedures using thigh tourniquets. ii Patients given bupivacaine reported tourniquet pain 25% of the time compared to 60% for patients given tetracaine despite longer tourniquet times in the bupivaCaine group. An explanation for the prolonged tourniquet

uccessful anesthesia for orthopedic procedures on the knee can be achieved by a variety of regional techniques. These not only offer an alternative or supplement to general anesthesia (GA), they add the following distinct advantages: Preemptive analgesic effects, ie, reduction of pain beyond the pharmacologic action of the local anesthetic agent,im3 reduction of the stress response.4 2. Catheters inserted during these blocks can be used for postoperative pain relief. 3. Allowance of earlier mobilization,5 reduction of deep venous thrombus formation, and pulmonary embolism.6 4. Reduced PACU time and cost.7
1.

The three most common interventions on the knee are arthroscopy, arthroplasty, and repair of the anterior cruciate ligament (ACL). In general, surgical time for these procedures only rarely exceeds the pharmacologic duration of longer lasting local anesthetic agents. However, open knee procedures are extremely painful; therefore, continuous catheters for continuous analgesia are frequently inserted. These catheters enable patients to tolerate mobilization immediately after surgery, This provides faster, less painful rehabilitation as well as superior functional reXlltS.5

This report describes our rationale for selecting one regional anesthetic technique over the other and what we believe are key points in order to achieve successful regional anesthesia [or surgical procedures on the knee.

From the Department of Anesthesiology Yale University School of Medicine New Haven, CT. Address reprint requests to Peter G. Atanassoff, MD, Department of Anesthesiology, Yale University School of Medicine, 333 Cedar Street, PO B o x 2 0 8 0 5 1 , N e w H a v e n , C T 06520-8051. Copyright 0 1999 by W.B. Saunders Company 1084-208x/99/0302/0006$1 0.00/O

Techniques in Regional Anesthesia and Pain Management, Vol 3, No 2 (April), 1999: pp 107-l 12

107

tolerance with bupivacaine is that tetracaine offers less prolonged blockade of C fibers. l2 Other maneuvers shown to improve tourniquet tolerance time are the addition of an opioid, the addition of epinephrine, larger volumes of local anesthetic, and avoidance of glucose in the local anesthetic solution.13 When glucose was added to bupivacaine, the incidence of tourniquet pain at similar sensory levels was 37% versus 13% when the same volume of plain isobaric bupivaCaine was used. The addition of an intrathecal opioid extends and outlasts the effect of even long-acting local anesthetics. The addition of morphine 0.25 to 0.3 mg to the local anesthetic extends the duration of the anesthesia and analgesia in a dose dependent fashion.14

Epidural Anesthesia
Epidural anesthesia provides both surgical anesthesia and effective postoperative pain relief and is the most extensively used regional anesthesia technique for procedures on the lower extremities. When employed for knee replacements, this technique has been shown to reduce perioperative mortality due to pulmonary embolism.6 This reduction in postoperative mortality from pulmonary embolism is consistent with reports of decreased incidence of deep venous thrombosis in patients receiving epidural anesthesia. Sonographically detected lower extremity deep venous thrombosis (DVT) is more frequent in patients who had received general rather than epidural anesthesia.* Both spinal and epidural techniques improve blood flow to the lower limbs by inducing arterial and venodilation. Venous stasis is thereby reduced.16 Epidural anesthesia may have the added benefit of generating high local anesthetic plasma concentrations which have been shown to decrease platelet aggregation. Earlier ambulation and discharge times may also play an important role in preventing deep venous thrombosis formation. Effective continuous epidural analgesia with local anesthetics with or without opioids allows early postoperative ambulation, use of continuous passive motion (CPM) machines, and active physical rehabilitation immediately following surgery. In contrast, even high dose epidural fentanyl (5 pg/kg) was unable to sufficiently relieve pain for passive mobilization following knee surgery18 Of note is that effective epidural analgesia with low concentrated local anesthetic solutions such as lidocaine 0.25% does not interfere with normal muscle function; muscle strength improved most likely due to the reduction in pain, and thereby prevented some of the postoperative muscle atrophy19 To access the epidural space the patient is positioned in the sitting or lateral decubitus position. Although the effect of gravity is controversial, reliability of blockade of Sl, which is necessary for knee procedures, is probably increased with the patient in the sitting position. Alternatively, patients who have a history of fainting or who are heavily premeditated should have their catheter inserted while in the lateral position. To increase involvement of the Sl segment, patients should then be placed in a modified reverse Trendelenburg position (Hammock position) prior to injection of the local anesthetic. When patients were randomly allocated to one of two groups with supine horizontal position versus supine 30 trunk elevation with 30 leg elevation (hammock position), the latter position markedly improved the percentage of patients (60% v 13%) having adequate anesthesia for knee surgery20

Higher volumes of local anesthetic solutions compared to spinal anesthesia are necessary to ensure sensory and motor blockade of the knee. A combination of lidocaine 2% with epinephrine (5 pg/mL) and ropivacaine 0.75% or l%, or lidocaine 2% with epinephrine (5 pg/mL) and bupivacaine 0.5% or 0.75% in a 1:l ratio provide both rapid onset and long duration of blockade. Postoperative pain following major knee surgery can be effectively managed with patient controlled epidural analgesia (PCEA). The patient receives a basic continuous infusion and may additionally self-administer bolus doses. Dilute solutions of local anesthetic with the addition of an opioid such as morphine, fentanyl, or hydromorphone have been shown to provide effective pain relief in the absence of motor blockade.21 Very dilute bupivacaine (1132%) combined with hydromorphone (10 pg/mL) at a rate of 10 to 12 mUhour is used for pain relief in the authors institution and enables the patients to undergo early postoperative rehabilitation.

Combined SpinaVEpidural

Anesthesia

Combined spinal-epidural anesthesia (CSE) offers distinct advantages for surgery on the knee. First, it provides the rapid onset and reliable sensory and motor blockade of spinal anesthesia. Second, there is the option to improve an inadequate spinal blockade or extend the dermatomal level. Third, it permits repeat injections through an indwelling epidural catheter, and fourth the needle-through-needle technique confirms proper positioning of the Tuohy needle and therefore may improve the rate of successful catheter placement. The technique usually involves placing an epidural needle into the epidural space first and then advancing a long 27-G spinal needle through the Tuohy needle into the subarachnoid space. Placing the Tuohy needle in the midline rather than paraspinally is preferred by the investigators because it seems TABLE 1. Data on 15 Consecutive Cases of Epiduroscopy
Migration of Catheter Single DUral Puncture With Spinal Spinal Level Needle
L3-L4 L-L3 -

Case Age No. (Yr) Sex 1 2 3 4 5 6 81 79 87 79 69 72 F F F F F M

Multiple Dural Punctures With Spinal Needles


-

Single Dural Puncture With Tuohy Needle


-

+
+ + + + -

Failure
L3-L4 u - 3 L3-L4 b-L3 L3-L4 L1-Lz L-L3 L3-L4

7 8

96 62

F M

w-3 L3-L4 w - 3 L1-LP

9 10 11 12 13 14 15

92 70 39 79 68 89 94

F M F F M F F

Failure Failure
L3-L4 b-L3

+ -

Failure
L3-L4 LA-2

La-L4 LYL3

1 -

Reproduced with permission from Holmstrom Bet al.* ATANASSOFF AND HARTMANNSGRUBER

108

to lead to a higher success rate of spinal needle placement. Provided the spinal needle is correctly positioned, a local anesthetic with or without the addition of an opioid is injected intrathecally. Following removal of the spinal needle the epidural catheter is inserted. In the investigators institution, patients frequently receive intrathecal hypobaric bupivacaine 0.5% 1 to 3 mL with epinephrine (5 pg/mL) and epidural hydromorphone 0.5 to 1 mg. This combination provides long lasting intraoperative surgical anesthesia that extends into the postoperative period. Further postoperative pain relief is then achieved by the previously mentioned combination of dilute bupivacaine and hydromorphone. Theoretically, there is the concern of accidental passage of the epidural catheter through the kale of the &~a( ~WXVX~ site.22 This possibility is very small as demonstrated in cadavers by epiduroscopy where even after multiple dural punctures with 25 to 26 G spinal needles no migration of epidural catheters was seen (Table 1).

Femoral/Sciatic

Nerve

Block

Many anesthesiologists prefer spinal and/or epidural techniques over peripheral nerve blocks for lower extremity surgery 23 Because of the risk of permanent neurologic damage from epidural hematoma formation, neuraxial techniques are contraindicated in anticoagulated patients. Low molecular weight heparin, in particular, has elicited discussions on safety of centroneuraxis blockade following publication of a series of

neuraxial hematomas, which resulted in long-term or permanent paralysis. Several incidences occurred when the first dose of low molecular weight heparin was administered within 12 hours after surgery Epidural hematomas were also described following epidural catheter removal within 10 to 12 hours after dosing low molecular weight heparin, or when subsequent dosing was not delayed at least 2 hours after epidural catheter removal (See Horlocker, Regional Anesthesia and Analgesia in the Orthopedic Patient Receiving Thromboprophylaxis).24 Because of these concerns anesthesiologists may shift away from neuraxial towards peripheral nerve blocks. Further advantages of peripheral nerve blocks are that hemodynamic changes as seen with spinal and epidural anesthesia are less likely to occur, and normal bladder and bowel function are preserved. A theoretical disadvantage of lower extremity peripheral nerve blockade is larger volumes of local anesthetic that are required. However, the risk of systemic toxicity is not proportionately increased because of the decreased uptake of local anesthetics from peripheral sites.25 Four nerves innervate the knee, ie, the femoral, the lateral femoral cutaneous, the obturator, and the sciatic nerves. The first three arise from Ll-L4, the sciatic nerve originates from LS-S2. The femoral nerve is blocked inferior to the inguinal ligament and immediately lateral to the femoral artery. The lateral femoral cutaneous nerve is anesthetized as it emerges from the fascia lata inferomedial to the anterior superior iliac spine. The obturator nerve is blocked as it emerges from the obturator canal (Fig 1). An alternative to these individual

Fig 1. Obturator nerve block technique. Reprinted with permission from Brown: Atlas of Regional Anesthesia, Philadelphia, PA, Saunders, 1992.

REGIONAL ANESTHESIA FOR KNEE SURGERY

109

Fig 2. Laying across the forceps, the femoral (middle) and lateral femoral cutaneous (right) nerves are dyed blue following femoral intraneural injection of 40 mL of methylene blue dye. The obturator nerve (left) is not dyed. Reprinted from J Clin Anesth, 7, Ritter JW, Femoral nerve sheath for inguinal paravascular lumbar plexus block is not found in human cadavers, 470-473, 1995, with permission of Elsevier Science.

Fig 3. Sciatic nerve block: classic technique and positioning. Reprinted with permission from Brown: Atlas of Regional Anesthesia, Philadelphia, PA, Saunders, 1992.

110

ATANASSOFF AND HARTMANNSGRUBER

3-in-1 blocks added to a general anesthetic compared to those who received general anesthesia only.31,32

Intrarticular Anesthesia
oscopy and ACL repair are frequently performed on an outpatient basis. Knee arthroscopy is performed in many centers using intraarticular local anesthetics with monitored anesthesia care. Volumes of 50 to 60 mL of a local anesthetic solution (ie, bupivacaine 0.25%) are required to ensure sufficient surgical anesthesia. Even with those high volumes, plasma levels remain 10 to 15 times below a nontoxic plasma level. This is presumably due to slow absorption through the synovia and considerable washout of the local anesthetic after the arthroscopy 33 Furthermore, peak serum bupivacaine concentrations can be reduced by adding epinephrine and injectA I ing the local anesthetic solution after tourniquet inflation.34 questionnaire sent to patients who had undergone arthroscopy as an outpatient procedure showed that the degree of satisfaction after local or spinal anesthesia was the same.33 A retrospective review examined a series of knee arthroscopic procedures that were completed using local, general, or regional anesthesia to evaluate the efficacy of these anesthetic techniques. Surgical time, complications or failures, procedures successfully performed, recovery room time, postoperative stay, and patient satisfaction were recorded. Local anesthesia with intravenous sedation compared favorably with the other techniques: surgical time was not increased, a large variety of operative procedures were successfully completed, recovery time was significantly shortened, and patient satisfaction remained high. It was concluded that this technique may offer several advantages over other types of anesthesia for knee arthroscopy, including improved cost effectiveness.35 In the presence of inflammation, peripheral opioid receptors may become accessible. Systemically ineffective doses administered into the knee joint after surgery have been shown to elicit potent and long lasting postoperative analgesia.36 The low lipid solubility of morphine and therefore its slow uptake into the circulation were postulated to account for the high degree of analgesia. In contrast, two other studies could not confirm the existence of peripheral opioid receptors in the knee joint after arthroscopy, described, however, the analgesic advantages of intraarticular local anesthetics.37 In conclusion there are several viable options of regional anesthetic techniques for knee surgery They all have in common substantial preemptive analgesic effects and may contribute to an overall reduction of patient morbidity

SITE

Fig 4. Sciatic nerve block: anterior technique. Reprinted with permission from Brown: Atlas of Regional Anesthesia, Philadelphia, PA, Saunders, 1992. blocks is the 3-in-1 block which theoretically blocks the three nerves by a single injection, provided sufficient volume of local anesthetic is injected. 26 For knee arthroscopy the 3-in-1 block provides a greater degree of muscle relaxation and a longer postoperative analgesia than the femoral nerve block alone.27 Electromyographic and anatomical studies, however, have shown that the obturator nerve may not be blocked even when using up to 50 mL of local anesthetic solution2sa29 (Fig 2). This may not be an important factor in ACL repair or knee arthroscopy, but as the obturator nerve provides sensory innervation to the medial aspect of the thigh and knee, separate blockade of this nerve may be warranted for knee arthroplasty To provide complete anesthesia for the knee, a sciatic nerve block must accompany the 3-in-1 block. The sciatic nerve can be located by a variety of approaches. The classic posterior approach described by Labat (Fig 3), the anterior approach (Fig 4), or the lithotomy approach. Combined 3-in-l/sciatic nerve blocks for knee arthroscopy provides excellent intraoperative and postoperative analgesia and reduces postoperative complication rates.25,30 For postoperative pain management following open-knee procedures, the 3-in-1 block is an excellent adjunct to general anesthesia. Two studies have documented lower pain scores with less narcotic usage in patients who received continuous

References
1. Woolf CJ, Chong MS: Preemptive analgesia-treating postoperative pain by preventing the establishment of central sensitization. Anesth A n a l g 77:362-379, 1993 2. Kissin I: Preemptive analgesia: terminology and clinical relevance. Anesth Analg 79:809-810, 1994 3. Atanassoff PG, Jarrett JM: The physiological and pharmacological aspect of pre-emptive analgesia. Anaesth Pharm Physiol Rev 4:6-20, 1996 4. Gold MS, DeCrosta D, Rizzuto C, et al: The effect of lumbar epidural and general anesthesia on plasma catecholamines and hemodynamits during abdominal aortic aneurysm repair. Anesth Analg 78:225230,1994

REGIONAL ANESTHESIA FOR KNEE SURGERY

111

5. Ecker ML, Lotke PA: Postoperative care of the total knee patient. Ortho Clin North America 205562, 1989 6. Sharrock NE, Cazan MG, Hargett MJL, et al: Changes in mortality after total hip and knee arthroplasty over a ten-year period. Anesth A n a l g 80:242-248, 1995 7. Lintner S, Shawen S, Lohnes J, et al: Local anesthesia in outpatient knee arthroscopy: a comparison of efficacy and cost. Arthroscopy 12:482-488, 1996 8. Pitkaenen M: Continuous spinal anaesthesia. Curr Opin Anaesthesiol 5:676-680, 1992 9. Rigler ML, Drasner K, Krejcie TC, et al: Cauda equina syndrome after continuous spinal anesthesia. Anesth Analg 72:275-281, 1991 10. Lambert LA, Lamber DH, Strichartz GR: Irreversible conduction block in isolated nerve by high concentrations of local anesthetics. Anesthesiology 80:1082-l 093, 1994 11. Concepion MA, Lamben DH, Welch KA, et al: Tourniquet pain during spinal anesthesia: A comparison of plain solutions of tetracaine and b u p i v a c a i n e . A n e s t h A n a l g 67:828-832, 1988 12. Stewart A, Lamben DH, Concepion MA, et al: Decreased incidence of tourniquet pain during spinal anesthesia with bupivacaine: a possible e x p l a n a t i o n . A n e s t h A n a l g 67:833-837, 1988 1 3 . Halaszynski TM, Hartmannsgruber MWB: Anatomy and physiology of a spinal and epidural anesthesia. Sem Anesth Periop Med 17:24-37, 1998 14. Bailey PL, Rhondean S, Schafer PG, et al: Dose-response pharmacology of intrathecal morphine in human volunteers. Anesthesiology 79149-59, 1993 15. Davidson HC, Mazzu D, Gage BF, et al: Screening for deep venous thrombosis in asymptomatic postoperative orthopedic patients using color Doppler sonography: analysis of prevalence and risk factors. Am J Roentgen 166:659-662, 1996 16. Modig J, Borg T, Karlstrom G, et al: Thromboembolism after total hip replacement: Role of epidural and general anesthesia. Anesth Analg 62:174-l 80, 1983 17. Borg T, Modig J: Potential anti-thrombotic effects of local anesthetics d u e t o t h e i r i n h i b i t i o n o f p l a t e l e t a g g r e g a t i o n . Acta A n a e s t h e s i o l Stand 29:739-742, 1985 18. Pierrot M, Blaise M, Dupuy A, et al: Peridural analgesia with high doses of fentanyl: Failure of the method for early postoperative kinesitherapy in knee surgery, Can Anaesth Sot J 29587-592, 1982 19. Arvidsson I, Eriksson E, Knutsson E, et al: Reduction of pain inhibition on voluntary muscle activation by epidural analgesia. Orthopedics 9:1415-1419, 1986 t 20. Ponhold H, Kulier AH, Rehak PH: 30 degree trunk elevation of the patient and quality of lumbar epidural anesthesia. Effects of elevation in operations on the lower extremities. Anaesthesist 42:788-792, 1993 21. Sinatra RS: Acute pain management and acute pain services, in Cousins MJ, Bridenbaugh PO (eds): Neural Blockade. Philadelphia, PA, Lippincott-Raven, 1998, pp 793-835

22. Holmstrom B, Rawal N, Axelsson K, et al: Risk of catheter migration during combined spinal epidural block: Percutaneous epiduroscopy s t u d y . A n e s t h A n a l g 80:747-753, 1995 23. Hadzic A, Vloka JD, Kuroda MM, Koorn R, Birnbach DJ: The practice of peripheral nerve blocks in the United States: a national survey. Reg Anesth Pain Med 23:241-246, 1998 24. Horlocker TT, Heit JA: Low molecular weight heparin: biochemistry, pharmacology, perioperative prophylaxis regimens, and guidelines for regional anesthetic management. Anesth Analg 85:874-885, 1997 25. Elmas C, Atanassoff PG: Combined inguinal paravascular (3-in-l) and sciatic nerve blocks for lower limb surgery. Reg Anesth 2:88-92, 1993 26. Winnie AP, Ramamurthy S, Durrani Z: The inguinal paravascular technic of lumbar plexus anesthesia: The 3-in-1 block. Anesth Analg 52:989-996,1973 27. Bonicalzi V, Gallino M: Comparison of two regional anesthetic techniques for knee arthroscopy. Arthroscopy 11:207-212,1995 28. Atanassoff PG, Weiss BM, Brull SJ, et al: Electromyographic comparison of obturator nerve block to three-in-one block. Anesth Analg 81:529-533, 1995 29. Ritter JW: Femoral nerve sheath for inguinal paravascular lumbar plexus block is not found in human cadavers. J Clin Anesth 7:470-473, 1995 30. Edwards ND, Wright EM: Continuous low dose 3-in-1 nerve blockade for postoperative pain relief after total knee replacement. Anesth A n a l g 75:265-267, 1992 31. Serpell MG, Millar FA, Thomson MF: Comparison of lumbar plexus block versus conventional opioid analgesia after total knee replacem e n t . A n a e s t h e s i a 46:275-277, 1991 32. Eriksson E, Haggmark T, Saat-tok T, et al: Knee arthroscopy with local anesthesia in ambulatory patients. Methods, results and patient c o m p l i a n c e . O r t h o p e d i c s 9:186-l 8 8 , 1 9 8 6 33. Solanki DR, Enneking FK, lvey FM, et al: Serum bupivacaine concentrations after intraarticular injection for pain relief after knee arthroscopy. Arthroscopy 8:44-47,1992 34. Shapiro MS, Safran MR, Crockett H, et al: Local anesthesia for knee arthroscopy. Efficacy and cost benefits. Am J Sports Med 23:50-53, 1995 35. Raja SN, Dickstein RE, Johnson CA: Comparison of postoperative analgesic effects of intraarticular bupivacaine and morphine following arthroscopic knee surgery. Anesthesiology 77:1143-1147, 1992 36. Stein C, Comisel K, Haimerl E, et al: Analgesic effect of intraarticular morphine after arthroscopic knee surgery. N Engl J Med 325:11231126,199l 3 7 . Heard SO, Edwards WT, Ferrari D, et al: Analgesic effect of intraarticular bupivacaine or morphine after arthroscopic knee surgery: A randomized, prospective, double-blind study. Anesth Analg 74:822826,1992

112

ATANASSOFF AND HARTMANNSGRUBER

1655

Anesthesiology 1999; 91:1655 60 1999 American Society of Anesthesiologists, Inc. Lippincott Williams & Wilkins, Inc.

A New Anterior Approach to the Sciatic Nerve Block


Jacques E. Chelly, M.D., Ph.D.,* Laurent Delaunay, M.D.

Background: Although several anterior approaches to sciatic nerve block have been described, they are used infrequently. The authors describe a new anterior approach that allows access to the sciatic nerve with the patient in the supine position. Method: Sciatic nerve blocks were performed in 22 patients. A line was drawn between the inferior border of the anterosuperior iliac spine and the superior angle of the pubic symphysis tubercle. Next, a perpendicular line bisecting the initial line was drawn and extended 8 cm caudad. The needle was inserted perpendicularly to the skin, and the sciatic nerve was identied at a depth of 10.5 cm (9.513.5 cm; median and range) using a nerve stimulator and a 15-cm b-beveled insulated needle. After appropriate localization, either 30 ml mepivacaine, 1.5% (group 1 knee arthroscopy; n 16), or 15 ml mepivacaine, 1.5%, plus 15 ml ropivacaine, 0.75%, (group 2 other procedures; n 6) was injected. Results: Appropriate landmarks were determined within 1.3 min (0.52.0 min). The sciatic nerve was identied in all patients within 2.5 min (1.25 min), starting from the beginning of the appropriate landmark determination to the stimulation of its common peroneal nerve component in 13 cases and its tibial nerve component in 9 cases. A complete sensory block in the distribution of both the common peroneal nerve component and the tibial nerve component was obtained within 15 min (530 min). A shorter onset was observed in patients who received mepivacaine alone compared with those who received a mixture of mepivacaine plus ropivacaine (10 min [525 min]

vs. 20 min [10 30 min]; P < 0.05). Recovery time was 4.6 h (2.55.5 h) after mepivacaine administration. The addition of ropivacaine produced a block of a much longer duration 13.8 h (5.223.6 h); P < 0.05. No complications were observed. Conclusions: This approach represents an easy and reliable anterior technique for performing sciatic nerve blocks. (Key words: Peripheral nerve block; lower extremity surgery; mepivacaine; ropivacaine.)

Additional material related to this article can be found on the ANESTHESIOLOGY Web Site. Go to the following address, click on Enhancements Index, and then scroll down to nd the appropriate article and link. http://www.anesthesiology.org

* Professor, Director of Regional Anesthesia and the International Regional Anesthesia Research Center at the Medical School, Department of Anesthesiology, The University of Texas Medical School-Houston. Staff Anesthesiologist, Clinique Ge ne rale, Annecy, France. Received from the Department of Anesthesiology, The University of Texas Medical School-Houston, Houston, Texas. Submitted for publication March 24, 1999. Accepted for publication July 9, 1999. Supported by the Department of Anesthesiology, The University of Texas Medical School-Houston, Houston, Texas. Address reprint requests to Dr. Chelly: Department of Anesthesiology, The University of Texas Health Science Center at Houston, 6431 Fannin, MSB 5.020, Houston, Texas 77030-1503. Address electronic mail to: jchelly@anes1.med.uth.tmc.edu

ALTHOUGH the combination of sciatic nerve and 3-in-1 blocks is an alternative to general or neuroaxial blocks for patients undergoing surgery of the lower extremities,1,2 sciatic nerve blocks are performed infrequently. Recent surveys have indicated that sciatic nerve blocks are the least performed by anesthesiologists.35 Reasons for this situation include lack of adequate training and the claim that sciatic blocks are difcult to perform. Several approaches have been described that depend on the position of the patient. The sciatic nerve can be approached with the patient supine or in the Sims position. In the supine position, the sciatic nerve can be accessed laterally or anteriorly. The posterior approach was rst described by Labat6 in 1930 and improved by Winnie.7 Although the posterior approach is the most commonly performed, it necessitates patient repositioning, which limits its use in patients with compromised mobility caused by severe arthritis, obesity, or trauma. In 1959, Ichiyanahi8 described a lateral approach of the sciatic nerve block. Today, this approach is performed commonly in children.9 The anterior approach, as rst described by Beck10 in 1963, may be difcult to perform because the appropriate femoral anatomical landmark, e.g., the greater trochanter, is not easily identied in obese patients. In addition, its identication may be extremely painful in patients with lower extremity fractures. Therefore, its use also has been limited. More recently, Raj et al.11 described a lithotomy approach with different landmarks (the midpoint of a line drawn between the greater trochanter and the ischial tuberosity) that allows more reliable access to the sciatic nerve after anterior exion of the legs. In patients with limited mobility, the exion of the leg requirement also represents an important limitation. Therefore, the usefulness

Anesthesiology, V 91, No 6, Dec 1999

1656 J. E. CHELLY AND L. DELAUNAY

of this approach is limited in patients with severe hip and knee arthritis and trauma. To perform sciatic nerve blocks while the patient is in the supine position, we developed a new anterior approach that necessitates neither repositioning nor identication of the greater trochanter. We studied the feasibility of this new anterior approach in 22 patients. Because the patients enrolled in this study underwent either knee arthroscopy (short procedure) or lower extremity surgeries estimated to last about 2 h, we also studied the onset and duration of two local anesthetic solutions; i.e., 1.5% mepivacaine for knee arthroscopy and a mixture of 1.5% mepivacaine and 0.75% ropivacaine (volume/volume) for the longer procedure.
Fig. 1. Anatomic landmarks for the anterior approach to the sciatic nerve.

Methods
This prospective clinical study was approved by the Institutional Review Board of The University of Texas Medical School-Houston. After preoperative evaluation and discussion of anesthetic options, informed consent was obtained. Except for two patients, the blocks were performed preoperatively in the recovery room. In two patients, the sciatic nerve block was performed after the rst hour of recovery after total knee replacement and rodding of the tibia. In these two cases, patients also consented preoperatively. Twenty-two patients were included and separated into two groups according to the surgical indication: knee arthroscopy (group 1; n 16), and other procedures (group 2; n 6), including anterior crucial ligament reconstruction (1), tibial plateau reconstruction (1), rodding of the tibia (1), total hip replacement (1), total knee replacement (1), and partial menisectomy (1). The sciatic nerve blocks were performed immediately after a paravascular 3-in-1 block1 in 20 patients, and after a paravascular 3-in-1 block followed by the placement of a femoral catheter to be used postoperatively for continuous infusion of 0.2% ropivacaine at a rate of 1216 ml/h to provide postoperative analgesia in the patients scheduled for either total knee or hip replacement.12 After placement of the proper monitors (blood pressure, electrocardiography, pulse oximetry), a peripheral intravenous catheter was established for the infusion of Ringers lactate, and baseline vital signs were recorded. Oxygen was delivered via face mask. Patients were then sedated as follows: 50 100 g intravenous fentanyl combined with 1.0 4.0 mg intravenous midazolam or propofol 20 30 mg intravenously, or both.
Anesthesiology, V 91, No 6, Dec 1999

Anatomic Landmarks With the patient supine and the lower extremity in the neutral position, a line was drawn between the inferior border of the anterosuperior iliac spine and the superior angle of the pubic symphysis tubercle. From this anterosuperior iliac spinepubic symphysis line, a perpendicular dissector line was drawn in the middle and extended 8 cm caudad (g. 1) to dene the site of introduction of the needle. During sterile conditions, a 15-cm insulated b-beveled Stimuplex needle (B-Braun/ McGaw Medical, Bethlehem, PA) connected to a Stimuplex-Dig nerve stimulator (B-Braun/McGaw Medical) was introduced perpendicularly to the skin after subcutaneous local anesthesia with 1 ml lidocaine, 1%, (Astra USA, Inc., Westboro, MA). The needle on its way to the sciatic nerve might come in proximity to the femoral nerve13 (g. 2); therefore, the nerve stimulator was initially set-up to deliver a current of 1.0 mA. At a depth of 35 cm, movements of the patella were observed. Because patella movements might represent the motor response to the stimulation of the femoral nerve, the current was decreased to 0.6 mA. If patella movements ceased, the needle was introduced 2 cm deeper, with the current of the nerve stimulator increased to 5 mA. Within a depth of 9.513 cm, the sciatic nerve was identied via the motor response related to the stimulation of its common peroneal nerve component (dorsiexion or eversion of the foot) or its tibial nerve component (plantar exion and inversion of the foot and exion of the toes). The current was then decreased, and the needle orientation was optimized to obtain the same response with a current equal to or lower than 0.7 mA (moved slightly medially, laterally, or deeper according to the

1657 ANTERIOR SCIATIC BLOCK

Fig. 2. Computed tomography scan of the thigh 1 cm above the lesser trochanter showing the relation among the femoral artery (A), vein (V) and nerve (N), the sartorius muscle (M), and the 15-cm insulated needle introduced perpendicular to the skin using our proposed landmarks.23

The intensity of the sensory and motor block was assessed every 5 min up to 45 min after being performed and every hour after surgery until sensory function recovered in either the common peroneal or tibial territory to determine the recovery time. Preoperatively and postoperatively, a three-level scale was used to evaluate the intensity of the sensory and motor block (no block, partial and complete block). In each case, the surgery was conducted in the absence of general anesthesia. The sensory block was assessed by application of ice to the dorsal aspect of the foot (common peroneal nerve) and to the plantar aspect of the foot (tibial nerve). The sensory block was considered complete if the patient did not feel the cold. Patients were asked to perform a dorsi and plantar exion to assess the intensity of the motor block (normal motor function, partial block, and complete block). The motor block was considered complete when a motor block was observed in both the common peroneal and the tibial territories. The duration of the sensory and motor block was dened as the time between the performance of the block and the recovery of sensory and any motor function, respectively. An unpaired t test was used to compare the demographic data between groups and are presented in table 1. In addition, the MannWhitney rank-sum test was performed to compare between groups the times to onset and completeness of the sensory and motor blockades. was set at 0.05. Data are presented as the median (range).

need). To satisfy postoperative analgesic requirements, group 1 patients received 30 ml mepivacaine, 1.5%, (Astra USA, Westboro, MA), whereas group 2 patients who required longer postoperative pain control received 15 ml mepivacaine, 1.5%, plus 15 ml ropivacaine, 0.75%, (Astra USA, Westboro, MA). After injection of a 1-ml test dose to verify that the injection was not neuronal (acute pain elicited by the injection), the solution of local anesthetics was injected slowly after negative blood aspiration, aspirating after every 5 ml to conrm that the needle remained extravascular. If the patella movements were maintained at 0.6 mA, the insulated needle was reoriented slightly medially. If the needle ended its course on the femur, it was retracted to the level of the skin (failed attempt), after which the skin was pulled 1 or 2 cm medially and the needle reintroduced vertically. At completion of the local anesthetic injection, patients were monitored for another 45 min before transfer to the holding areas, where they waited until the time of surgery.
Anesthesiology, V 91, No 6, Dec 1999

Results
Table 1 shows patient demographics. Determination of the appropriate landmarks required 1.1 min (0.4 2.0 min). The sciatic nerve was identied in all patients within 2.9 min (1.2 6.1 min) after two (one to three) attempts, and the sciatic nerve was found at a depth of 10.5 cm (9.513 cm). In 13 patients, the common peroneal nerve was rst stimulated, whereas stimulation of the tibial nerve was elicited in 9 patients. The current before injection was decreased to 0.5 mA (0.5 0.7 mA). A complete sensory block developed faster in the common peroneal than in the tibial territory. After 5 min, a complete common peroneal sensory block was observed in 50% of patients and in all patients after 20 min. In contrast, 10 min was necessary for a complete sensory block in the tibial territory in 50% of the patients and in all patients within 30 min (P 0.05). In these conditions, the overall onset time for a complete sensory

1658 J. E. CHELLY AND L. DELAUNAY

Table 1. Patient Demographics


Group 1 Group 2 Overall

Age (yr) Weight (kg) Height (cm) Duration of Surgery (min) N


Data are median (range). * P 0.05.

53 (1969) 72 (5387) 170.5 (160185) 19 (1528) 16

50.5 (4076) 76 (6895) 175 (172185) 75* (65155) 6

53 (1976) 72 (5395) 172.5 (160185) 19 (15155) 22

block in both the common peroneal and the tibial territories was 12.5 min (530 min). However, the onset time varied according to the anesthetic solutions. It was 10 min (525 min) after administration of mepivacaine alone and 20 min (10 30 min) after administration of the combination of mepivacaine and ropivacaine (P 0.05), indicating that with mepivacaine alone the onset was shorter. The overall duration of the block was 5.0 h (2.523.6 h). The block lasted 4.6 (2.55.5 h) with mepivacaine alone compared to 13.8 h (5.223.6 h) after the mixture of mepivacaine and ropivacaine (P 0.05). A complete common peroneal and tibial motor block was observed in 50% of patients after 20 min. The common peroneal and tibial motor blocks were completed in all patients after 45 min. In these conditions, the overall onset time for a complete motor block in both the common peroneal and the tibial territories was 25 min (5 45 min). The onset time did not vary with the anesthetic solutions. The overall duration of the motor block was 4 h (2.322.6 h). The motor block lasted 3.8 h (2.35.5 h) with mepivacaine alone compared with 11.4 h (4.4 22.6 h) after the mixture of mepivacaine and ropivacaine (P 0.05). Also, there was no hemodynamic compromise. No blood aspiration, paresthesia, or sensory and other motor decits were observed during and after the performance of these sciatic nerve blocks.

Discussion
These data show that this new anterior approach is effective for blocking the sciatic nerve. By using the proposed anatomic landmarks and a nerve stimulator, the sciatic nerve was identied in all patients, and a sensory and motor block was obtained in all patients. Using the anterior approach as described by Beck,10 Manani et al.14 reported a 14% rate of failure. However, it is important to recognize that we used a nerve stimulator, whereas Manani et al.14 used a paresthesia techAnesthesiology, V 91, No 6, Dec 1999

nique. Davies and McGlade15 reported that for a posterior approach to the sciatic nerve the use of paresthesia resulted in less than 50% positive identication of the sciatic nerve versus 92% with a nerve stimulator. Therefore, the identication of the sciatic nerve is greatly facilitated by the use of a nerve stimulator. The combination of sciatic and 3-in-1 blocks is particularly appropriate for lower extremity surgery necessitating the use of a tourniquet.12 Furthermore, the use of a single-injection sciatic block in combination with a continuous femoral nerve block or a continuous lumbar plexus block is especially useful for total knee replacements.12 In this regard, Allen et al.16 recently reported that, in a group of 15 patients, the block of the sciatic nerve was not necessary for postoperative pain control in patients who underwent total knee replacement. In our group of patients, a sciatic nerve block was performed after total knee replacement because of pain in the posterior aspect of the knee. The effectiveness of the sciatic nerve block in controlling pain17 suggests that, in contrast to the concept developed by Allen et al.,16 blockade of the sciatic nerve may be important in patients undergoing total knee replacement. Therefore, it appears that additional data are necessary to determine the role of sciatic nerve blocks in postoperative pain control after total knee replacement. It is established that the time of onset for sciatic nerve block using a posterior approach is approximately 20 30 min.18 This is most likely related to the large size of the nerve, necessitating more time for diffusion of the local anesthetic solution. Our data indicate that a similar time requirement exists with our anterior approach. However, the onset and duration of the blocks also depend on the local anesthetic mixture. Thus, the onset and duration of the blocks were shorter with mepivacaine compared with the onset and duration when ropivacaine was added. These data conrmed the data already published regarding the duration of sciatic nerve blocks with mepivacaine19 and ropivacaine.20

1659 ANTERIOR SCIATIC BLOCK

Fig. 3. Comparisons of our anatomic landmarks and those described by Beck.10 ASIS anterosuperior ischemic spine; GT greater trochanter; LT lesser trochanter; PT pubic tuberosity; PS pubic symphysis; LB lower border.

ing femoral landmarks and is based on the use of pelvic landmarks that are easily identied, even in obese and trauma patients; i.e., the lower border of the anterior superior iliac spine and the superior angle of the pubic symphysis tubercle. It is especially interesting that the anterior approach originally described by Beck10 and our description, using different pelvic landmarks, identify the same site for introduction of the needle (g. 3). In many patients, two attempts were necessary for the identication of the sciatic nerve. With the rst attempt, the needle ended its course on the femur. Although this rst attempt was counted as a failed attempt, it is important to recognize that it allowed an estimate of the depth at which the sciatic nerve could be found during the next attempt. Therefore, the distance between the skin and the sciatic nerve equals the distance between the skin and the femur plus the estimated thickness of the femur (3 or 4 cm). Furthermore, at the level of the lesser trochanter, the position of the sciatic nerve is known to be medial and posterior to the femur. Hadzic et al.22 demonstrated that if the needle is above the lesser trochanter, an internal rotation facilitates the location of the sciatic nerve, whereas an external rotation facilitates an approach of the sciatic nerve below the lesser trochanter. Although this rotation technique was not necessary in this group of patients, it has been used successfully in other patients, especially the internal rotation, suggesting that with our proposed landmarks the sciatic nerve is approached from above the lesser trochanter.

Conclusion
The short onset time combined with a block of several hours duration, favors the performance of these blocks at some interval from surgery. This is especially interesting when considering that Pavlin et al.21 showed that the use of peripheral nerve block as the sole anesthesia technique reduces the duration of hospital stays for outpatient procedures by approximately 1 h. Even a delay between the performance of the block and the beginning of surgery is unlikely to affect the effectiveness of these blocks. In 1963, Beck10 rst described anatomic landmarks that allowed localization of the sciatic nerve via an anterior approach based on the identication of the greater trochanter. As previously indicated, the greater trochanter is not always easy to identify in obese patients, and attempts to locate it may be very painful for trauma patients. Our approach does not necessitate usAnesthesiology, V 91, No 6, Dec 1999

This easy and reliable anterior technique for performing sciatic nerve blocks is an alternative to the more traditional approaches, especially in patients with limited mobility. This approach should therefore facilitate the performance of sciatic nerve blocks.

References
1. Winnie AP, Ramamurthy S, Durrani Z: The inguinal paravascular technique of lumbar plexus anesthesia: The 3-in-1 block. Anesth Analg 1973; 52:989 96 2. Jankowski CJ, Horlocker TT, Rock MJ, Stuart MJ: Femoral 3-in-1 nerve block decreases recovery room time and charges and time to hospital discharge after outpatient knee arthroscopy (abstract). Reg Anesth Pain Med 1998; 60:23S 3. Hadzic A, Vloka JD, Kuroda MM, Koorn R, Birnbach DJ: The practice of peripheral nerve blocks in the United States: A national survey. Reg Anesth Pain Med 1998; 23:241 6

1660 J. E. CHELLY AND L. DELAUNAY

4. Bouaziz H, Mercier FJ, Narchi P, Poupard M, Auroy Y, Benhamou D: Survey of regional anesthetic practice among French residents at time of certication. Reg Anesth 1997; 22:218 22 5. Smith MP, Sprung J, Zura A, Mascha E, Tetzlaff JE: A survey of exposure to regional anesthesia techniques in American anesthesia residency training programs. Reg Anesth Pain Med 1999; 24:11 6 6. Labat G: Its technique and clinical applications, Regional Anaesthesia, 2nd edition. Philadelphia, Saunders Publishers, 1924, pp 4555 7. Winnie AP: Regional anaesthesia. Surg Clin North Am 1975; 54:86192 8. Ichiyanagi K: Sciatic nerve block: Lateral approach with patient supine. ANESTHESIOLOGY 1959; 20:601 4 9. Dalens B, Tanguy A, Vanneuville G: Sciatic nerve blocks in children: Comparison of the posterior, anterior, and lateral approaches in 180 pediatric patients. Anesth Analg 1990; 70:1317 10. Beck GP: Anterior approach to sciatic nerve block. ANESTHESIOLOGY 1963; 24:222 4 11. Raj PP, Parks RI, Watson TD, Jenkins MT: A new single-position supine approach to sciatic-femoral nerve block. Anesth Analg 1975; 54:489 93 12. Singelyn FJ, Deyaert M, Joris D, Pendeville E, Gouverneur JM: Effects of intravenous patient-controlled analgesia with morphine, continuous epidural analgesia, and continuous three-in-one block on postoperative pain and knee rehabilitation after unilateral total knee arthroplasty. Anesth Analg 1998; 87:88 92 13. Bo WJ, Meschan I, Krueger WA: Male pelvis and perineum, Atlas of Basic cross-sectional Anatomy. Philadelphia, WB Saunders, 1980, p 192

14. Manani G, Angel A, Civran E, Tegazzin V, Dal Vecchio A, Giron GP: Sciatic nerve block by the anterior and posterior approach for operations on the lower extremity. A comparative study. Acta Anaesthesiol Belg 1982; 3:18393 15. Davies MJ, McGlade DP: One hundred sciatic nerve blocks: A comparison of localisation techniques. Anaesth Intens Care 1993; 21:768 16. Mansour NY, Bennetts FE: An observational study of combined continuous lumbar plexus and single-shot sciatic nerve blocks for post-knee surgery analgesia. Reg Anesth 1996; 21:28791 17. Allen HW, Liu SS, Ware PD, Nairn CS, Owens BD: Peripheral nerve blocks improve analgesia after total knee replacement surgery. Anesth Analg 1998; 87:937 18. Coventry DM, Todd JG: Alkalinisation of bupivacaine for sciatic nerve blockade. Anesthesia 1989; 44:46770 19. Fanelli G, Casati A, Aldegheri G, Beccari P, Bertl M, Leoni A, Torri G: Cardiovascular effects of two different regional anaesthetic techniques for unilateral leg surgery. Acta Anaesthesiol Scand 1998; 42:80 4 20. Casati A, Fanelli G, Borghi B, Torri G: Ropivacaine or 2% mepivacaine for lower limb peripheral nerve blocks. ANESTHESIOLOGY 1999; 90:104752 21. Pavlin DJ, Rapp SE, Polissar NL, Malmgren JA, Koerschgen M, Keyes H: Factors affecting discharge time in adult outpatients. Anesth Analg 1998; 87:816 26 22. Hadzic A, Reiss W, Dilberovic F, April EW, Koorn R, Thys DM, Vloka JD: Rotation of the leg inuences ability to approach the sciatic nerve through the anterior approach. Reg Anesth Pain Med 1998; 23:38

Anesthesiology, V 91, No 6, Dec 1999

Anatomical Landmarks for Femoral Nerve Block: A Comparison of Four Needle Insertion Sites
Jerry D. Vloka, MD, PhD*, Admir Hadz ic , MD, PhD*, Leon Drobnik, April Ernest, PhD, Wojciech Reiss, MD*, and Daniel M. Thys, MD*
MD, PhD, *Departments of Anesthesiology, St. LukesRoosevelt Hospital Center, Columbia University College of Physicians and Surgeons, New York, New York; Medical School, Poznan, Poland; and Department of Anatomy, Columbia University, New York, New York

The site for needle insertion in femoral nerve block varies significantly among various descriptions of the technique. To determine the site with the highest likelihood of needle-femoral nerve contact, femoral nerve block was simulated in a human cadaver model (17 femoral triangles from 9 adult cadavers). Four 20-gauge 50-mmlong styletted catheters were inserted at four frequently suggested insertion sites for femoral nerve block. At the levels of inguinal ligament and the inguinal crease, the catheters were inserted adjacent to the lateral border of the femoral artery and 2 cm lateral to the femoral artery. During anatomical dissection, we studied the number of catheter-nerve contacts for each of the four insertion sites, and relationships between the femoral nerve and other anatomical structures of relevance to femoral nerve block. Insertion of the needle at the level of the inguinal crease, next to the lateral border of the femoral

artery resulted in the highest frequency of needlefemoral nerve contacts (71%). Of note, the femoral nerve was significantly wider (14.0 vs 9.8 mm) and closer to the fascia lata (6.8 vs 26.4 mm) at the inguinal crease than at the inguinal ligament level. We conclude that needle insertion at the inguinal crease level immediately adjacent to the femoral artery produced the highest rate of needle-femoral nerve contacts. The main factors influencing this result include the greater width of the femoral nerve and the more predictable femoral artery-femoral nerve relationship at the inguinal crease level, compared with the inguinal ligament level. Implications: Insertion of a needle at the inguinal crease level and immediately adjacent to the lateral border of the femoral artery results in a high rate of needlefemoral nerve contact. (Anesth Analg 1999;89:146770)

he femoral nerve block remains an infrequently used anesthesia technique despite its many clinical indications (1). Possible obstacles to its wider use may include unfavorable local practice settings (2), the need for additional blocks, and inconsistent surgical anesthesia, which are frequently reported with this method (35). Suggested needle insertion sites vary widely among published descriptions of the technique (6 9). However, it is not known whether these insertion sites are equally effective for prompt localization of the femoral nerve. In this study, based on human cadaver models, we compared the frequency of needle-femoral nerve contacts on single needle insertions at four frequently used insertion

sites. We also studied the spatial relationship among anatomical structures of relevance to femoral nerve block, and postulated how these relationships might influence the performance of the block.

Methods
This experiment was performed on the lower extremities of nine adult cadavers, all dead between 6 and 18 mo. All had been embalmed for anatomical purposes using a solution of phenol (13%) as the principal fixative, and glycerin (28%) for retention of water content. Cadavers were positioned supine on the dissecting table with the long axes of the legs horizontal to the table and the feet forming a 90 angle to the horizontal plane. The anterior superior iliac spine, pubic tubercle, and the midline of the pubic symphysis were marked by colored pins. The inguinal crease was identified as a skin fold distal to the inguinal ligament and parallel to it, medially intersecting the junction of the scrotal (labial) fold with the thigh (Fig. 1).
Anesth Analg 1999;89:146770

Presented in part at the 23rd Annual Meeting of the American Society of Regional Anesthesia in Seattle, WA, May 16, 1998, and published in abstract form (Reg Anesth Pain Med 1998;23:53). Accepted for publication July 27, 1999. Address correspondence and reprint requests to Admir Hadz ic , MD, PhD, Department of Anesthesiology, St. LukesRoosevelt Hospital Center, 1111 Amsterdam Ave., New York, NY 10025. Address e-mail to ah149@columbia.edu.
1999 by the International Anesthesia Research Society 0003-2999/99

1467

1468

REGIONAL ANESTHESIA AND PAIN MANAGEMENT VLOKA ET AL. NEEDLE INSERTION SITE FOR FEMORAL NERVE BLOCK

ANESTH ANALG 1999;89:146770

Table 1. The Incidence of Needle-Nerve Contacts After Simulation of Femoral Nerve Block at Four Different Insertion Sites % Inguinal ligament Adjacent to FA* 20 mm lateral to FA Inguinal crease Adjacent to FA 20 mm lateral to FA 12 24 71 0

FA Femoral artery, inguinal ligament a ligamentous stretch between the anterior-superior iliac spine and medial pubic tubercle, inguinal crease a skin fold distal to the inguinal ligament and parallel to it, medially intersecting the junction of the scrotal (labial) fold with the thigh. * P .05; P .01; P .001.

Figure 1. Anatomical landmarks for femoral nerve block; needle insertion sites. IL inguinal ligament (line between the anterosuperior iliac spine and the pubic tubercle), IC inguinal crease (a natural skin fold 4 6 cm below the inguinal ligament), FA femoral artery, numbers 1 through 4 needle insertion sites.

The skin and fat of the anterior thigh were removed to the level of the fascia lata to identify the femoral triangle with its borders: sartorius muscle laterally, adductor longus muscles medially, and inguinal ligament superiorly (10). Subcutaneous tissue was removed gradually until the anterior wall of the femoral artery was identified. At this level, the femoral nerve could not be identified in any of the cadavers. Four needles with 20-gauge 50-mm-long catheters over the needle were inserted through four different insertion sites to simulate the femoral nerve block (Fig. 1, Table 1). The needles with catheters were advanced perpendicularly to the horizontal plane to a depth of 50 mm or until bone was contacted. The needles were then removed, but the catheters were left in situ. Further anatomical dissection of the inguinal triangle was conducted to determine the number of catheter-femoral nerve contacts for each of the four insertion sites. The contact between the catheter and the femoral nerve was considered positive if the catheter had contacted

or pierced the femoral nerve. During anatomical dissections, preservation of the fascia lata was given special attention to facilitate precise measurements of the distances from fascia lata to femoral nerve and femoral artery. After dissection of the femoral triangles was completed, the sizes of the femoral artery and femoral nerve were obtained and distances of the femoral nerve to femoral artery and fascia lata were measured (Table 2). Because the femoral nerve has an oval-like profile, the shortest femoral nerve-femoral artery and femoral nerve-fascia lata distances were used. Data are expressed as means sd for continuous measures. The differences in the spatial relationship (distances) of the relevant anatomical structures at the inguinal ligament and inguinal crease levels were analyzed using the paired t-test. As the needle-femoral nerve contacts were evaluated in the same 17 cadaver legs, analyses were conducted by a series of McNemars 2 tests (one leg was excluded because extensive scarring in the femoral region precluded precise measurements). The relationships of the four insertion sites were tested in all pair-wise combinations. For instance, the number of needle-femoral nerve contacts at one point (e.g., insertion at the inguinal ligament and adjacent to the femoral artery) was compared to the number of contacts observed at each of the other three insertions. Statistical analyses were performed using the Statistical Package for the Social Sciences (SPSS for Windows, version 5.0.2, Chicago, IL).

Results
The study material comprised 17 femoral triangles from nine cadavers (four female, five male) of average body frame and nutrition. As shown in Table 1, the proportion of needle-femoral nerve contacts when the needle was inserted at the inguinal crease and immediately to the lateral border of the femoral artery was significantly higher than that observed at any of the other three tested points. Moreover, the rates of

ANESTH ANALG 1999;89:146770

REGIONAL ANESTHESIA AND PAIN MANAGEMENT VLOKA ET AL. NEEDLE INSERTION SITE FOR FEMORAL NERVE BLOCK

1469

Table 2. Relationship of the FN and the FA to the Surrounding Structures in the Femoral Triangle at Various Levels Level Distance (mm) FAfascia lata Femoral nervefascia Femoral nerve width Inguinal crease 3.2 1.5 6.8 2.7 14.0 2.0 Inguinal ligament 19.7 6.4 26.4 7.0 9.8 2.0 P value 0.01 0.01 0.01

Values are mean sd. FN femoral nerve, FA femoral artery, inguinal ligament a ligamentous stretch between the anterior-superior iliac spine and medial pubic tubercle, inguinal crease a skin fold distal to the inguinal ligament and parallel to it, medially intersecting the junction of the scrotal (labial) fold with the thigh.

needle-femoral nerve contacts of the other three insertions were not significantly different by pair-wise comparisons. There were no needle-femoral nerve contacts when the needle was inserted at the inguinal crease (20 mm lateral to the femoral artery). In all cadaver legs, the femoral nerve was medial to this insertion point. The femoral nerve was more superficial (closer to the fascia lata) at the level of the inguinal crease than at the level of the inguinal ligament (6.8 2.7 vs 26.4 7.0 mm; P 0.01) (Table 2). The femoral nerve was also significantly wider (14.0 2.0 vs 9.8 2.0 mm; P 0.01) at the inguinal crease than at the inguinal ligament level. It should be noted that the femoral artery-femoral nerve relationship was more consistent at the inguinal crease than at the inguinal ligament. At the inguinal crease level, the femoral nerve was immediately adjacent to the femoral artery. Indeed, in 70.6% of the cases, the femoral nerve was partially covered by the femoral artery. In contrast, at the level of the inguinal ligament, the femoral artery partially covered the nerve in only 11.7% of cases and the distance between the two structures varied from 0.0 to 13.0 mm (median 8.5).

Discussion
The preferred site of insertion in femoral nerve block varies widely. Moore (8) suggested an insertion site at 2.5 cm below the inguinal ligament and just lateral to the femoral artery. In another common technique, the needle is inserted at the lateral border of the femoral artery, but at the inguinal ligament level (9). Our results suggest that the femoral nerve is most consistently localized when the needle is inserted at the inguinal crease level and immediately lateral to the femoral artery. Based on anatomical measurements, this is mainly because of the intimate and predictable femoral nerve-femoral artery relationship, as well as the greater width of the femoral nerve at this level.

These anatomical characteristics, along with the superficial position of the femoral nerve at the inguinal crease level, should facilitate performance of femoral nerve block in patients, with fewer attempts required to localize the femoral nerve. In patients, however, the use of a nerve stimulator may substantially facilitate identification of the femoral nerve at other sites. It is important to note that some of the branches of the femoral nerve to the sartorius muscle and neighboring skin start departing the main trunk of the nerve as the nerve emerges under the inguinal ligament. However, the bulk of the femoral nerve remains enveloped by the epineural sheath at the inguinal crease level. The limitation of the study was the possibility that the anatomy of embalmed cadavers was somewhat distorted. Specifically, the blood vessels did not contain blood, and were partially collapsed. However, the embalming process typically affects veins more than arteries because of the thicker and more rigid arterial vessel wall, and the diameter of the arteries should be affected similarly at both the inguinal ligament and inguinal crease levels. Additionally, the mere finding of the catheter tip-nerve contact in our cadaveric study is no guarantee of a successful blockade in patients. However, our preliminary data of a follow-up clinical study testing the ease of performing femoral nerve block through the approach at the inguinal crease level are in agreement with the anatomical results. In this study, a needle connected to a low-output nerve stimulator was inserted at the inguinal crease level, immediately lateral to the femoral artery and directed in the sagital plane at 60 degrees cephalad. In 100 consecutive patients, this technique resulted in identification of the femoral nerve (rhythmic movement of the patella) on one or two attempts in more than 90% of patients (11). Injection of 30 milliliters of local anesthetic after the quadriceps muscle twitch was obtained at 0.4 milliampere or less resulted in a 100% success rate (inability to extend the leg in the knee joint and insensitivity to touch in the anterior thigh and medial leg below the knee). In summary, insertion of the needle at the inguinal crease level and immediately to the lateral border of the femoral artery in adult cadavers resulted in a high frequency of needle-femoral nerve contacts. Additional advantages of approaching the femoral nerve at this location are the greater width and superficial location of femoral nerve, which both should facilitate localization of the femoral nerve.

The authors thank Gorica Hadz ic and Henry Shih for their assistance in preparation of this manuscript.

1470

REGIONAL ANESTHESIA AND PAIN MANAGEMENT VLOKA ET AL. NEEDLE INSERTION SITE FOR FEMORAL NERVE BLOCK

ANESTH ANALG 1999;89:146770

References
1. Hadz ic A, Vloka JD, Kuroda MM, et al. The use of peripheral blocks in anesthesia practice: a national survey. Reg Anesth Pain Med 1998;23:241 6. 2. Kopacz DJ, Bridenbaugh LD. Are anesthesia residency programs failing regional anesthesia? The past, present, and future. Reg Anesth 1993;18:84 7. 3. Macrae WA. Lower limb blocks. In: Wildsmith JAW, Armitage EN, eds. Principles and practice of regional anesthesia. Edinburgh: Churchill Livingstone, 1987:155 67. 4. Moore DC. Block of the sciatic and femoral nerves. In: Moore DC, ed. Regional block. Springfield: Charles C. Thomas, 1961: 219 31. 5. Brown DL. Femoral block. In: Brown DL, ed. Atlas of regional anesthesia. Philadelphia: W.B. Saunders, 1992:89 95. 6. Carron H, Korbon GA, Rowlingson JC. Lumbar plexus block. In: Carron H, Korbon GA, Rowlingson JC, eds. Regional anesthesia: techniques and clinical applications. Orlando: Grune & Stratton, 1984:2731.

7. Anker-Moller E, Spangsberg N, Dahl JB, et al. Continuous blockade of the lumbar plexus after knee surgery: a comparison of the plasma concentrations and analgesic effect of bupivacaine 0.250% and 0.125%. Acta Anaesthesiol Scand 1990;34:468 72. 8. Moore DC. Block of the sciatic and femoral nerves. In: Moore DC, ed. Regional block. 4th ed. Springfield: Charles C. Thomas, 1965:281 8. 9. Brown DL. Regional anesthesia. In: Kirby RR, Gravenstein N, eds. Clinical anesthesia practice. Philadelphia: W.B. Saunders, 1994:514 41. 10. Khoo ST, Brown TCK. Femoral nerve block: anatomical basis for a single injection technique. Anaesth Intensive Care 1983; 11:40 2. 11. Vloka JD, Hadzic A, Reiss W, et al. Femoral nerve block: the sartorius muscle twitch can be used as a reliable guide to obtain a quadriceps muscle response. Anesthesiology 1998;89:3.

AMBULATORY ANESTHESIA
SECTION EDITOR PAUL F. WHITE

SOCIETY

FOR

AMBULATORY ANESTHESIA

A Comparison of Spinal, Epidural, and General Anesthesia for Outpatient Knee Arthroscopy
Michael F. Mulroy, MD, Kathleen L. Larkin, MD, Peter S. Hodgson, MD, James D. Helman, MD, Julia E. Pollock, MD, and Spencer S. Liu, MD
Department of Anesthesiology, Virginia Mason Medical Center, Seattle Washington

We compared general, epidural, and spinal anesthesia for outpatient knee arthroscopy (excluding anterior cruciate ligament repairs). Forty-eight patients (ASA physical status IIII) were randomized to receive either propofol-nitrous oxide general anesthesia with a laryngeal mask airway with anesthetic depth titrated to a bispectral index level of 40 60, 1520 mL of 3% 2-chloroprocaine epidural, or 75 mg of subarachnoid procaine with 20 g fentanyl. All patients were premedicated with 0.035 mg/kg midazolam and 1 g/kg fentanyl and received intraarticular bupivacaine and 1530 mg of IV ketorolac during the procedure. Recovery times, operating room turnover times, and patient satisfaction were recorded by an observer using an objective scale for recovery assessment and a verbal rating scale for satisfaction. Statistical analysis was performed with analysis of variance and 2. Postanesthesia care unit

discharge times for the general and epidural groups were similar (general 104 31 min, epidural 92 18 min), whereas the spinal group had a longer recovery time (146 52 min) (P 0.0003). Patient satisfaction was equally good in all three groups (P 0.34). Room turnover times did not differ among groups (P 0.16). There were no anesthetic failures or serious adverse events in any group. Pruritus was more frequent in the spinal group (7 of 16 required treatment) than in the general or epidural groups (no pruritus) (P 0.001). We conclude that epidural anesthesia with 2-chloroprocaine provides comparable recovery and discharge times to general anesthesia provided with propofol and nitrous oxide. Spinal anesthesia with procaine and fentanyl is an effective alternative and is associated with a longer discharge time and increased side effects. (Anesth Analg 2000;91:860 4)

utpatient arthroscopic knee surgery can be performed with general or regional anesthesia. Epidural anesthesia (1) and peripheral nerve block (2) have both provided more rapid discharge than general anesthesia in previous reports, whereas spinal anesthesia with bupivacaine (3) has also been shown to provide discharge times comparable to general anesthesia. Recent data, however, suggest that spinal and epidural anesthesia require longer discharge times than the newer shorter-acting general anesthetic drugs (propofol and sevoflurane) (4,5). Published reports have not compared optimal techniques in each category. We hypothesized that the ideal selection of local anesthetic drugs for either epidural or spinal anesthesia would provide recovery times comparable to those obtained with short-acting general anesthesia and result in equal patient satisfaction. We compared
Accepted for publication May 17, 2000. Address correspondence and requests for reprints to Michael F. Mulroy, MD, Department of Anesthesiology, B2-AN, 1100 Ninth Ave., PO Box 900, Seattle, WA 98111. Address e-mail to anemfm @vmmc.org.

discharge times, side effects, operating room (OR) efficiency, and patient satisfaction levels of three anesthetic techniques, each performed with an ideal drug, for outpatient arthroscopy in a prospective, randomized fashion.

Methods
After institutional review board approval, patients scheduled for elective outpatient unilateral knee arthroscopy, not including anterior cruciate ligament repair, were asked to participate in a prospective, randomized comparison of anesthetic techniques. All patients were between 21 and 60 yr of age, ASA physical status IIII, and 110 kg. Patients were English speaking, available for follow-up phone interviews, and had no medical contraindication to any of the three anesthetic techniques (allergy, bleeding disorders, localized infection, neurologic disease). After informed consent, patients were randomized by sealed envelopes to receive either general, epidural, or spinal anesthesia. All patients were premedicated with a
2000 by the International Anesthesia Research Society 0003-2999/00

860

Anesth Analg 2000;91:8604

ANESTH ANALG 2000;91:860 4

AMBULATORY ANESTHESIA MULROY ET AL. ANESTHESIA FOR KNEE ARTHROSCOPY

861

maximum of IV 0.035 mg/kg midazolam and 1 g/kg fentanyl. General anesthesia was performed with 2 mg/kg IV propofol induction and 60% nitrous oxide by laryngeal mask airway with a continuous infusion of propofol titrated to maintain the bispectral index monitor reading between 40 and 60. The infusion was discontinued when the trocars were removed from the knee. Epidural anesthesia was performed in a standard fashion at the L2-3 or L3-4 interspace with the operative knee in the dependent position. Skin infiltration was performed with 1% lidocaine, and a test dose of 3 mL of 1.5% lidocaine with 15 g epinephrine was injected. If there was no evidence of IV or subarachnoid injection, 15 mL of 3% 2-chloroprocaine was injected in 5 mL increments, with an additional 5 mL added after 10 min, if the block height was below T-10 (level needed to provide anesthesia for thigh tourniquet discomfort). Spinal anesthesia was performed at the L2-3 or L3-4 interspace with a 25-gauge Whitacre needle with the operative knee dependent. After free flow of cerebrospinal fluid, 75 mg of procaine plus 10 20 g of fentanyl diluted with an equal volume of cerebrospinal fluid was injected, and the patient turned supine. Patients receiving spinal or epidural anesthesia who requested sedation were given an intraoperative infusion of propofol (2550 g kg1 min1). All patients received 1530 mg IV ketorolac. The surgeon injected 50 mL of 0.25% bupivacaine into the knee joint at the completion of the procedure. When the surgical procedure was performed by the same surgeon sequentially in the same OR, turnover time was measured. This was defined as the time interval from the departure of the previous patient from the OR until the completion of anesthesia preparation. All patients were transferred by stretcher to the Phase I postanesthesia care unit (PACU). When vital signs were stable for two measurements and block level (for regional block patients) was below T-8, they were transferred to the Phase II area. Discharge time was recorded as the time from admission to PACU until the patient met all discharge criteria from Phase II. These included mental alertness, stable vital signs, absence of nausea, control of pain, ability to ambulate, and (for regional techniques) voiding. Side effects measured were the incidence of hypotension (systolic blood pressure 100 mm Hg requiring treatment with ephedrine at the discretion of the anesthesiologist) or bradycardia (heart rate 60 requiring treatment with atropine at the discretion of the anesthesiologist) in either the OR or in the PACU, nausea or vomiting, pain severe enough to require IV narcotics in the PACU, or pruritus requiring treatment. Patient satisfaction (rated on a verbal scale of 5 very satisfied, 4 satisfied, 3 neutral, 2 dissatisfied, 1 very dissatisfied) and specific side effects were

evaluated by follow-up phone call 24 h after the procedure. For data analysis, discharge time was considered the primary outcome variable, with a difference considered significant at the P 0.05 level. The secondary variable was the efficiency, as estimated by the room turnover time. These were compared by analysis of variance for the three groups and confirmed with Wilcoxons nonparametric test. The frequency of side effects among the groups was compared by 2 analysis with Fishers exact test. A power analysis indicated that a sample size of 16 patients per group was required to show a 45 min difference in discharge time among groups at the P 0.05 level with 80% power. This was based on a similar study of procaine spinal anesthesia for knee arthroscopy performed in the same outpatient unit with a standard deviation of 47 min for discharge time (6).

Results
Between March 25, 1999, and December 21, 1999, we recruited 192 patients to participate. One hundred forty-one patients refused randomization (95 requesting regional anesthesia and 46 preferring general anesthesia); 51 patients accepted randomization. There were no differences in age, weight, sex, ASA physical status, midazolam administration, or fentanyl administration among groups (Table 1). Three were excluded from analysis (one overnight admission, two protocol violations). All operations were performed by one of three surgeons. There were no major surgical or anesthetic complications. Anesthesia was satisfactory for incision in all three groups; one patient in the epidural group was given an intraoperative bolus of propofol at the request of the surgeon to provide greater muscle relaxation. Four of 16 epidurals were redosed during the case. Of 32 patients receiving regional anesthesia, 13 were given propofol infusions for sedation (9 epidural, 4 spinal). In the PACU, nine patients required IV narcotics for pain (five general, two epidural, two spinal). Antiemetics were necessary for three spinal patients. Sedation was reported as prolonging Phase I stay in one patient in the general group. Antipruritics were given to seven patients, all in the spinal group (P 0.001). Voiding was required for discharge after regional anesthesia. Time from injection to voiding was shorter in the epidural group versus the spinal group (144 22 vs 193 50 min, P 0.0013). The time in the PACU was also significantly shorter for general and epidural as compared with spinal (general, 104 31 min; epidural, 92 18 min; spinal, 146 52 min; P 0.0003 with Scheffe s adjusted P value of 0.008 for general versus spinal and 0.0006 for epidural versus spinal) (Table 2).

862

AMBULATORY ANESTHESIA MULROY ET AL. ANESTHESIA FOR KNEE ARTHROSCOPY

ANESTH ANALG 2000;91:860 4

Table 1. Demographic Data General (n 16) Epidural (n 16) Spinal (n 16)

42 14 42 13 46 13 Age (yr)a Weight (kg)a 83 15 79 14 82 16 Male (%) 56 69 63 ASA physical status 12/4/0 12/3/1 15/1/0 (I/II/III) Midazolam (mg)a 1.7 0.7 2.4 1.5 2.2 1.0 Fentanyl (g)a 71.6 28.7 55.5 33.8 36.3 41.1 Operative time (min)a 32 17 27 11 31 11
a

Average sd.

In follow-up evaluation, 44 of 48 patients were satisfied with the care they received. (Table 2) Four patients complained of mild/moderate backache on follow-up phone call (3 epidural, 1 spinal). No patients had backache immediately after block resolution or back pain radiating to the legs (transient neurologic symptoms [TNS]) after discharge. One patient in the spinal group had a positional headache which did not require treatment. Minor side effects were infrequent (Table 2). A total of 42 patients said they would choose the same anesthetic again; three general anesthesia patients said they would not because they would want to watch the procedure.

Discussion
General and regional anesthesia techniques have been used for outpatient knee arthroscopy. To date, however, there has not been a direct comparison of the new general anesthetics to ideal regional anesthetic techniques. Parnass et al. (1) used isoflurane for general anesthesia, rather than the newer short-acting anesthetics. New general anesthetics including propofol, desflurane, and sevoflurane allow for faster emergence after general anesthesia than thiopental and isoflurane, and the potential for earlier discharge in the outpatient setting. Pavlin et al. (4) reported that general anesthesia with these drugs allowed an earlier discharge compared with spinal and epidural anesthesia in her practice (184 vs 202 minutes for men, 185 vs 213 minutes for women), but she did not specifically study knee arthroscopy, nor the use of short duration neuraxial regional anesthesia techniques. Luttropp et al. (5) reported that for knee arthroscopy, general anesthesia with either propofol or sevoflurane provides faster recovery than spinal anesthesia (116 and 141 vs 176 minutes), although with a higher frequency of postoperative nausea and vomiting with sevoflurane. Regional anesthesia has been used successfully for outpatient knee arthroscopy. Peripheral nerve block of the lumbar plexus block at the groin provides satisfactory anesthesia and rapid discharge (2,7), as is consistent with other comparisons of peripheral nerve

blocks to general anesthesia (4). Peripheral nerve blocks, however, may take longer to perform, be less familiar to the practitioner, have a higher failure rate (7), and may have a slow onset of anesthesia (8). Neuraxial blocks are simpler to perform and are used more often in the outpatient setting. Epidural anesthesia allows the titration of short duration local anesthetics to provide potentially rapid discharge. Parnass et al. (1) demonstrated faster discharge with lidocaine epidural anesthesia than general anesthesia with isoflurane (159 vs 208 minutes). Dahl et al. (9) compared general anesthesia with propofol to spinal and epidural anesthesia, but did not report actual discharge times and used the longer-acting mepivacaine to provide epidural anesthesia. Epidural anesthesia with chloroprocaine provides faster resolution compared with lidocaine (127 vs 195 minutes) (10), and may be a superior choice. Although back pain has been reported with this drug, it appears to be related to the use of large doses (11), and may not be a problem when doses of 25 mL or less are used. Thus, epidural anesthesia with chloroprocaine may offer the current optimal clinical practice for knee arthroscopy. Spinal anesthesia is also advocated for outpatient surgery because of its reliability and simplicity, but suffers the limitation of a single injection technique. Larger doses may be required for adequate anesthesia, but may be associated with longer duration of action. Luttropp et al.s (5) relatively large dose of spinal anesthesia with 70 80 mg of lidocaine may account for longer discharge time compared with general anesthesia. Selection of an optimal drug and dose is problematic because of a dearth of comparative studies. Lidocaine has been the traditional short-acting spinal anesthetic. Small-dose (40 mg) subarachnoid lidocaine provides discharge times equivalent to epidural chloroprocaine (12), but with a 10% failure rate. Reports of TNS after use of lidocaine for lithotomy and arthroscopy procedures have led to a search for alternative spinal anesthetics (13). Ben-David et al. (14) found that 5 mg bupivacaine with 10 g of fentanyl avoided TNS and provided satisfactory anesthesia for knee surgery, but with discharge times comparable to reports of duration of spinal anesthesia with 50 mg lidocaine (15,16). Procaine has been regarded as a shorter-acting spinal anesthetic, and has been studied recently as an alternative to lidocaine. Hodgson and Liu (6) reported that subarachnoid 100 mg hyperbaric procaine is associated with less TNS, but a higher failure rate and longer discharge time than 50 mg lidocaine. Axelrod et al. (17) have reported shorter recovery times with doses of 60 and 80 mg procaine when fentanyl is added, but block heights below T-10 with the 60 mg dose. From these data, we estimated that a 75 mg dose of procaine with fentanyl might provide less risk of TNS and the best chance of a

ANESTH ANALG 2000;91:860 4

AMBULATORY ANESTHESIA MULROY ET AL. ANESTHESIA FOR KNEE ARTHROSCOPY

863

Table 2. Results General (n 16) Turnover time (min) Time to voidb (min)a Time to discharge (min)a Hypotension/bradycardia in the operating roomd Postanesthesia care unit IV Narcotics Antiemetics Antiemetics Antipruritic Follow-up Headache Pain control Back/leg pain Satisfaction scores: 5 (very satisfied) 4 (satisfied) 3 (neutral) 2 (dissatisfied)
Average sd. Time from postanesthetic care unit admission to void. c HR 60, systolic blood pressure 100 mm Hg requiring treatment. d One patient had a positional headache; none required treatment. *P 0.05.
a b

Epidural (n 16) 23 6 (n 5) 80 16* 92 18 2 2 0 0 0 3 1 2 12 3 1

Spinal (n 16) 28 9 (n 3) 135 51* 146 52* 2 2 3* 3* 7* 3c 1 1 8 6 1 1

24 6 (n 8) NA 104 31 2 5 0 0 0 1 1 0 12 3 1 0

reliable block above T-10 with optimal discharge duration for outpatient spinal anesthetic. In our study of the three techniques, propofol/ nitrous oxide general anesthesia provided a duration of recovery similar to that after epidural anesthesia with 2-chloroprocaine. Spinal anesthesia required on average 42 to 54 minutes longer for recovery than the general or epidural group. The spinal group also had a higher incidence of side effects, specifically pruritus, which was attributed to the addition of the fentanyl to the procaine, as well as one occurrence of positional headache in the 16 patients. Room turnover times were not different among the groups, although the number of procedures performed sequentially in the same OR was small in our study, and does not allow significant conclusions. The level of patient satisfaction was equally high among all three groups. No patients in the epidural group had lower back pain after chloroprocaine, nor were there any symptoms of TNS in the spinal group. No patients had urinary retention in the PACU. Omitting the requirement for voiding would have reduced the discharge times for both the epidural and spinal group, but further study is needed to substantiate the safety of such a practice (18). It appears that the use of either propofol general anesthesia or chloroprocaine epidural anesthesia provide rapid discharge for outpatient knee arthroscopy. The choice between these two may be based on the patients desire to be awake and alert during the surgical procedure. There are several potential limitations of our study. First, it was impossible to perform this study in a

blinded fashion because of the nature of the anesthetics. Outcome data points (discharge times), however, were collected by an observer every 15 minutes using objective PACU recovery criteria. Room turnover times were likewise an objective time measure. A second problem was a limited number of comparisons in room turnover times. In our outpatient surgical center, surgeons often do not follow themselves in the same room. Nevertheless, for the limited number of observations, there was no difference in the turnover times among groups. Epidural anesthesia in our setting is performed in an induction area outside the OR, while the OR itself is being prepared for the next case. Subsequent transfer of the patient to the OR allows additional time for onset of anesthesia. Longer turnover times may be found if such an area is not available. Although our study may provide a valid comparison of ideal general and epidural anesthesia techniques, we may not have identified the ideal local anesthetic or dose for outpatient spinal anesthesia. The shortest previously reported duration is with 40 mg of lidocaine (12), but this dose was associated with a 10% failure rate and lidocaine has been associated with symptoms of TNS. Bupivacaine has been recommended as an alternative, but its duration is equivalent to 50 mg lidocaine anesthesia. We chose a relatively small dose of procaine with fentanyl to reduce the probability of TNS and shorten discharge, but still found discharge times comparable to 50 mg lidocaine, with a high side effect profile that included nausea, pruritus, and positional headache. Further study is needed to determine whether there is a

864

AMBULATORY ANESTHESIA MULROY ET AL. ANESTHESIA FOR KNEE ARTHROSCOPY

ANESTH ANALG 2000;91:860 4

drug or dose regimen that would provide an equivalent duration and side effect profile to epidural chloroprocaine. We had a relatively low recruitment to our study (51 of 192), which might suggest a bias of the participants. However, we found that those who did consent to randomization appeared to be least prejudiced as to the relative advantages of the techniques. The patients with strong convictions about the choice of anesthetic did not participate in the study. On the other hand, this patient prejudice that we encountered could be a significant factor in the choice of anesthetic. The majority of our patients who refused randomization (95 of 141) preferred being awake, observing the arthroscopic procedure on the video screen. In an outpatient population, this preference for alertness may serve as a primary criteria for the choice of anesthetic technique, based on our data of equivalent side effects and recovery times. In conclusion, we performed a prospective, randomized comparison of general, epidural and spinal anesthesia for outpatient knee arthroscopy. Epidural anesthesia with chloroprocaine and general anesthesia with propofol-nitrous oxide provided equally effective intraoperative conditions and PACU discharge times in our outpatient center. Spinal anesthesia with 75 mg of procaine with fentanyl in this setting was associated with an average of 42 to 54 minutes longer discharge times than the other two techniques, and a higher incidence of side effects. Further study will be needed to identify an appropriate spinal anesthetic that would produce discharge times equivalent to the results obtained with general or epidural anesthesia in this setting. Either general or epidural anesthesia provide satisfactory anesthesia for outpatient knee arthroscopy and the choice of anesthetic technique may be primarily dependent on the patients desire to be alert and participatory during the surgical procedure.
The authors wish to thank Deanna Boltz, RN, for help in collection of data, and David Kerr for statistical analysis.

2. Patel NJ, Flashburg MH, Paskin S, Grossman R. Regional anesthetic technique compared to general anesthesia for outpatient knee arthroscopy. Anesth Analg 1986;65:1857. 3. Ben-David B, Levin H, Solomon E, et al. Spinal bupivacaine in ambulatory surgery: The effect of saline dilution. Anesth Analg 1996;83:716 20. 4. Pavlin DJ, Rapp SE, Polissar NL, et al. Factors affecting discharge time in adult outpatients. Anesth Analg 1998;87:816 26. 5. Luttropp HH, Orlanders K, Ikonomidou E. Spinal, sevoflurane, or propofol anesthesia for outpatient knee arthroscopy [abstract]. Anesthesiology 1998;89:A41. 6. Hodgson PS, Liu SS, Batra MS, et al. Procaine compared to lidocaine for incidence of TNS. Reg Anesth Pain Med 2000;25: 218 22. 7. Goranson BD, Lang S, Cassidy JD, et al. A comparison of three regional anaesthesia techniques for outpatient knee arthroscopy. Can J Anaesth 1997;44:371 6. 8. Casati A, Fanelli G, Borghi B, Torri G. Ropivacaine or 2% mepivacaine for lower limb peripheral nerve blocks. Anesthesiology 1999;90:104752. 9. Dahl V, Gierloff C, Omland E, Raeder JC. Spinal, epidural or propofol anaesthesia for outpatient knee arthroscopy? Acta Anaesthesiol Scand 1997;41:13415. 10. Kopacz D, Mulroy MF. Chloroprocaine and lidocaine decrease hospital stay and admission rate after outpatient epidural anesthesia. Reg Anesth 1990;15:19 25. 11. Stevens RA, Urmey WF, Urquarht BL, Kao TC. Back pain after epidural anesthesia with chloroprocaine. Anesthesiology 1993; 78:4927. 12. Urmey WF, Stanton J, Peterson M, Sharrock NE. Combined spinal-epidural anesthesia for outpatient surgery: doseresponse characteristics of intrathecal isobaric lidocaine using a 27-gauge Whitacre spinal needle. Anesthesiology 1995;83: 528 34. 13. Freedman JM, Li DK, Drasner K. Transient neurologic symptoms after spinal anesthesia. Anesthesiology. 1998;89:633 641. 14. Ben-David B, Solomon E, Levin H, et al. Intrathecal fentanyl with small-dose dilute bupivacaine: better anesthesia without prolonging recovery. Anesth Analg. 1997;85:560 5. 15. Chiu AA, Liu S, Carpenter RL, et al. The effects of epinephrine on lidocaine spinal anesthesia: a crossover study. Anesth Analg 1995;80:7359. 16. Liu S, Chiu AA, Carpenter RL, et al. Fentanyl prolongs lidocaine spinal anesthesia without prolonging recovery. Anesth Analg 1995;80:730 4. 17. Axelrod EH, Alexander GD, Brown M, Schork MA. Procainespinal anesthesis: a pilot study of the incidence of transient neurologic symptoms. J Clin Anesth. 1998;10:404 9. 18. Pavlin DJ, Pavlin KG, Gunn HC. Voiding in patients managed with or without ultrasound monitoring of bladder volume after outpatient surgery. Anesth Analg 1999;89:90 7.

References
1. Parnass SM, McCarthy RJ, Bach BR, et al. Beneficial impact of epidural anesthesia on recovery after outpatient arthroscopy. Arthroscopy 1993;9:915.

Anterior Approach to the Sciatic Nerve Block: The Effects of Leg Rotation
Jerry D. Vloka, MD, PhD*, Admir Hadz ic , Daniel M. Thys, MD*
MD, PhD*,

Ernest April,

PhD,

and

*St. Lukes-Roosevelt Hospital Center, Columbia University College of Physicians and Surgeons, New York, New York; and College of Physicians and Surgeons, Columbia University, New York, New York

In the anterior approach to the sciatic nerve block, the femur often obstructs the passage of the needle toward the sciatic nerve. In this study, by using a human cadaver model, we assessed how internal and external rotation of the leg influences the accessibility of the sciatic nerve with the anterior approach. Ten lower extremities from five adult cadavers were studied. Needles were used to simulate the anterior approach to the sciatic nerve block. The effect of leg rotation on the needle plane required to reach the sciatic nerve was studied with legs in the neutral position and then with internal and external rotation (45) of the legs. During needle

placement in the neutral position, the needle could not be fully advanced to the level of the sciatic nerve because of obstruction by the lesser trochanter in 80% of attempts. Medial redirection of the needle (1015) allowed it to pass the lesser trochanter but brought the tip of the needle too medial to the sciatic nerve. Internal rotation of the leg facilitated passage of all needles inserted at the level of the lesser trochanter. We conclude that internal rotation of the leg may significantly facilitate needle insertion in the anterior approach to sciatic block. (Anesth Analg 2001;92:460 2)

he anterior approach to the sciatic nerve block has several advantages over the posterior (1) or lithotomy (2) approaches. With the anterior approach, the block can be performed with the patient in the supine position, the limb need not be flexed (3,4), and both sciatic and femoral blocks can be placed with the patient in the same position (5). In the anterior approach, the needle is inserted through the anteromedial thigh, inferior to the inguinal ligament, and advanced posteriorly to the sciatic nerve sciatic nerve that lies directly behind the femur (Fig. 1). Ideally, the needle passes just medial to the femur and contacts the sciatic nerve. However, the needle often encounters the femur before reaching the sciatic nerve. Although the classical description of the block suggests that the needle simply should be walked off the bone in the event of needle-femur contact, this maneuver often results in displacement of the needle tip too medially and thus away from the nerve (4). Because rotation of the femoral shaft could change the configuration of the insertion plane required to reach the
Accepted for publication October 24, 2000. Address correspondence and reprint requests to Admir Hadz ic , MD, PhD, Department of Anesthesiology, St. Lukes-Roosevelt Hospital Center, 1111 Amsterdam Avenue, New York, NY 10025. Address e-mail to ah149@columbia.edu.

sciatic nerve, the purpose of this study was to determine whether internal and external rotation of the leg would improve the accessibility of the sciatic nerve by using the anterior approach.

Methods
We studied ten lower extremities from five adult cadavers. The cadavers had been embalmed for anatomical purposes with a solution of phenol (13%) as the principal fixative and glycerin (28%) to retain water content. The cadavers were placed in the supine position with the lower extremity in the neutral position. The landmarks for the anterior approach to sciatic nerve block were drawn according to the classic description by Beck (3) (Fig. 1). Three lines were drawn: an inguinal ligament line between the anterio-superior iliac spine and the pubic tubercle (inguinal ligament), a transtrochanteric line that passed through the greater trochanter (identified by dissection) and parallel to the inguinal ligament, and a transecting line perpendicular to the inguinal ligament that passed through the intersection of the medial one third and the lateral two thirds of the inguinal ligament line. The transecting line was extended caudally, and the needle insertion
2001 by the International Anesthesia Research Society 0003-2999/01

460

Anesth Analg 2001;92:4602

ANESTH ANALG 2001;92:460 2

REGIONAL ANESTHESIA AND PAIN MEDICINE VLOKA ET AL. ANTERIOR SCIATIC NERVE BLOCK AND LEG ROTATION

461

Figure 2. The needle is shown inserted retrogradely (posterioranterior) through the sciatic nerve to demonstrate the insertion plane in the anterior approach to sciatic nerve block.

block was performed first at the level of the lesser trochanter. The needles were inserted in a retrograde direction (i.e., posterior to anterior) in a parasagittal plane to determine the part of the femur causing obstruction to the needle path. The insertions were first done with the feet forming a 90 angle to the horizontal plane, and then they were repeated with the feet rotated 45 internal and then 45 external in relation to the sagittal plane. The same experiment was then repeated but with insertion of the needles at a level 2 cm distal to the lesser trochanter. The number of needlebone contacts and successful needle passes were recorded at each leg rotation.

Figure 1. Anterior approach to sciatic nerve block: a diagram with landmarks. ASIS anterior superior iliac spine; PT pubic tubercle; PS pubic spine; GT greater trochanter; LT minor (lesser) trochanter.

Results
The majority of needles (80%) inserted anteriorposteriorly through the classic insertion point for sciatic nerve B contacted the lesser trochanter on the first attempt. Similarly, 80% of needles inserted through the center of the sciatic nerve in the posterior-anterior direction at the level of the lesser trochanter (with the foot forming a 90 angle to the horizontal plane) could not be advanced in the parasagittal plane toward the anterior thigh. This is because the bony prominence of the lesser trochanter intercepted the passage of the needle in both directions. Upon contacting the femur through Becks insertion point (3), medial redirection of the needle allowed it to pass by the lesser trochanter but brought the tip of the needle too medially to the sciatic nerve and closer to the femoral artery. Consequently, none of the medial reinsertions of the needle resulted in needle-sciatic nerve contact. The internal rotation of the leg by 45 facilitated passage of the needle inserted at the level of the lesser trochanter by opening the insertion plane, resulting in needle-nerve intersection in all 10 legs (100%). However, the external rotation resulted in contact with the

site was marked on the anterior thigh at the intersection of the transecting line and transtrochanteric lines. Needles (16-gauge, 15 cm long) were inserted at this site perpendicular to the horizontal plane until the needles either exited the posterior thigh or could not be advanced because of contact with the femur. The posterior thigh was then dissected to expose the sciatic nerve and to determine the number of needle-sciatic nerve and needle-bone contacts. After exposing the sciatic nerve, the effect of leg rotation on the needle plane required to reach the sciatic nerve through the anterior approach was studied. Needles (16-gauge, 15 cm long) were inserted through the center of the exposed sciatic nerve and advanced in the posterior-anterior direction in the parasagittal plane until the needle tips exited the marked insertion site on the anterior thigh or encountered the femur (Fig. 2). This simulation of the needle path by using the anterior approach to sciatic nerve

462

REGIONAL ANESTHESIA AND PAIN MEDICINE VLOKA ET AL. ANTERIOR SCIATIC NERVE BLOCK AND LEG ROTATION

ANESTH ANALG 2001;92:460 2

femur and prevented needle-nerve contact in all insertions (100%). In contrast to the insertion at the lesser trochanter, needles inserted at the level 2 cm inferior to the lesser trochanter encountered the femur in 40% of cases in the neutral position. At this level, the internal rotation of the leg by 45 had an entirely opposite effect and obstructed passage of the needle in all but one leg (90%), whereas external rotation facilitated needle passage and resulted in needle-nerve contact in all legs (100%).

Discussion
Using a cadaver model, we confirmed the clinical impression that in the anterior approach to sciatic nerve block, the femur often obstructs the path of the needle. However, walking off the femur, as the classic description of anterior approach to sciatic nerve block by Beck (3) suggests, results in the needle tip being too medial to and away from the sciatic nerve. Thus, progressive medial reinsertion of the needle with slight lateral angulation is usually required to reach the sciatic nerve behind the femur. Unfortunately, multiple attempts and medial reinsertion of the block needle increases the risk of puncturing the femoral nerve, femoral artery, or both. Because of this, the anterior approach may be less desirable in patients with peripheral vascular disease or those treated with anticoagulantsthe very patients in whom this block can have important advantages over neuraxial anesthesia. Our study demonstrates that during needle advancement with the anterior approach to sciatic nerve block (anterior to posterior in the sagittal plane), the path of the needle is often obstructed by the lesser trochanter. However, slight withdrawal of the needle from its position where it contacts the femur, followed by internal rotation of the leg and reinsertion of the needle in the same plane, facilitates the needle passage toward the sciatic nerve. This finding has been confirmed in a clinical study by Chelly and Delauney (4). When internal rotation fails to facilitate further advancement of the needle, external rotation should be attempted because the needle may erroneously be inserted below the lesser trochanter, in which case the

rotation has entirely the opposite effect. In this situation, the external rotation of the medial ridge of the trianglelike profile of the femoral shaft swings the ridge antero-laterally and opens the plane required for the block needle to pass medially to the femur. If these maneuvers fail, progressively more medial insertions should be attempted. The identification of the greater trochanter in this study was done by dissection, because its identification is difficult to perform on embalmed cadavers. Although this may somewhat limit the clinical applicability of these findings, we believe that our landmarks were sufficiently close to those obtained by palpation in patients. Furthermore, our preliminary clinical data are in agreement with the anatomical findings. In conclusion, if our results are applicable to clinical practice, the lesser trochanter often obstructs the path of the needle to the sciatic nerve in the anterior approach to sciatic nerve block. In this case, slight withdrawal followed by reinsertion of the needle in the same plane after internal rotation of the leg by 45 may significantly facilitate passage of the needle toward the sciatic nerve. When this is unsuccessful, our results further suggest that external rotation can also be helpful when the needle is erroneously inserted below the lesser trochanter, as may occur in patients with more obscure landmarks.
The authors thank Alan Santos, MD, for his editorial remarks and Drs. Chelly and Delauney for their kind contribution of anterior approach to sciatic nerve block diagram for this publication.

References
1. Labat G. Regional anesthesia: its technique and clinical applications. 2nd ed. Philadelphia: WB Saunders, 1928:4555. 2. Raj PP, Parks RI, Watson TD, Jenkins MT. A new single-position supine approach to sciatic-femoral nerve block. Anesth Analg 1975;54:489 93. 3. Beck GP. Anterior approach to sciatic nerve block. Anesthesiology 1963;24:222 4. 4. Chelly JE, Delauney L. A new anterior approach to the sciatic nerve block. Anesthesiology 1999;91:1655 60. 5. Vloka JD, Hadz ic A, Drobnik L, et al. Anatomical landmarks for femoral nerve block: a comparison of four needle insertion sites. Anesth Analg 1999;89:146770.

1999 American Society of Anesthesiologists, Inc. Volume 91(1) July 1999 pp 8-15

Effects of Perioperative Analgesic Technique on the Surgical Outcome and Duration of Rehabilitation after Major Knee Surgery [Clinical Investigations] Capdevila, Xavier MD, PhD; Barthele t, Yves MD; Biboulet, Philippe MD; Ryckwaert, Yves MD; Rubenovitch, Josh MD, BSc; d'Athis, Francoise MD (Capdevila, Barthelet, Biboulet, Ryckwaert, Rubenovitch) Assistant Professor. (d'Athis) Professor and Head, Department of Anesthesiology and Intensive Care Medicine. Received from the Department of Anesthesiology, Lapeyronie University Hospital, Montpellier, France. Submitted for publication April 30, 1998. Accepted for publication January 27, 1999. Support was provided solely from institutional and/or departmental sources. Presented in part of the annual meeting of the American Society of Anesthesiologists, San Diego, California, October 18-22, 1997. Address reprint requests to Dr. Capdevila: D.A.R. A, Hopital Lapeyronie, 371 Av du Doyen G. Giraud, 34295 Montpellier Cedex 5, France. Address electronic mail to: x-capdevila@chu-montpellier.fr Abstract Background: Continuous passive motion after major knee surgery optimizes the functional prognosis but causes severe pain. The authors tested the hypothesis that postoperative analgesic techniques influence surgical outcome and the duration of convalescence. Methods: Before standardized general anesthesia, 56 adult scheduled for major knee surgery were randomly assigned to one of three groups, each to receive a different postoperative analgesic technique for 72 h: continuous epidural infusion, continuous femoral block, or intravenous patient-controlled morphine (dose, 1 mg; lockout interval, 7 min; maximum dose, 30 mg/4 h). The first two techniques were performed using a solution of 1% lidocaine, 0.03 mg/ml morphine, and 2 [micro sign]g/ml clonidine administered at 0.1 ml [middle dot] kg-1 [middle dot] h-1 . Pain was assessed at rest and during continuous passive motion using a visual analog scale. The early postoperative maximal amplitude of knee flexion was measured during continuous passive motion at 24 h and 48 h and compared with the target levels prescribed by the surgeon. To evaluate functional outcome, the maximal amplitudes were measured again on postoperative day 5, at hospital discharge (day 7), and at 1- and 3-month follow-up examinations. When the patients left the surgical ward, they were admitted to a rehabilitation center, where their length of stay depended on prospectively determined discharge criteria. Results: The continuous epidural infusion and continuous femoral block groups showed significantly lower visual analog scale scores at rest and during continuous passive motion compared with the patient-controlled morphine group. The early postoperative knee mobilization levels in both continuous epidural infusion and continuous femoral block groups were significantly closer to the target levels prescribed by the surgeon than in the patient-controlled morphine group. On postoperative day 7, these values were 90 [degree sign] (60-100 [degree sign]) (median and 25th-27th percentiles) in the continuous epidural infusion group, 90 [degree sign] (60-100 [degree sign]) in the continuous femoral block group, and 80 [degree sign] (60-100 [degree sign]) in the patient-controlled morphine group (P < 0.05). The durations of stay in the rehabilitation center were significantly shorter: 37 days (range, 30-45 days) in the continuous epidural infusion group, 40 days (range, 31-60 days) in the continuous femoral block group, and 50 days (range, 30-80 days) in the patient-controlled morphine group (P < 0.05). Side effects were encountered more frequently in the continuous epidural infusion group. Conclusion: Regional analgesic techniques improve early rehabilitation after major knee surgery by effectively controlling pain during continuous passive motion, thereby hastening convalescence.

Key words: Local anesthetics; morphine; pain relief; physiotherapy. This article is accompanied by an Editorial View. Please see: Todd MM, Brown DL: Regional anesthesia and postoperative pain management: Long-term benefits from a short-term intervention. Anesthesiology 1999; 91:1-2. ALTHOUGH pain control occupies an unargued position in postoperative management, many questions concerning the role of analgesia on the postoperative outcome, beyond the fundamental humane aspect, remain to be resolved. Several authors have reported the importance of pain management in controlling postoperative complications in high-risk patient populations. [1-3] The beneficial effects of analgesia on functional rehabilitation and the duration of convalescence have been suggested repeatedly but demonstrated in only few publications. [4,5] Liu et al. [6] showed that epidural analgesia, associated with early ambulation and feeding, improved postoperative outcome after elective colon surgery. Kehlet [7] and Kehlet and Dahl [8] highlighted the importance of analgesia in optimizing postoperative rehabilitation. The authors insist on the need to develop techniques that allow early functional recuperation and emphasize the importance of integrating analgesia into multimodal rehabilitation programs. Many teams have included early mobilization using continuous passive motion (CPM) into their rehabilitation regimens after knee surgery. Although CPM improves the functional outcome, and as such may decrease the length of hospital stay, [9] it causes severe pain during mobilization. Few studies have explored the effects of postoperative analgesia on the functional prognosis of major knee surgery. [5,10,11] None of them have evaluated the benefits of postoperative continuous plexus or femoral blocks, and they could not show a decreased length of stay in patients who received epidural infusions. The aim of this study was to evaluate prospectively the influence of three postoperative analgesic techniques on the functional outcome and subsequent duration of hospitalization (hospital and rehabilitation center) after major knee surgery and use of CPM; i.e., continuous epidural infusion (CEI), continuous femoral block (CFB), and intravenous patient-controlled analgesia with morphine (PCA). Methods After our institutional review board give its approval and patients provided written informed consent, we enrolled 56 patients into this prospective study. All patients were classified as either American Society of Anesthesiologists physical status I or II, ranged in age from 18 to 75 yr, and were scheduled for total knee replacement or arthrolysis. Protocol One of three surgeons belonging to the same team and using the same techniques performed all of the operations. Before surgery, each patient was randomized to one of three postoperative analgesia groups: CEI, CFB, or PCA. All patients were premedicated with 0.5 mg oral alprazolam. The patients of the CEI and CFB groups were prepared for regional analgesia using 20-gauge catheters (Vygon, Les Ullis, France) placed under surgical aseptic conditions. For the CEA patients, the catheter was threaded 3 cm into the epidural space after medial puncture of the L2-L3 or L3-L4 vertebral interspaces with an 18gauge Tuohy needle (Braun, Melsungen, Germany). For the CFB patients, the puncture was performed using Winnie's landmarks with a 16-gauge, 80-mm nontraumatic needle (Krebbs; Pajunk, Ulm, Germany) linked to a neurostimulator (Stimuplex, Braun) by a sterile cable. The femoral nerve was localized by a required motor response, ascension of the patella, obtained at less than 0.5 mA. The catheter insertion length was 12-15 cm. Local anesthetics were not administered via the catheters in either group before the postoperative period. Intraoperative general anesthesia, standardized for the three study groups, was induced with 5 mg/kg intravenous thiopental, 1 [micro sign]g/kg sufentanil, and 0.5 mg/kg atracurium. All patients were intubated, and controlled ventilation was applied for the duration of surgery. Anesthesia was maintained using 60% nitrous oxide in oxygen, 0.75%-1.5% isoflurane end-tidal concentration, and 0.3 [micro sign]g/kg sufentanil given over 60 min, followed by a 0.15 [micro sign]g [middle dot] kg-1 [middle dot] h-1 continuous infusion, which was stopped 30 min before the end of surgery. Postoperative Analgesic Management and Discharge Criteria

Pain was evaluated during the study period using a visual analog scale (VAS) ranging from 0 mm (no pain) to 100 mm (worst imaginable pain). The postoperative analgesia protocol was initiated in the post-anesthesia care unit and continued in the surgical ward. The CFB patients received a 25-ml bolus of 2% lidocaine with 1/200,000 epinephrine and 2 mg morphine without preservatives via the femoral catheter. The resulting blockade was tested using the pinprick technique, assuring at least a femoral if not 3-in-1 blockade. The block was maintained by the continuous infusion of an analgesic solution, containing 1% lidocaine, 2 [micro sign]g/ml clonidine, and 0.03 mg/ml morphine administered at 0.1 ml [middle dot] kg-1 [middle dot] h-1 . If after 30 min pain control was considered insufficient (i.e., a VAS of 40 mm), a subcutaneous injection of morphine (0.1 mg/kg) was administered as rescue analgesia and repeated at 6-h intervals as required. The CEI patients received 2 mg morphine without preservatives and 5-ml doses of 2% lidocaine with 1/200,000 epinephrine, via the epidural catheter, until a T10 level was determined using the pinprick method. The epidural blockade was maintained by the continuous infusion of the same solution used to maintain the femoral blockade in the CFB group, administered at the same rate. If after 30 min pain control was considered insufficient (i.e., a VAS of 40), a subcutaneous injection of 0.1 mg/kg morphine was administered as rescue analgesia and repeated at 6-h intervals as required. The PCA patients received an initial intravenous infusion of morphine (2-mg doses at 5-min intervals) titrated manually until VAS scores of 30 mm were obtained. At this time, a PCA pump (Ivac, San Diego, CA) was connected, delivering 1-mg doses with a 7-min lockout period and a maximum dose of 30 mg in 4 h. The first pain evaluation under the influence of PCA morphine was performed 30 min later. If after 1 h pain control was considered insufficient (i.e., a VAS of 40 mm), the intermittent doses were increased to 1.5 mg and the patients were encouraged to use the PCA as often as possible. The patients stayed 12 h in the post-anesthesia care unit. During the 48 h after surgery, all patients received 2 g propacetamol and 100 mg ketoprofen, infused intravenously during 15 min at 8-h and 12-h intervals, respectively. All supplementary subcutaneous injections of morphine administered in the CFB and CEI groups were noted. On the morning of postoperative day 3, PCA, CFB, and CEI were discontinued and the catheters were removed. Pain was evaluated at rest and during early mobilization. The resting pain levels were determined and recorded 1 (H1 ), 6 (H6 ), 12 (H12 ), 24 (H24 ), and 48 (H48 ) h after the onset of analgesia. Early rehabilitation was initiated on the day after surgery (day 1) using a motorized variable amplitude splint (Kinetec, Tournes, France) and maintained during 10-12 h per day. In accordance with the surgical team, knee flexions of 40 [degree sign] and 50 [degree sign] were progressively attempted during 30 min on days 1 and 2, respectively. Pain during mobilization was evaluated by a physiotherapist during this 30-min onset period of CPM, at 24 h, and at 48 h. Excessive pain (i.e., a VAS of 60) was treated by decreasing the amplitude of flexion as necessary. If severe pain (i.e., a VAS of 80) was encountered despite the protocol's maximal analgesic levels and decreases in the amplitude of flexion, mobilization was deferred. All deferments required within the hour after the onset of mobilization were noted. The median maximal amplitude of flexion obtained at 24 h and 48 h were noted, allowing comparison with the target levels prescribed by the surgeon. All pain evaluations were associated with the surveillance and recording of possible side effects arising from the analgesic protocol. General effects included arterial hypotension (> 20% decrease in the preoperative mean blood pressure value), respiratory depression (respiratory rate, <or= to 8 breaths/min), sedation (0 = awake, 1 = sleepy but awakened by oral order, 2 = sleepy but awakened by nociceptive stimulation, 3 = not awakened), urinary retention (impossibilit y to urinate, requiring a urinary catheter to empty the bladder), nausea, vomiting, pruritus, and dysesthesia (paresthesia, numbness). Local complications included hematomas, catheter occlusions, kinks, or displacements. A member of the surgical team, blin ded to the postoperative analgesic technique, determined the maximal amplitude of knee flexion achieved on postoperative day 5 and on discharge from the surgical ward on day 7. All patients were admitted to a rehabilitation center when they were discharged from the surgical ward on day 7. A daily postoperative rehabilitation and assessment program was established for each patient based on four targets: joint mobility (including 2 h of CPM twice daily and manual mobilizations conducted by a physiotherapist), muscle force (quadriceps muscle force was evaluated daily using an isometric force dynamometer and trained using a weighted pulley system), motor function (rehabilitation was performed by having the patient walk an inclined plane, crouch, and climb steps), and absence of local complications (thrombophlebitis, inflammation). The length of stay in the rehabilitation center was determined by a blinded physiatrist. The objective criteria used

for discharge from the rehabilitation center included knee flexion of 110 [degree sign], knee extension of 0 [degree sign], lower limb flexion of 90 [degree sign] with 0 [degree sign] of knee extension, and the ability to walk an inclined plane without aid and to climb and descend 10 stairs. Statistical Analysis Data were analyzed using SAS version 6.11 software (SAS Institute, Cary, NC). The quantitative anthropometric, hemodynamic, and morphine consumption values were expressed as the mean +/- SD. Pain scores and knee mobilization values were expressed as medians (25th-75th percentiles), and the length of stay in the rehabilitation center was expressed as the median (and range). The repeated-measures aspect of this study was evaluated using analysis of variance. A Kruskall-Wallis test was used to compare the quantitative parameters of the three analgesic techniques at each evaluation. When a significant difference was encountered, the groups were compared two at a time, and Bonferroni correction was applied. Categorical data were compared using the chisquare or Fischer exact tests, as appropriate. A significance threshold of P < 0.05 was retained. Results Patient demographics, type of surgery, and duration of surgery (Table 1) were similar in all three patient groups: PCA (n = 19), CFB (n = 20), CEI (n = 17).

Table 1. Arthropometric Characteristics, Types, and Duration of Surgery

The resting VAS scores of the CEI group were significantly less than those of the PCA group at all test times and significantly less than those of the CFB group from H6 to H12 , but they were similar at H (1). The CFB group's VAS scores were significantly less than those of the PCA group at H1 and from H24 onward. During mobilization (Figure 1), the VAS of the CEI and CFB groups showed no significant differences during the study period. Both were significantly less than those of the PCA group.

Figure 1. Comparison of the visual analog scale values during continuous passive motion (CPM) of the three groups at (A) 24 h and (B) 48 h. The box represents the 25th-75th percentiles; the dark line is the median; the extended bars represent the 10th-90th percentiles, and the circles represent values outside this range. *P < 0.01 versus continuous femoral block (CFB) and continuous epidural infusion (CEI).

One subcutaneous morphine injection was required on day 1 by two CEI and six CFB patients. No significant differences were noted between the two groups' supplemental morphine consumption. Neither group required morphine supplements on day 2. The PCA group's mean morphine consumption was 36 +/- 13 mg and 31 +/- 15 mg on days 1 and 2, respectively. Side effects (Table 2) were most often noted in the CEI group, with a significantly elevated incidence of urinary retention, dysesthesia, and arterial hypotension (Table 3). A high incidence of sedation, defined as the need to call the patient by name to incite awakening, was noted in all groups in the post-anesthesia care unit.

Table 2. Principal Side Effects in the Three Analgesic Techniques

Table 3. Arterial Hypotension and Mean Arterial Blood Pressure Values in the Three Analgesic Groups

At 24 and 48 h, the CEI and CFB groups achieved the prescribed mobilization levels significantly more frequently and required deferred mobilization less frequently than did the PCA group (Table 4). Similarly, the maximal amplitude of knee flexion reached on day 5 and at discharge from the hospital was significantly greater in the CEI and CFB groups (Table 5). No significant differences were noted for the knee flexion values among the three groups at the 1- and 3-month follow-up examinations. The duration of stay in the rehabilitation center, needed to reach the target levels prescribed by the physiatrist, was significantly different between the regional analgesia and the PCA groups: CEI, 37 days (range, 30-45 days); CFB, 40 days (range, 31-60 days); PCA, 50 days (range, 30-80 days; P < 0.05).

Table 4. Quality of Early Rehabilitation ([degree sign])

Table 5. Functional Outcome: Knee Flexion ([degree sign]) at Day 5, upon Discharge from the Surgical Ward (Day 7), and at 1and 3- month Follow-ups Discussion After major knee surgery, analgesia provided by CEI or CFB is more effective during early motorized mobilization than by intravenous PCA. These regional analgesia techniques allow for more intense early rehabilitation and accelerate functional recuperation and shorten the total (hospital and rehabilitation center) duration of institutional stay. One aspect of the study design deserves comment. The pain evaluations during the postoperative days were not performed under blinded conditions because of the clinical setting of this study. Rigorous scientific methods would have required placing a femoral and epidural catheter and joining a PCA pump to the peripheral venous catheter in all patients. Because only the analgesic technique tested in each group would be used, evident ethical reasons restrained our application of this method. In contrast, blinded conditions were applied to the evaluations of functional outcome and the length of stay in the rehabilitation center. Analgesia Our results confirm the efficacy of the three tested analgesic techniques in controlling resting pain after knee surgery. [12-17] Of the three, PCA remains slightly less efficient. This difference could be caused, in part, by autoadministration with the classically maintained pain level tolerated by the patients. [18] In contrast, our findings concerning early rehabilitation using continuous motorized mobilization highlight the differences among the three techniques. Both continuous epidural and femoral blockades, optimized by their association with nonopioids in a multimodal analgesic protocol, provided more effective pain control during early mobilization, as evident by lower VAS scores. The efficacy of CFB during rehabilitation, although certainly of interest, has rarely been addressed in the literature. We found only a few contradictory studies that evaluated the postoperative analgesic effect of CFB during punctual mobilization. [12-14,17] [double dagger] In accordance with the findings of Hirst et al., [17] most of our patients reported pain in the region behind the knee. The sciatic nerve, not affected by CFB because of its sacral plexus origin, participates in the knee's innervation and certainly played a role in this pain. In contrast with the findings of Hirst et al., [17] our patients' pain was well controlled during mobilization. Our choice of analgesic regimen and the mobilization technique that we used may explain this difference. Although Hirst et al. [17] used bupivacaine alone to obtain blockade, we used a combination of lidocaine, clonidine, and morphine in the current study to obtain femoral blockade, which provided an additive if not synergistic effect to the resulting analgesia. Although the superior efficacy of CEI has been demonstrated in controlling dynamic pain when local anesthetics are associated with adjuvants, [18,19] their effects in plexus blocks are still being studied. [20-24] When standard doses are used, 1% lidocaine provides better motor blockade than bupivacaine 0.25% or 0.125%, more effectively avoiding quadriceps muscle spasm, which is cited as the cause and consequence of postoperative pain that hinders rehabilitation. [25] In addition, although as yet unconfirmed, regional analgesia would block the massive afferent nociceptive input thought to trigger increased excitability of the peripheral nociceptors and the dorsal horn neurons. Consequently, the increased reflex excitability leading to quadriceps muscle spasm, in response to even nonnociceptive input, would be controlled. [26] Furthermore, muscle relaxation was enhanced by replacing the punctual manipulations performed by physiotherapists with motorized CPM, a more gentle and progressive mobilization. [9,27] As such, the anticipated increase in pain often incriminated in the failure of rehabilitation using punctual mobilization was avoided. [15,28] [double dagger] In accordance with the literature [15,28,29] [double dagger] and despite the use of CPM, the pain scores of the PCA group were significantly greater than those of the groups treated by regional analgesia. The greater analgesic efficacy noted in the regional analgesia groups was associated with greater knee flexion values (i.e., better quality rehabilitation).

Although not statistically significant, lower VAS scores were expressed by patients in the CEI group than by those treated with CFB, suggesting a greater analgesic effect provided by CEI. Although epidural analgesia remains the reference technique, CFB has been shown to be safe and to cause only few minor adverse effects, [12,13] thus providing the best balance of analgesia and side effects. In contrast to CEI, femoral blocks cause fewer episodes of low blood pressure and urinary retention; the latter, difficult to accept in the context of functional surgery, also increases the risk of infection arising from the use of urinary catheters. Rehabilitation The novelty of this study lies in the 48-h postoperative analysis of the quality of rehabilitation, functional outcome, and length of stay in the rehabilitation center. In contrast to PCA, the analgesic quality of the regional blocks allowed patients to consistently achieve the mobilization levels targeted by the surgeons. The optimized rehabilitation led to earlier functional recuperation reflected in the knee flexion values noted on day 5 and at discharge from the hospital, which were greater than those of the PCA group. Similar findings have been reported in the literature. Syngelyn and Gouverneur [12] noted significantly better knee flexion from days 1 to 10 after total knee replacement surgery using CFB and CEI rather than PCA. As in the current study, a 10 [degree sign] advantage was noted in the regional analgesia groups on day 10. Comparing intramuscular morphine and CEI with local anesthetics during CPM after the same surgery, Pettine and Wedel [10] noted mean knee flexion values of 86 [degree sign] and 93 [degree sign], respectively, at discharge from the hospital 10-12 days after surgery. Moiniche et al. [11] performed the same comparison after total hip or knee replacement. When comparing the postoperative outcomes following unilateral primary total knee replacement under either epidural or general anesthesia, Williams-Russo et al. [4] reported that epidural anesthesia was associated with more rapid achievement of postoperative goals. All rehabilitative milestones were reached earlier, with patients climbing stairs significantly earlier. However, none of the authors could show that accelerated functional recuperation resulted in decreased hospital stays. Pettine and Wedel [10] programmed a minimal hospital stay of 10 days, masking the prolonged rehabilitation caused by persistent pain in the intramuscular morphine group. Moiniche et al. [11] did not include early intensive rehabilitation in the study protocol, leaving the analgesic advantage of epidural analgesia unexploited. Because the hospital stay in our study was limited to 7 days, analysis was centered on the duration of stay in the rehabilitation center. The knee flexion values noted in the regional analgesia groups at discharge from the hospital, increased by 10 [degree sign] compared with the PCA group, significantly shortened the rehabilitation period required to reach the objective functional discharge criteria used by the physiatrist. In accordance with Pettine and Wedel, [10] follow-up examinations at 1 and 3 months showed no functional differences among the groups, as reflected by knee flexion values. Although resolution of pain during the follow-up period allowed the three group's functional results to be homogenized, early intensive rehabilitation, facilitated by regional analgesia, accelerated functional recuperation. A multimodal recovery program including regional analgesic techniques and CPM should be privileged after major knee surgery. The augmented analgesic effect of the regional techniques on pain during early mobilization permitted the rehabilitative advantages of CPM to be maximized. When compared with the use of PCA, functional recuperation was accelerated, and the overall hospital stay was shortened. Similar findings were reported after colon surgery, [6] leading to Kehlet's [7] recommendation of regional analgesia techniques in 1994. The choice of regional techniques should be based on a careful evaluation of the benefits and risks. The routine use of CEI after knee surgery should be limited because shortening the global hospital stay justifie s neither the discomfort of side effects nor the danger of potentially serious complications. In contrast, CFB seems to have all the qualities necessary to become the primary choice for regional analgesia after major knee surgery. [double dagger] Hord AH, Roberson JR, Thompson WF, Cohen DE: Evaluation of continuous femoral nerve analgesia after primary total knee arthroplasty (abstract). Anesth Analg 1990; 70:S164. REFERENCES 1. Beattie WS, Buckley DN: Epidural morphine reduces the risk of postoperative myocardial ischaemia in patients with cardiac risk factors. Can J Anaesth 1993; 40:532-41 [Medline Link] [Context Link] 2. Mangano D, Siliciano D, Hollenberg M, Study of Perioperative Ischemia Research Group: Postoperative myocardial ischemia: Therapeutic trials using intensive analgesia following surgery. Anesthesiology 1992; 76:342-53 [Context Link] 3. Yeager MP, Glass DD, Neff RK, Brinck-Johsen T: Epidural anesthesia and analgesia in high-risk surgical patients. Anesthesiology 1987; 66:729-36 [Medline Link] [Context Link]

4. Williams-Russo P, Sharrock NE, Haas SB, Insall J, Windsor RE, Laskin RS, Ranawat CS, GO G, Ganz SB: Randomized trial of epidural versus general anesthesia: Outcomes after primary total knee replacement. Clin Orthop 1996; 331:199-208 [Context Link] 5. Mahoney OM, Noble PC, Davidson J, Tullos HS: The effect of continuous epidural analgesia on postoperative pain, rehabilitation and duration of hospitalization in total knee arthroplasty. Clin Orthop Rel Res 1990; 260:30-7 [Medline Link] [Context Link] 6. Liu SS, Carpenter RL, Mackey DC, Thirlby RC, Rupp SM, Shine TSJ, Feinglass NG, Merzger PP, Fulmer JT, Smith SL: Effects of perioperative analgesic technique on rate of recovery after colon surgery. Anesthesiology 1995; 83:757-65 [Fulltext Link] [Medline Link] [Context Link] 7. Kehlet H: Postoperative pain relief. What is the issue ? Br J Anaesth 1994; 72:375-8 [Medline Link] [Context Link] 8. Kehlet H, Dahl JB: The value of 'multimodal' or 'balanced analgesia' in postoperative pain treatment. Anesth Analg 1993; 77:1048-56 [Medline Link] [Context Link] 9. Colwell CW, Morris BA: The influence of continuous passive motion on the results of total knee arthroplasty. Clin Orthop 1992; 276:225-8 [Medline Link] [Context Link] 10. Pettine KA, Wedel DJ: The use of epidural bupivacaine following total knee arthroplasty. Orthop Rev 1989; 18:894-901 [Medline Link] [Context Link] 11. Moiniche S, Hjortso NC, Hansen BL, Dahl JB, Rosenberg J, Gebuhr P, Kehlet H: The effect of balanced analgesia on early convalescence after major orthopaedic surgery. Acta Anaesthesiol Scand 1994; 38:328-35 [Medline Link] [Context Link] 12. Syngelyn FJ, Gouverneur JM: Influence of the postoperative analgesic technique on knee mobilization after total knee replacement (abstract). Anesthesiology 1997; 87:A775 [Context Link] 13. Anker-Moller E, Spansberg N, Dahl JB, Christensen EF, Schultz P, Carlson P: Continuous blockade of the lumbar plexus after knee surgery: A comparison of the plasma concentrations and analgesic effect of bupivacaine 0,250% and 0,125%. Acta Anesthesiol Scand 1990; 34:468-72 [Context Link] 14. Schultz P, Anker-Moller E, Dahl JB, Christensen EF, Spansberg N, Fauno P: Postoperative pain treatment after open knee surgery: Continuous lumbar plexus block with bupivacaine versus epidural morphine. Reg Anesth Pain Med 1991; 16:34-7 [Medline Link] [Context Link] 15. Loper KA, Ready LB: Epidural morphine after anterior cruciate ligament repair: A comparison with patient-controlled intravenous morphine. Anesth Analg 1989; 68:350-2 [Medline Link] [CINAHL Link] [Context Link] 16. Weller R, Rosenblum M, Conard P, Gross JB: Comparison of epidural and patient-controlled intravenous morphine following joint replacement surgery. Can J Anaesth 1991; 38:582-86 [Medline Link] [Context Link] 17. Hirst GC, Lang SA, Dust WN, Cassidy D, Yip RW: Femoral nerve block: Single injection versus continuous infusion for total knee arthroplasty. Reg Anesth Pain Med 1996; 21:292-7 [Medline Link] [Context Link] 18. Ferrante M, Orav EJ, Rocco AG, Gallo J: A statistical model for pain in patients-controlled analgesia and conventional intramuscular opioid regimens. Anesth Analg 1988; 67:457-61 [Medline Link] [Context Link] 19. Dahl JB, Rosenberg J, Hamsen BL, Hyortso NC, Kehlet H: Differential analgesic effect of low dose epidural morphine and morphine-bupivacaine at rest and during mobilization after major abdominal surgery. Anesth Analg 1992; 74:362-5 [Medline Link] [Context Link] 20. Gaumann D, Forster A, Griessen M, Habre W, Poinsot I, Della Santa D: Comparison between clonidine and epinephrine admixture to lidocaine in brachial plexus block. Anesth Analg 1992; 75:69-74 [Medline Link] [Context Link]

21. Viel E, Eledjam JJ, de la Coussaye JE, d'Athis F: Brachial plexus block with opioids for postoperative pain relief: Comparison between morphine and buprenorphine. Reg Anesth Pain Med 1989; 14:274-8 [Context Link] 22. Flory N, Van-Gessel E, Donald F: Does the addition of morphine to brachial plexus block improve analgesia after shoulder surgery? Br J Anaesth 1995; 75:23-6 [Medline Link] [Context Link] 23. Singelyn FJ, Gouverneur JM, Robert A: A minimum dose of clonidine added to mepivacaine prolongs the duration of anesthesia and analgesia after axillary brachial plexus block. Anesth Analg 1996; 83:1046-50 [Fulltext Link] [Medline Link] [Context Link] 24. Bernard JM, Macaire P: Dose-range effects of clonidine added to lidocaine for brachial plexus block. Anesthesiology 1997; 87:277-84 [Fulltext Link] [Medline Link] [Context Link] 25. Bonica JJ: Current status of postoperative pain therapy, Current Topics in Pain Research and Therapy. Edited by R. Dubnet. Tokyo, Yokota, 1983, pp 169-89 [Context Link] 26. Woolf C, Wall P: The brief and the prolonged facilitatory effects of unmyelinated afferent input on the rat spinal cord are independently influenced by peripheral nerve injury. Neuroscience 1986; 17:1199-205 [Medline Link] [Context Link] 27. Boitard J, Reboul C, Vidal J, Buscayret C: [Continuous passive motion: Preventive and curative use in knee diseases], Actualites en Reeducation Fonctionnelle et Readaptation. Edited by L Simon. Paris, Masson, 1984, pp 214-8 [Context Link] 28. Egan KJ, Ready LB: Patient satisfaction with intravenous PCA or epidural morphine. Can J Anaesth 1994; 41:6 -11 [Medline Link] [Context Link] 29. Raj PP, Knarr DC, Vigdorth H, Denson DD, Pither CE, Hartrick CT, Hopson CN, Edstrom HH: Comparison of continuous epidural infusion of a local anesthetic and administration of systemic narcotics in the management of pain after total knee replacement surgery. Anesth Analg 1987; 66:401-6 [Medline Link] [Context Link]

Accession Number: 00000542-199907000-00006 Copyright (c) 2000-2001 Ovid Technologies, Inc. Version: rel4.3.0, SourceID: 1.5031.1.149

Letters to the Editor


Visual Analog Scale Scores for Labor Pain
To the Editor: I read with interest the article by Ludington and Dexter (1) describing the use of a total pain score for analysis of analgesic modalities in labor. The concept of a total pain score for labor should not be interpreted by readers as a valid or even a useful measure of analgesia. Readers should be aware that the total pain score has never been used in contemporary obstetric analgesia trials, has never been validated for this purpose, and may represent a remarkable oversimplification of the labor process. This concept should not have been proposed as a valid technique in a medical intelligence article, but rather presented as a hypothesis that needs further study and verification. To ascribe one single total pain score for an entire labor would present extreme difficulties to researchers seeking to assess subtle but important differences among drugs and/or techniques. Obtaining total pain scores by multiplying retrospective visual analog scale (VAS) measurements by duration of labor has no basis in either physiology or any commonly used current research paradigm. Can you realistically compare squeezing your head tightly in a vice for 10 min to squeezing it lightly for 8 h and assume that the experience is the same because the time-weighted average VAS scores were similar? Moreover, the authors claim that VAS scores are true ratios. They claim that a score of 6 is twice as severe as a score of 3, and that the magnitude of difference between a score of 2 and 3 is the same as between 7 and 8, for example. Although this may be true in a purely mathematical sense only (and the authors should have acknowledged this limitation), it is hardly true in reality; in fact, it even defies common sense. Certainly no one would claim that a baby with an Apgar score of 6 is exactly twice as healthy as one with a score of 3. These are all ordinal, not interval, data. Parametric statistics may be used with robustness to analyze ordinal data, as shown by Dexter and Chestnut (2). However, the underlying numbers must still be recognized as ordered, not interval. Mathematical analysis notwithstanding, interpreting VAS data during labor as pure linear continuous data simply makes no physiological sense and represents a gross oversimplification of the complex factors associated with measurement of labor pain. To ascribe one single total pain score to an entire labor is equally inappropriate. William Camann, MD
Department of Anesthesia Brigham and Womens Hospital Boston, MA 02115

The visual analog scale is a vertical or horizontal line, usually 10 cm in length, with descriptive words such as no pain and worst pain imaginable written at the end points. The patient marks the line at the point that they consider represents the amount of pain they are experiencing at that moment. The psychometric basis for the visual analog scale relates to power functions relating (i) the subjective magnitude of pain to the true magnitude of the pain stimulus and (ii) the subjective magnitude of linear distance to the true distance along a line. In contrast, the Apgar score assigns numerical values to nonoverlapping clinical categories. The two are different from a psychometric perspective. Franklin Dexter, MD, PhD
Department of Anesthesia University of Iowa Iowa City, IA 52242

Repeated Use of the Cuffed Oropharyngeal Airway in an Infant for Radiation Therapy
To the Editor: We recently encountered a 6-mo-old girl diagnosed with rhabdomyosarcoma of the left mandible, requiring general anesthesia to receive twice-daily radiation therapy for 6 wk. This required an atraumatic airway management technique that provided an effective airway during a propofol general anesthetic. Because of anticipated irritationif not mucosal friabilitysecondary to the radiation dose, we sought to avoid using endotracheal intubation of the laryngeal mask airway. To address our concerns, and with the parents consent, a small cuffed oropharyngeal airway (COPA; 6 cm) custom-made with a larger bite block (Figure 1 inset) to provide a consistent open jaw angle for consistent, precise delivery of radiation dose, was provided (Mallinckrodt Medical, Inc., St. Louis, MO). The COPA also allowed administration of oxygen and monitoring of end-tidal CO2. Although we encountered no significant complications during the 49 anesthetics performed using this technique, it was necessary, at times, to reposition the device or to use a tongue depressor to ensure proper positioning of the COPA cuff. These manipulations were

References
1. Ludington E, Dexter F. Statistical analysis of total labor pain using the visual analog scale and application to studies of analgesic effectiveness during childbirth. Anesth Analg 1998;87:7237. 2. Dexter F, Chestnut DH. Analysis of statistical tests to compare visual analogue scale measurements among groups. Anesthesiology 1995;82:896 902.

In Response: I agree that to ascribe one single total pain score for an entire labor would present. . .difficulties to researchers seeking to assess subtle. . .differences among drugs and/or techniques. However, our article considered not the evaluation of new drugs or therapies, but rather statistical methodologies to measure their cost-effectiveness as labor analgesics. Consideration of both the magnitude and duration of pain (total pain) relief provided by a drug or therapy is necessary to perform a cost-effectiveness analysis. Additional studies are required to assess patients relative valuations for different magnitudes and durations of pain.
1999 by the International Anesthesia Research Society

Figure 1. A custom-made cuffed oropharyngeal airway used in an infant.


Anesth Analg 1999;88:14219

1421

1426

LETTERS TO THE EDITOR

ANESTH ANALG 1999;88:14219

lifting of the epiglottis by MMB-AE and exposure of the glottis were facilitated. Spyros D. Mentzelopoulos, MD, DEAA(P1) Marina V. Tsitsika, MD Evangelia A. Karamichali, MD, PhD
Department of Anesthesia Evangelismos General Hospital Athens, Greece

Assessment of Renal Effects of Sevoflurane in Elderly Patients Using Urinary Markers


To the Editor: Although sevoflurane anesthesia has been shown not to cause nephrotoxicity in adults with normal renal function, its effect in patients with impaired renal function or in the elderly has not been thoroughly investigated. Recently, urinary albumin has gained attention as a marker of postanesthetic renal injury in surgical patients (1), but there is no agreement on how the results of studies should be interpreted. At our institution, we have investigated the effects of inhaled anesthetics on renal function in elderly patients. We present the data of our study, in which urinary albumin, 1 microglobulin (MG), 2 MG, and N-acetyl-d-glucosaminidase (NAG) were used as markers of renal injury in elderly patients anesthetized with sevoflurane. Thirteen patients aged 70 yr undergoing gastrectomy were randomly assigned to receive either sevoflurane anesthesia (n 7; mean age 77.6 yr) or isoflurane anesthesia (n 6; mean age 78.5 yr). We used routine anesthetic techniques for our institution in this study, i.e., local epidural anesthesia combined with inhaled anesthesia (3 L/min air, 2 L/min oxygen, and either sevoflurane or isoflurane). A urine sample was collected via a catheter before anesthesia, and cumulative urine samples were then collected during and after anesthesia (immediately before surgery; 2 h after the start of surgery; at the end of surgery; and 3 h and Days 1, 3, and 7 after anesthesia). There were no significant differences in patient demographics. The mean minimum alveolar anesthetic concentration-hour was 5.1 for the sevoflurane group and 3.7 for the isoflurane group. The mean urinary albumin excretion was 65.0 mg/g creatinine (gCr) and 44.4 mg/gCr before anesthesia, which significantly increased to 147.9 mg/gCr and 196.9 mg/gCr 2 h after the start of surgery, sustained similar values 3 h after anesthesia, and returned to close to the preanesthetic values on Postanesthesia Day 1 in the sevoflurane group and the isoflurane group, respectively. No significant difference was found between the two groups, and the increase in urinary albumin excretion indicated that transient glomerular injury occurred with both anesthetics. The mean urinary 1 and 2 MG levels were 9.3 mg/gCr and 0.81 mg/gCr in the sevoflurane group and 7.4 mg/gCr and 0.70 mg/gCr in the isoflurane group before anesthesia, showed a gradual but significant increase in both groups during anesthesia, and reached 31.4 mg/gCr and 6.20 mg/gCr in the sevoflurane group and 44.1 mg/gCr and 10.9 mg/gCr in the isoflurane group by 3 h after anesthesia. The values of 1 and 2 MG then temporally dropped on Postanesthesia Day 1, but they increased again on Day 3 in both groups and decreased again on Day 7 (cause unknown). The mean urinary NAG also began to increase during anesthesia to significantly higher levels than the preanesthetic values (18.9 mg/gCr in the sevoflurane group and 16.9 mg/gCr in the isoflurane group), reached a peak 3 h after anesthesia (40.6 mg/gCr in the sevoflurane group and 35.7 mg/gCr in the isoflurane group), but returned to almost the preanesthetic values on Postanesthesia Day 1. The changes in the urinary enzyme levels indicated the presence of transient renal tubular injury in both groups. In conclusion, using urinary markers, we found that sevoflurane and isoflurane anesthesia in combination with epidural anesthesia resulted in similar degrees of mild, transient, glomerular, and tubular functional impairment in elderly surgical patients undergoing gastrectomy. Kokichi Hase, MD Kazuko Meguro, MD Takako Nakamura, MD
Department of Anesthesiology Tokyo Metropolitan Geriatric Hospital Tokyo 173-0015, Japan

References
1. Benumof JL, Cooper SD. Quantitative improvement in laryngoscopic view by optimal external laryngeal manipulation. J Clin Anesth 1996;8:136 40. 2. Mallampati SR, Gatt SP, Gugino LD, et al. A clinical sign to predict difficult tracheal intubation: a prospective study. Can J Anaesth 1985;32:429 34. 3. Samsoon GLT, Young JRB. Difficult tracheal intubation: a retrospective study. Anaesthesia 1987;42:48790.

Midfemoral Block: A New Lateral Approach to the Sciatic Nerve


To the Editor: A new lateral approach to the sciatic nerve is described as the midfemoral block. The surface landmarks are the palpable greater trochanter (GT), the great axis of the femur palpated. A line is drawn from the posterior margin of greater trochanter toward the knee, parallel to the femur. The puncture site is situated on this line at the middle of the thigh. At this point, in contrast to the popliteal fossa, the sciatic nerve is not yet divided, and is reached at 3 8 cm of depth. The procedure is conveniently performed while the patient is in the supine position with the limb raised on a pillow. The operator places one hand on the limb to move to zero rotation and exposes the line. With the other hand, the operator inserts a stimulating needle perpendicularly to the skin (Figure 1). The needle is advanced toward the sciatic nerve until evoking foot movements. The mean duration of complete sciatic blockade was 14 h after the administration of 20 30 mL of 0.5% adrenalinated bupivacaine. The midfemoral sciatic was combined with femoral block in 50 patients after total knee replacements and in 10 for foot surgeries. The combined blocks were performed with one pack of Plexus Mini-set (Pajunk, Geisingen, Germany). C. Pham Dang, MD
Service dAnesthe sie-Re animation Chirurgicale Ho tel-Dieu 44093 Nantes cedex 1, France

Figure 1. The midfemoral sciatic block is proceeding with the stimulating needle inserted on the line, perpendicularly to the skin. The line was drawn from the posterior margin of the greater trochanter (GT), down to the knee, parallel to the great axis of the femur. Note the continuous femoral block (CFB) previously achieved and dressed.

Comparison of Tw

nesthetic

echniques

In 1973. Winnie et ;il. described a technique : lr blocking the femoral. lateral femoral c u t a n e o u s a:3 the obturator nerves \kith one injection (perivascu.:r approach to the psoas compartment block. or Sir-i block). The uscfulnesx of the 3-in-l block for xthr Iscopic knee surgery was reported by Pate1 et al.

MATERIALS AND METHODS From January 1989 IG August 1989, 280 inpatients who were admitted to hospital for knee arthroscop> were included in this prospective study. The study entailed alternate-day assignment to treatment groups. so that 100 patients were randomly allocated to Group i to undergo Sin- 1 block and 180 patients were randomly allocated to Group 2 to have f femoral nax5

V. BONICALZI

ANIJ M. GALLINO

FIG 1. S?nsttivr innervation of the anterior knee. ( 1) Lateral cut;lneous ner, 2. (2) femoral nerve. I ZA) saphenous nerve ibranch of the fernor; nerve), (3) obturator nerve.

block alone. All the patients were between 15 and 79 years 010 2nd ASA category I or II; informed consent was obtk:ned from all patients during a preoperative examinzon. The details of the patients studied are provided in Table 1. All the patients received 0.5 mg intramu<<ular of Atropine and 5 to 10 mg of Diazepam orally (&pending on body weight) as premeditation. 15 to 6(1 .ninutes before estimated time of operation. All the blocks were performed by four trained anesthesiologsts with considerable practical experience in lower s~rremity blocks for orthopaedic interventions. The 3-ic-: blocks (Group I) were performed with the patients ..,ing supine. The inguinal area had been previously -7aved. A line connecting the pubic tubercle and the ;rrerior superior iliac spine was drawn. identifying the Lnguinal ligament. The femoral artery was

marked and a sterile held was prepared. A 72.gauge 5-cm short-bevel needle ~3s advanced lateral to the artery in a cephalad direction until paresthesia was elicited in the distribution ofthe femoral nerve. without use of nerve stimulator. The needle was held immobile while distal pressure was applied digitally to the femoral sheath (Fig 7). A total of 20 mL of local anesthetic solution (0.5% bupivacaine without epinephrine) wai injected incrementally after negative aspiration. If pain was present at the latera! hkin incision, the surgeon applied a supplementary intradermal and subcutaneous local anesthesia in the area of the lateral patellar portal by injecting 5 to 10 mL of0.5% bupivacaine. Care was taken to avoid an inadvertent intra-articular injection of the anesthetic solution. To perform the femoral nerve block (Group 2). the patient was placed in the supine position: the inguinai ligament and the femoral artery were marked as described above. A skin weal was raised I-cm lateral to the femoral artery, just below the inzuinal ligament, and a 7-Z-gauge G-cm short-bevel needle was advanced lateral to the line marking the femoral artery. When

REGIONAL BLOCKS FOR KNEE ARTHROSCOPY the needle reached the depth of the artery. a pulsation of the huh was visible. The needle was advanced until paresthesia was obtained. When a paresthesia was elicited, 10 mL of 0.5% hupivacaine without epinephrine solution :vas injected at that site. The patients who underwent the femoral block alone (Group 2) received a supplemental local intradermal and subcutaneous injection with 8 to 12 mL of 0.5% bupivacaine without epinephrine by the surgeon at the site of the lateral knee incision, to complete the procedure. The anesthesioiogicai procedures were carried out in an appropriate room. near the operating room, and the arthroxqic oFration began 25 to 30 minutes after the ejection (~;the local anesthetic. The tourniquet applied to the thisi was inflated when a pinprick test showed t h a t ;1 fUl1~ cutaneous anesthesia in the femoral nerve distribution had been established. In this way, the time of onset Q: Jnesthesia was also determined. The pressure in the :ourniquet was the patients arterial pressure plus 50 mx Hg. The tourniquet was routinely used in all paticni;. without pressure variations, Duration of ;:?Hation iid not exceed 70 minutes in any case. The !lrgical \z:lle field was then prepared so that the prozociure begin 5 to IO minutes after the tourniquet intlation. In GroU? I. 72 meniscectomies, 3 chondrectomies. 2 partial ~;novcctomies, I loose-body removal, I7 plicectomiex. 2nd 5 lateral releases were carried out. In Grou: 7. I38 meniscectomies, 2 chondrectomies. S partial \iri)vcctomies, 3 loose-body removals, I I pli :ctomie\. .:nd I3 lateral releases were carried out (Ta.i 7). If -;uicty and discomfort were present at any stage of If: procedure. patients received intravenous Ilunitraz.er:m. 0.0 I mg/kg, with intravenous fentanyl, 50 to 100 _ 2. il the patient complained of pain. The inc: ::ncc of pain correlated to operative maneuvres dtrrin: .\rthroscopy. the relationship between the type of ar.rroccopic operation. and the SLICC~SS 01 re(6onal ane -nesia: discomtort or pain (aching or burn: -( \insat: 7 in the distal extremity) due to the tourni-I:[. pel-k: ~-2ncc 01 cutaneous anesthesia to pinprick test and p -ioperative analgesia (assessed hy direct questionin; .is well iis the side effects of the blocks Ulcl the su:::-ons opinions on 1nu~cle relaxation were recorded i: - 111 patients. Age, weight, sex. drug closes. and durat1~ - of anesthesia and analgesia were cxanined bv SIL ::nts unpaired t-test. A11 results were anal!wti &r \..-!\tical differences between groups using 4 \~I .z.\[. P c .(X5 was I-egardcd as significant.
RESULTS

209

Group I Menlscectomies
Chondrecromies

cl-oup 2
S 23 14X Tn T 49 A I s2TnT I As.3TnT l Al s3 l-n T l AS 5 IlTnT 3 A -s 3 I3 Tn T 1 A-

P Value
c.01 .03 .8 NS .Ol .8 NS .8 NS .6 NS .8 NS .6 NS

s71TnT 13 A 3 s.?TnT.A s2 I.11 T-

tarr~ai

sinovcctornIc:s

Loose-body rcmov31

3c r
Xurator i of rhc

,A -sI Tn T., -s-I? 'l-n r 3

l'llctxlorllle

Lateral rdlexcs

Tll

.A I s'r I .A I

~\bbrevlations: Tn. iota/ Ilumber of patient>: S, p;lin from surgical li~;lneil\el-s-scdaIlon rcquesred: T, pain or discomfort from tourniquet--\ctlalton ~requesteci: A, -wner:il arwthesra [requested.

<aLlge o the I was thout lobile &nothetic ) was . pain geon


ltX>LlS m-td : w a s 311 o f

thetic dosage and the time of onset of anesthesia were significantly different between Group 1 and Group 2, who needed a smaller dose of anesthetic and had a more rapid onset of anesthesia (Table 3). Mean collective operative time was 41.5 It 13.2 ininutes (40.1 t 12.6 minutes in Group I and J2.2 f 13.5 minutes in Group 2). There was no statistically significant difference in mean operative times between the ? groups
o f putients.

1. t h e

uinal i Lie.:I1 to lent.


IlCld Jhen

The pal:: -:\ in both groups wcrc similar iu age. weight, ant ,5\; di\tributic)n (Table I ). The local anes-

in Group I. 75 patients experienced no pain resulting from the surgical muneuvres during the whole procedurc: 30 patients experienced pain on lateral knee incision and had a supplemental local anesthesia. Seventeen Of these 95 patients needed intravenous llunitrstepam and/or fentanyl because of the discomfort cxsed by the tourniquet. Pain or discomfort prevented the remaining 5 patients from tolerating the surgical maneuvres (rotatory and \.arus-valgus stress manoeuvrcs) or the tourniquet, and II general anesthesia K as needed to complete the operation (Table 4). In Group 2. 88 patients were pain free for the entire procedure: 3 1 patients needed intravenous ilunitratepam and/or fentanyl for the pain caused by the surgical maneuvres ix described above); 59 patient5 needed Ilunitratepam and/or fentanyl for discomfort or pain caused by the tourniquet: in 2 patients. a general anesthesia wx need4 (Table 4). No patients in Group 1 or 2 had any sensation of pain in the medial aspect of
the kntx.

210

V. B0NlCALZI AN13 M. GALLINO

T,\BLE 3. Meun Opercrtive Time, Dose of Local Anesthetic and Duratiotl (~Ane.sthesin/Analgesic
Group I Mean operative time Dose of local aneschzcic (mL) Onset or^ anesthesia (min) Duration of anesthesia (min) Duration of postop5rxive anal,o&a (min) 40.1 + 12.6 21.3 i: 3.5 28.7 _c 4.3 248 2 71 48.5 -t 96 Group 2 42.2 t 13.5 18.1 t 3.9 23.6 ir 5.8 823 F 67 382 2 103 P Value .2 NS c.01 <.Ol .I NS <.Ol

NOTE. Mean i- standard deviation.

No relation was found between the rate of success of regional anesthesia and the type of operation that was performed (Table 2). The surgeon was more satisfied with the degree of muscle relaxation with the 3in- 1 block than with femoral block alone, as the manipulation ,>,f knee joint was more difficult in the patients of Group 2. The duration of anesthesia was comparable for both groups. but the period of postoperative analgesia was significsntly longer for patients of the Group I (Table 3). No side effects were recorded in either group. DISCUSSION In th:; study. the number of patients in Group 2 is nearly :-.vice the number of patients in Group 1. We became iware of this abnormality during the statistical procesi:ng of the data. The abnormality is due to the fact th;: the study entailed alternate-day assignment to treatment groups and this may have introduced an unsuspf:ted bias. However. we believe that the study sample, ,vere obtained by a process that reasonably simiila1s, random selection from a larger population and tht: this manner of selecting patients for study does ni : jignitioantly distort the facts. Moreover we know trIt the calculation ritual for the test of signiticance cl.: be applied to any set of numbers. provided that the \rudy is not prone to strong effects of bias and uncontr led confounding variables.* For this reason we dec.:?d to perform the tests of significance with wr CllttL 'ILIT data show no st;ftikzaI tiifference in age, weight. _,d sex distribution between patients in either rroup, L : Lhtz subjects were AS!4 class I and II patients 2 I r~asox - Iy healthy patients). w e i-..e routinely pretnedicated our patients with atropinc end diazepam. A&opine was used in view of

the possibility of an inadvertent intravascular injection or a massive reassorbtion of bupivacaine (since a hi& dose of bupivacaine was injected near a great vessel) and of the risk of sinus bradycardia and sinus arrest induced by high blood levels of bupivacaine. BupivaCaine is proportionally more cardiotoxic than other l ocal anesthetics and the cardiac resuscitation following bupivacaine-induced toxicity is possible if massive doses of epinephrine and atropine are used. Diazepam was used to reduce anxiety, but it is also effective in controlling local anesthetic-induced seizures. It is known that regional anesthesia (both centroneuraxis blocks and peripheral nerve blocks) may be used for knee arthroscopy, provided that prolonged tourniquet times can be avoided. The controversy of regional versus general anesthesia with respect to patient safety is longstanding. As stated by Sharrock and Savarese: In most cases, the choice of regional or general anesthesia in orthopaedics depends on some or all of the following factors: patient preference, state of health of the patient, expertise of the anesthesiologist, duration of the procedure, and surgeons preference. Many surgeons worldwide perform knee arthroscopy under general anesthesia and it can seem a very satisfactory anesthetic technique; nevertheless it can cause major complications. Ross and Tinker state that the overall risk of death due to anesthesia is about I: 10.000; in the United States, there are 2,400 deaths every year primarily due to anesthesia. Morbidity caused by anesthesia ranges from remporary pressure injuries to comatose vegetative patients. A recent French study found an incidence of 1 major complication per 739 uses of anesthesia. Postanesthetic respiratory depression accounted for nearly half of the cases in which death or coma was totally attributable to anesthecia. In the

REGIONAL BLOCKS FOR KNEE ARTHROSCOPY ion igh iel) est vaIO-

211

lng ive zrn in -obe eci


Of

Eind or ne &e
O-

:r-

th :d le 3. ?l)f :)r e

United States, the incidence of death due to cardiac arrest is 0.9: 10.000 cases; 26% of cardiac arrest occurred in ASA class I or II patients. Of all patients, 18% experience one or more anesthetic complication. Nausea and vomiting constitute 50% of the total postoperative complications, and they may be troublesome in outpatient surgery: sore throat is the second most frequent postoperative complication and may represent injury to trachea and may progress to sequela such as laryngeal granuloma. There is some evidence that mortality and morbidity are lower for regional rather than general anesthesia. S~inat or cpidural lumbar anesthesia may be tiseful pr<jcedures for knee surgery, but they are not advisable for all patients. If the surgical procedure is short, spinal anesthesia is more practical because it takes effect more rapidly than epidural anesthesia. However, spinal anesthesia may be inappropriate for young patients because of the occurrence of post-lumbar-puncture headache Cup to 70% of cases). Spinal anesthesia affects the cardiovascular system more than general zsthesia but :he mortality rate in healthy patients undergoing spinal anesthesia is I : 10,000. The epidural block fails LO induce surgical anesthesia in 2.4% of cases and the risk of subarachnoid puncture is 0.5% to 2% of cases. - The 3-in-l block and the femoral nerve block are \dfe, simple to perform, and they have some advantage\. buch as lack ofcomplete sympathectomy and prolonged postoperative pain relief. In our seriey. more satisfactory anesthesia was ;ieved with ths j-in- I block than with femoral nerve biock alone. The tolerance of the tourniquet was significantly better :n Group 1 patients, who maintained sensibility only :n the area of the sciatic and posterior cutaneous nerve,. than in Group 2 patients, who had a more extender: \ensitive area under the toiirniyuct. The LIX of Iriravenous Runitrazepam and/or fenranyl was reque,:sd by only 17% of the patients of the ,-oup 1, in cozrrast with 50% of the patients of the i)up 2 who ne-&d sedation. This means, in our opin!on. thak the toic:;ince to surgical manipulation (rotaM-y and \arus- .iigus stress manruvres essential to correct diagno\l and surgery) was significantly better ior patients in Cr-~up 1In these patients, pain was not e\oked evcn LV~Z-, stress or traction was applied 10 the jynovial membilne during endo\zopic surgery. The ?er&ption of \( _-.L m:trtcu\~rcs lvithin the joint. esperily if intlamc:. may cause discomfort. but this is -iIdly easily tc: ;rablc to 3 cooperative patient. In 0~11~ cupericncc, anc~-hesia and muscle relaxation were IllOre complete . , 3211 3-in-1 block was performed. The t'crnor-al - :TVL block alone could offer the ad-

vantages of a reduction of the drug dose needed to produce the block and a reduction of the time of onset of the anesthesia. Nevertheless these advantages have no clinical importance because the total amount of the anesthetic drug for the 3-in-1 block is no greater than with the maximum single dose (3 mg/kg).. The shorter onset of anesthesia has also no clinical importance; it is only a difference of 5 minutes. In our series, we observed a relatively high number of conversions to general anesthesia in both groups of patients, without a statistic difference between groups. We cannot exclude that anxiety and/or fear are determining factors in requiring general anesthesia in our randomized series. Some studies have shown that the pain tolerance threshold is significantly different between ethnic groups. x-20 Italian patients seem to be more interested than other ethnic groups in immediate pain relief and they tend to describe a variety of sensations as pain., In our experience, most patients believe that general anesthesia is the sole acceptable way of being anesthetized. An anxious patient could be frightened simply of being in an operating theatre to undergo surgery. and a general anesthesia may be requested for that reason alone. On the other hand. in many cases patients found it desirable to watch their operation on the video screen. The peripheral nerve block does not have to be completely resolved before discharging inpatients,! and we discharged them from the recovery room as soon as possible. given their remarkable cardiovasculn~ and respiratory stability. With outpatients, the use of a short or medium acting local anesthetic could be more appropriate: however. complete recovery of sensation is not a prerequisite for discharge. In our experience no local or general side effects were encountered.

We conclude that the j-in-1 tzchniyue offers a deeper intraoperativc anesthesia and ;i superior muscle relaxation compared with the femoral nerle block alone. The local anesthetic dosage does not reach the maximum single dose (i.e.. 3 g/k?. or 223 mg), even when a lateral supplemental local block is performed by the surgeon, and so the risk of nervotls or cardiovascular complications is nearly ahient. As previously observed by Pate1 et al.. the lateral aspect of the thigh ma> be dilficult to anesthetize. but :t loc:i! infiltration is sufficient to overcome this problem. In our patients who underwent the 3-in- 1block. a suppiementary local lateral anesthesia wai needed in only

712

V. BONICALZI AND hi. GALLINO


lx technique of lumbar plexus anasthesia: The 3 in I block.

20% of cases. If the femoral nerve block alone is performed, the lateral local anaesthesia is mandatory. With the 3-in-1 block, the surgeon is able to maneuvre free of difficulties during surgery because of a more diffuse and profound anesthesia. With the femoral nerve block alone, this was not always the case and, in most cases, the tourniquet is not so well tolerated as it is with the 3-in-1 block. The more prolonged postoperative analgesia also plays a role in choosing the 3-in-I block. We conclude that in our experience knee arthroscopy can be accomplished under femoral nerve block plus local lateral anesthesia, but the 3-in-1 block is the most reliable, safe, and effective regional anesthetic technique for knee arthroscopy.
Acknowledgment:
his linguistic We are indebted to Dr. J. Hoskins for

S, Grossman R. A regional anesthetic technique compared to general anesthesia for outpatient knee arthroscopy. Anesth Anulg 1986:65: I85- 187. 8. Colton T, Civetta JM. A primer for understanding the medical literature. In: Civetta JM, Taylor RW, Kirby RR, eds. Critird care. Philadelphia: Lippincott, 1988: 1699. I7 14. 9. DeJong RH, Ronfeld RA, DeRosa RA. Cardiovascular effects of convulsant and supraconvulsant doses of amide local anesthetics. Anesrh Ad! 1982; 6 1:3- 11. IO. Strichartz GR, Covmo BJ. Local anesthetics. In: IMiller RD, ed. A!ze.rrhesiu. Ed 3. New York: Churchill Livingstone, 1990:437-

7. Pure1 NJ, Flashburg MH, Paskin

.Anesrh

Analg 1973;52:989-996.

-170.
I I. Moore DC, Balfour RI, Fitzgibbons D. Dinzepam treatment of local anaesthetic-induced seizures. Anesthesiology 1979: 50:

3.54-457.
12. Ross AF, Tinker JH. Anesthesia risk. In: Miller RD. ed. ;\!le.rr/i+ .~in. Ed 3. New York: Churchill LivIngstone. 1990:715-7-12. 13. Tiret L, Desmonts JM, Hatton F, Vourch G. Complication ~SSOciated with anesthesia-A prospective survey in France. CUZ Anize.sth Sot J 1986; 33:336-342. 13. Keenan RL, Boyan CP. Cardiac arrest due to anesthesia: A study of incidence and causes. JAMA 1985;253:2373-2377. 15. Atkinson RS, Rushman CR, Lee JA. A synopsis of anaesthesia. Ed IO. Bristol: Wright, I987:662-721. 16. Brown DL, Wedel DJ. SpInal. epldural, and caudal anesthesia. In: Miller RD, ed. &e.srhe.s~~. Ed. .3. New York: Churchrli Livingstone, 1990: I377- 1405. 17. Eledjam JJ, Bruelle P, Vie1 E. DeLaCoussa>e J-E. ,Anesthc\ic et analgesic peridurales. Encycl Med Char---\nesthcsie Reanlmation. Paris: Editions Tcchniqurs, i90?:36-.315-A- IO. 18. Bates MS, Edwards WT, Anderson KO. Ethnocultural intluences on variation in chronic pam perception. /urn 1993:52: 101-I 12. IO. Zola IK. Culture and symptoms: -\n analysts of patients pre\enting complaints. Am Socioi Kc\> 1966; 3 I 6 15-h 19. 20. Stembach RA. Tursky B. Ethnic differences among housewives In psychophysical and skin potential response\ [o electric \hock. Prvch,)p/7~.sloio.r? 1965: I .2-! i-219. 21. Brown DL. We&l DJ. Introducrlon to re_elonal anasrhr\la 11, Miller RD, ed. Ar~r.vrhr.sirr. Ed 3. New Yorh: Churchill Llrln;stone. 1990: 1369-137.5.

assistance.

REFERENCES
I. Shcxock NE. Savarese JJ. Anesthesia for orthopaedic surgery. In: Lliller RD. ed. ,Ane.cti?esiu. Ed 3. New York: Churchill Livingironr. 1990: 1951-1967. 3. Bon~ca JJ. The rnanclgerne~~ ojpain. Ed 2. Philadelphia: Lea & Fr::ger. I990. i .-1r.cnson KS, Rushman GB, Lee 54. A synopsis of anaesthesia. Ed 0. Bristol: Wright, 1987:644-651. 1. W<rsl DJ, Brown DL. Nerve blocks. In: Miller RD, ed. .&e.srheSUM. Ed 3. New York: Churchill Livingstone, 1990: 1107-1437. 5 Gk. no M, Bonicalzi V. Gmziano G, Clemente M, Dcltino U. Lo. .xregional anaesthesia for knee arthroscopy. J Sporrs Trcu~-

mi;: .! Rel Res 1990; 12:-1 I-38.

6. Wj-llr AP, Ramamurthy S, Durriani Z. The inguinal parnvascu-

Femoral Nerve Block as an Alternative to Parenteral Narcotics for Pain Control After Anterior Cruciate Ligament Reconstruction
Brian S. Edkin, M.D., Kurt P. Spindler, M.D., and John F. K. Flanagan, M.D.

Summary: Anterior cruciate ligament ( ACL) reconstruction is asociated

with signiticunt postoperative pain, ~lsuallyrequiring parentersl narc0ticc A prospectivt2 study 01 arthroscopicaliy assisted autograft patzllar tendon XCLR was initiated using Winnies three-in-one femoral nerve block (FNB) as the primary means of postoperative pain control. Patient satisfaction and absence of pxenteral narcotic use indicated clinical success. Of 21 patients studied, 93% had no parenteral narcotics administered lollowing FNR. Ninety-live percent ot patients believed FUR was beneficial and would request another. The average duration of pain control was 29 hours and the majority of patients (79%) believed discharge w;ls possible within 13 hours. There wcrc two patients who railed to respond to FNHI t8r) and no major complications. FNB is ;I safe. reliable. and ctfective form ot analgesia f o l l o w i n g ACLR. eliminating the n e e d f o r parenter: n a r c o t i c s . I(ey Words: Femoral nerve block-Anterior cruciate ligament recoIih[ructioiiPain-Nurcotics-Ketorolac. -

FNR PAIN CONTROI, parenteral narcotic administration as an indicator ot clinical efficacy. In addition, a retrospective questionnaire was administered after the perioperative period to evaluate individual patient experience with this technique. Because only the femoral nerve distribution was specifically assessed for anesthesia following block placement (i.e., obturator and lateral femoral cutaneous distributions were not tested), we elected to use the term femoral nerve block (FNB), rather than threein-one nerve block, in this study.

RCL RECONSTRUCTION

405

MATERIALS AND METHODS Twenty-five <onsecutive primary arthrocopicully assisted ACL reconstructions using autograft middle onethird pateliar tendon grafts were performed on 25 pz tients by a single surgeon (K.P.S.) between 12/l j/c)2 and 7/19/93. One patient. who preferred to use his intraoperative epidural for postoperative analgesia, was excluded from ihe study. The remaining 24 patients comprised the ,tudy group where follow-up was obGned. There \izre 9 female and I5 male subjects between the age5 of I6 and II years. Seventeen of the in.juries were considered acute (operation within 3 months ot in.jup I and 7 chronic. Ten meniscai procedures (I medial rnzniscus repair. 5 partial lateral meniscectomies, 2. partial m e d i a l menisccctomies, a n d 2 Lomhined par-r!,4 medial and lateral menisccctoniy) were perlormei: coincident with ACL reconstrLlctior1. rhe patients UC:= anesthetized by general endotracheal cnlesthcsiu (2 I ;dtients), ~ CTcnerai mask (2 patients). and cpldural ( I paL:=nt). No procedures were \ame day \urgerv. After approylate ancsthcsia was administered, anteromedial and i:terolatcrril portals were fashionccl i;)r arthroscopy ali~: iii,jection with ii I : I niiuture of 0.X% hupivicaine \hi:zout epinephrine and 1% iidocaine with :pinephrine ( I 100.000). Fifty milliliters of this mixure wils injec::d intraarticularly before the incision was made as ti:~cribetl by Weiker et al. All patients had an autogrs:- middle-third patellar tendon XCL reconstruction v, -h 3 two-incision technique using ai Acufex (Xcufc : Vicrosurgical. Inc.. Mansticld, CIA) rear-entry guic; system. A medial parapatellar \kin incision \vas ux ;J for isolation of the middle-third ptcllar tendon ;r.:ft. and a small Ialrrai incision was :Ised for drillir : the femoral tunnel using outsidei l l technique. I tourniquet w a s LIXL~ in all cases and two drains wer: placed at the end of each procedure. Upon cornpIer 7 of the operation, all operated knees were placed in ;-I Air Cast cryocuff (Air Cast. Summit.

ill

in
ild stlee ill

NJ) and knee immobilizer before transfer to the recovery room. Informed consent was obtained by the anesthesia Acute Pain Service for FNB placement. Following completion of the reconstructive procedure, the groin region of the operative side was prepared with betadine and draped with sterile linen. The femoral artery was palpated just inferior to the inguinal ligament and a 22-gauge insulated needle (Stimex: Becton, Dickinson, and Co., Franklin Lakes, NJ) was inserted lateral to the arterial pulse. INerve stimulation was used to elicit a strong quadriceps twitch at a level of 0.5 mA or less, confirming needle placement within the femoral sheath. An injection of0.5% bupivicaine with epinephrine ( I :100,000) was administered in a dose ranging from 2 to 3 mg/kg. All patients were placed on a similar postoperative analgesia protocol. Standard postoperative pain orders included as needed administration of intramuscular ketorolac tromethamine (Toradol; S y n t e x Luboratories, Inc., Palo Alto, CA) and oral narcotic preparations (acetaminophen with hydrocodone or oxycodone). Parenteral narcotics were to be used only in cases of failure of the nerve block and supplemental analgesia. The time, route, and dose of all analgesics administered were recorded. Prcoperatively, patients were counseled in general terms regarding the use of FYI3 as an analgesic supplement. To avoid the introduction of potential biases. no formal discussion was made with patients regardin_ 2 alterations in postoperative nnal~esic protocols or expected patterns of analgesic usage. The patients were mobilized within 24 ~OUI-s of surgery and were instructed in the USC ot crutches and a standard set 01 exercises by the physical therapy staff. No attempt was made to standardize the time of discharge. Upon discharge. the patients were given a prescription for an acetaminophc~~ihycirocodone oral preparation. RESULTS Twenty-three nerve blocks were placed postoperatively in patients who were awake in the Post Anesthesia Care Unit (PACU) and one in the operating room upon completion of the operation. The average time from the end of the operation to placement ot the block was 50 minutes (range, IO to 10-t minutes). Parentcral analgesic use in the PACU before and after FNl3 is compared in Table I Nine patients required narcotics or ketoroiac before block placement. No parenteral narcotics were administered in the PACU following placement 01 the block. Five (3 I %) patients received

--

B. S . EDKIN E T AL.

Pre-FNB

Post-FNB

Nnrcotics Yes
No 18 (75%)

Ketorolac 5 (21%) I9 (797c)

Narcotics
0 (0%) 24 ( 100%)

Ketorolac
5 (21%) I9 179%/c

6 (25%)

ketorolac following block placement. including one patienr who had received narcotics before FNB. For-tysix percent of patients required no supplemental anslc~esia in the PACU before or after block placement. c .Ail patients were admitted to the orthopaedic ward postoperatively. Only two patients 18%) received parenteral narcotics after transfer from the PACU. The first patient, although documented as initially receiving signlticant relief following FNB. required multiple injectlons of morphine sulfate and meperidine beginning 3 hl>urs postoperatively. A second patient received mo?hine sulfate by patient controlled analgesia pump (PC.\) for approximately 7 hours. beginning 16 hours after block placement. He was discharged home on an oral narcotic preparation the morning following sut-= germ F&in in the remainder of patients was adequaicly conrrI>lled with combinations of parentera ketorolac and t-al narcotic deri\,ative/XPAP preparations (Table 2). T,.Lentv-five percent of patients required only or;11 preF;rations (oxycodone/APAP o r hydrocndonc/ AP.2.P) after discharge from the PACIJ. An additional four 7atients received only a single dose of parenteral ketor-)lac in addition to oral preparations following discharge from the P,ACU (Table 3). Thus. -1-23, of pati,:[s required zero or one dose of non-narcotic analgesii ,,,ia a parenteral route after leaving the recovery rouIT J3: <XI levels of bupivicnine following FNB plnccmen. kere observed in a small sample of our patient pops 2tion (n = 3; 1 3_cc). Peak serum levels averaged I .0-I _ z/mL (range. 0.884 to 1.20) and were seen hetwec- I and I .5 hour\ post injection. It has been sugTestC: that toxicitv occurs in milder forms with ccrum i) levei- JS t o w a5 i .6 pg/mL. with morc severe side cffec-~ at le\,els exceedins 2.3 PglmL a n d d b/g/ n1L. c Howe\ er, several studies have reported Icvels betwr:n these values without toxic symptoms. Al. 3atients were contacted by phone regarding their post< -erativr experience. Data collected from the telephone ntervielv is summarized in Table .3. The average perce ;-d duration of significant relief from FNB was

29 hours (range, 0 to 3.5 days). Twenty-three rrf 24 patients (95%) believed FNB to be very (n -= 21~ or somewhat (n = 2) beneficial and would rcrir:r:at another nerve block if undergoing a similar sur:;::al procedure in the future. The only patient who dec! t was considered to have a failed block attempt, ir ing required significant doses of parenteral narcdrtjc beginning approximately 3 hours from the time of F placement. Interestingly. one patient who respctn somewhat beneficial to this question used nc> upplemental pain medications during the entire ho(,rir!al course. The average time of hospitalization was I .X ..,/v with 7 discharges on postoperative day one, 14 on iiay two, and 3 on day three. XII patients were retro%f?estively questioned regarding the possibility of carlicr discharge. A majority of patients believed that IhLY could have comfortably been discharged earlier (l*;lbfc 3). Three of the five patients declining early dischW$ reported pain as a factor in their decision. The two ,)/ho did not have significant pain cited a general fcclili 1 weakness and dependence as the main 1~ l(ir?. Most patients experienced adequate pain relicl fr(M ora1 narcotic preparations following discharge (l;tblc 3). Two minor complications occurred, with no d(JlJ:. ent evidence of more serious local or systemic f,i(lC effects. Two patients developed a mild contact tlcrrrl;ltitis at the sitr, of injection, presumably secontl:ti / I0 the local preparation in the groin region. Two 1~1111 :,I) (both of whom underwent general anesthesia) cxl)~:~I enced urinary retention requiring limited cathctcf i/z tion in the immediate postoperative period.

FNBs and three-in-one nerve block< have bee11 il((( for lower extremity anesthesia and analgesia untl~.~ I variety of operative and nonoperative condition\ \I though a majority of repon; in the literature ha\c (I( scribed favorable results, \,ariability in operative/ill/III\ site. operative procedure. technique of FNB ac1111llil tration, type and concentration of anesthetic. and Ii Y((

Narc0t1cs Yes NO 2 ( 8 !; ) 21 (92: )

FNB PAlN CONTROL ACL RECONSTRUCTION f 34 ?I) uest $21 med


1av-

107

.Another Yes = 96% No = 4%

Beneticialf Very = 88% Somewhat = 8% Not = 4%

Possible D/Q SD = 33% N D - A M = 36% Other = 21%

Oral Meds$ None = i3% Adequate = 71% Marginal = So/n Inadequate = 84

Duration//

Avg = 29 h
Range = O-84 h

)tics TNB tded iuplital iays da> )ec-lie1 hey tble urge VI10 : of )rs.
0111

Abbreviations: SD. same dav: ND. next dav: i AM, other. ii would you have another 6NB? j- Is the FNB beneticiai? _ When would you be able to be discharged? 3 Efiicacy of oral meds foilowIng discharge. I Recall of subjective duration of pain relief.

ible jaride
Xlto

nts 3%
La-

ICI -

versus per weight dosing make it difticult to extrapolate data and compare results. However, these reports have provided us with useful information regarding the general efiectiveness. or ineffectiveness, of various methods and means of FNB administration. Lynch et al. studied 208 patients who received intermittent femoral nerve blockade using an indwelling catheter t?)llo~. ing ACL reconstruction (operative technique not spe<itied). They concluded that this technique provide; safe and reliable analgesia, improves patient mobility. has a high patient acceptance and [reduces: systemic analgesic demand following [ACL] repair. Edwards and Wright reported similar results using continuous bupivicaine infusion via an indwelling catheter following total knee replacement, reporting decreased opiate requirements and lower pain scores in :he first 24 hours. Hord et al.s reported significantly &creased pain scores in the first 3-4 hours following coE:Inuous block administration after primary total kcsz arthroplasty. Total morphine use by PCA. as well ~5 pain scores on the second postopcrative day. did ?ot differ- significantly when compared with controls. :Xingrose and Cross prospectively compared blocks: and unblocked patient groups following ACL rczonstruction (unspecified operative procedure) and f1 ilnd a significant difference in narcotic use among ptrlents receiving FNB. A study by Matheny et al. _ jmparing PCX and continuous lumbar plexus block LPB) following ACL reconstruction, found that par;-nts receiving LPB used dramatically less parenterai narcotic and had significantly fewer narcotic-associate; side effects. In contrast 1 these positive results, Tierney et al. us&l 20 IllL t - 0.15% plain bupivicaine to effect an FNB followic; knee ligament reconstruction (unhpecitied proc:jure). They found that narcotic requirements were rsclced in the recovery room and that time to first narcotiU lose was prolonged in the study group. However, the; lid not find a significant difference in

the total dose of narcotic used by the FNB group overall. No serious complications were reported in these series, as in our series. The systemic toxic side effects of bupivicaine for local or regional anesthesia are well described. In addition, there have been sporadic reports of more localized complications. including transient and prolonged femoral nerve palsy following in.jection of safe doses of anesthetic for femoral and three-in-one nerve blocks.- Signs of toxicity include central nervous system reactions (includin,(7 nausea, vomiting, chills, restlessness, anxiety, dizziness, blurred vision, tremors, and seizures) and cardiovascular system reactions (including myocardial depression, diminished cardiac output, hypotension, bradycardia, heartblock, and ventricular dysrhythmias), as well as rare signs of local and systemic hypersensitivity. The more severe central nervous system and cardiovascular side effects are usually associated with higher plasma concentrations possibly related to accidental intravenous or arterial administration. unusually rapid uptake from the site of administration, decreased degradation, or excessive administration. Several studies have reported safi: and effective administration for peripheral nerve block using doses that exceed the recommended 2 mgikg Misra et al. administered 0.5% bupivicaine with epinephrine ( 1:200,000) at 3 mg/kg for Femoral threein-one combined with sciatic nerve block for lower extremity surgery, achieving concentrations of up to 3
In countrle\ other than the United States. the doaage of 0.25% and (X5?+ concentrations of bupivicaine is ? mg/kg, with the mxx~mum o f 1 0 0 mg wIthout eplnephnne a n d I50 mg w i t h eplnephrelne I :X0.000. In the United States, imaximum doses of such concentrutiolls have not been determined, and the package insert states: 1~) closi experience to date 1s with single doses of Marcame up to 175 nlg with eplnephrine I :7_00,000, and I75 mg without epinephrine: more or le5\ drug may be used depending on the IndivxhAlzatlon of each ca\e: and (b) In climcal stutiia to date. total dally dox\ ha\ze been up to 100 mg. Until further experience is galned, this dew hould not be exceeded in 7-l hours.

-LO8

B. S. EDKIN ET AL other more severe toxicities in the gastrointestinal, ca rdiovascular, hematologic, nervous, renal, and hepatic systems. Hypersensitivity has also been reported. We do not regularly use parenteral ketorolac for longer than 5 days duration or the oral preparation for larger than 1 to 2 weeks secondary to a reported increased incidence of associated side effects. By eliminating parenteral narcotic use as an indicator of success, we have outlined a protocol for the safe, effective, and reliable administration of FNB for ACL reconstruction. We are currently using this technique with good clinical results on patients undergoing ACL reconstruction using patellar tendon or hamstring and following patellar realignment using the modified Elmslie-Tritlat procedure. Prospective studies are currently underway evaluating the usefulness of FNB following outpatient ACL reconstruction, and as the primary anesthetic during routine arthroscopy. We have shown that blockade of the femoral nerve, when administered using the described technique. can provide a reliable, safe, and effective alternative form of analgesia in the immediate postoperative period fo!lowing ACL reconstruction. The complication protiie and patient tolerance for this procedure are excellent. With s~~cessf~~l FNB, parenteral narcotic administration is eliminated and patient pain relief is excellent. FNB is a valuable adjunct to ACL reconstruction surger-v. REFERENCES

PglmL. !4oore et al. administered up to 400 mg (6 to IO mg/kg) bupivicaine with epinephrine ( 1:320,000) for sciatic. femoral, and lateral femoral cutaneous nerve blocks, achieving levels of up to 3.16 ,q/mL. Robison et al. found that peak plasma concentrations were significantly diminished in patients receiving anesthetic containing epinephrine. No systemic side effects were reported in these studies. By evaluating the available historical data, as well as our own clinical experience, we attempted to design our FNB protocol based on factors that appear to be important in obtaining consistent, effective, and durable analgesia. These include an anatomic approach to the femoral triangle, the use of a nerve stimulator to isolate the optimal injection site, and a per weight dose of a 0.35 bupivicaine preparation containing epincphrine. Previous studies. as well as our complication profile and serum level determinations, suggest that safe and effective analgesia can be achieved with up to 3 mg/kg of 0.5% bupivicaine with 1:200.000 epinepl?rine. Using FNB. with ketorolac tromethamine and oral narcotic analgesics as supplements, we have essentially eliminated the need for narcotic use via the parenteral route. This combination has provided a highly effective means of controlling postoperative pain following the operatiLt procedure described. It is associated with a high degree of patient approval (95%). favorable coniplicstion profile, and low rate of failure (8%). Although not specifically evaluated in this study, there appear< to be less association with the undesirable side effects ot parenteral narcotics, such as sedation. naujea, and generalized malaise. In our experience, patients appear to tolerate the postoperative period better and seen to mobilize more readily than when parenteral narcotics are used as the primary means of pain <ontroi. ;ub.jective data collected demonstrate that patients s~~ulci tolerate earlier discharge using this protocot. Ifrftitctivety used in this manner, cost savings can uttimarsib be realized, especially with the utilization ~>f day -urgery centers tor ACL reconstruction. We have ~u~quently evaluated these savings in a proypecti\i Yrudy. Our u+ of ketorolac ior breakthrough pain has prob,lbly imzrovrd our results in a favorable way: to what degree [Jnnot be determined in this study. We will continu: to use it in our protocol because it appears to pro\ 122 acceptable supplemental analgesia without (he und-\irable \ide-effects of parenteral narcotics. Although :<I obvious toxicity resulted from our use of ktorotl~. a spectrum of bide effects ranging from the more cc iinion gasti-ointestinal complaints associated ii,ith oi.72~ non\tcroidat ;unti-inNalnmatory drugs to

:ar-

Sic
We ger

ger ;ed cakfe, CL lue ZL md mtlY ng


inte.

an
i-m ~,Iile nt. ant.

lr-

r-al fir LtZ ch xl xi lit i,,l(, e :1. I.< I,;

The Division of the Sciatic Nerve in the Popliteal Fossa: Anatomical Implications for Popliteal Nerve Blockade
Jerry D. Vloka, MD, PhD*, Admir Hadz ic , Daniel M. Thys, MD
MD, PhD*,

Ernest April,

PhD,

and

Departments of *Clinical Anesthesiology and Anesthesiology, St. Lukes-Roosevelt Hospital Center; and the Department of Anatomy, Columbia University College of Physicians and Surgeons, New York, New York

The sciatic nerve (SN) originates from the L4-S3 roots in the form of two nerve trunks: the tibial nerve (TN) and the common peroneal nerve (CPN). The TN and CPN are encompassed by a single epineural sheath and eventually separate (divide) in the popliteal fossa. This division of the SN occurs at a variable level above the knee and may account for frequent failures reported with the popliteal block. We studied the level of division of the SN in the popliteal fossa and its relationship to the common epineural sheath of the SN. The level of division of the SN sheath into TN and CPN above the

knee was measured in 28 cadaver leg specimens. The SN was invariably formed of independent trunks (TN and CPN) encompassed in one common epineural sheath. The SN divided at a mean distance of 60.5 27.0 mm (range 0 to 115 mm) above the popliteal fossa crease. We conclude that the TN and CPN leave the common SN sheath at variable distances from the popliteal crease. This finding and the relationship of the TN and CPN sheaths may have significant implications for popliteal block. (Anesth Analg 2001;92:2157)

locks of the sciatic nerve (SN) in the popliteal fossa are associated with a variable success rate (13). With this technique, also known as a popliteal block, the needle is inserted 50 70 mm above the popliteal fossa crease and advanced toward the SN (2,4). The tip of the needle ideally should be positioned next to the main trunk of the SN before its separation into the tibial nerve (TN) and common peroneal nerve (CPN). Injection of local anesthetic in the vicinity of only one of these components may result in an incomplete block. We examined the anatomical variations of the SN in the popliteal fossa and determined the optimal distance from the popliteal crease for the needle to be placed in the popliteal block. The SN derives its fibers from the L4-S3 spinal segments and is almost 2 cm wide at its origin near the sacral plexus. Two separate nerve trunks (TN and CPN) enveloped by a common fascial sheath (epineural sheath) can be distinguished from the onset (5,6). These two trunks leave the pelvis (together with the posterior cutaneous nerve of the thigh) through the sacro-sciatic foramen between the tuberosity of the ischium and the

greater trochanter of the femur. The TN and CPN eventually diverge, with the TN descending medially through the popliteal fossa into the back of the leg and the CPN diverging laterally from the midline to pass behind the head of the fibula and lateral to its neck (7). The SN provides motor branches to the hamstrings and all muscles below the knee. The SN also provides the sensory innervation to the posterior thigh and entire leg and foot below the knee (except the medial aspect, which is innervated by the saphenous nerve, a branch of the femoral nerve).

Methods
The lower extremities of 15 adult cadavers were obtained for dissection of the popliteal fossae. The cadavers had been deceased from 6 to 18 mo and were free of gross pathology. They were embalmed for anatomical purposes in a solution of phenol (13%) as the principal fixative and glycerin (28%) for retention of water content. The popliteal fossa crease was identified, and the skin and subcutaneous tissue overlying the popliteal fossa were removed to the level of the superficial fascia of the hamstring muscles. Further dissection of the more caudad aspect of the fossa was performed to identify the more superficial CPN. The epineural sheath of the CPN was dissected proximally to the main trunk of the CPN until the point at which
Anesth Analg 2001;92:2157

Accepted for publication August 23, 2000. Address correspondence and reprint requests to Admir Hadz ic , MD, PhD, Department of Anesthesiology, St. Lukes-Roosevelt Hospital Center, 1111 Amsterdam Avenue, New York, NY 10025. Address e-mail to ah149@columbia.edu.
2001 by the International Anesthesia Research Society 0003-2999/01

215

216

REGIONAL ANESTHESIA AND PAIN MEDICINE VLOKA ET AL. DIVISION OF THE SCIATIC NERVE IN THE POPLITEAL FOSSA

ANESTH ANALG 2001;92:2157

its sheath merged with that of the TN into the single epineural sheath of the SN. This point was defined as the division of the SN. For each dissected leg, distances from the division of the SN to the popliteal fossa crease were measured with calipers (degree of accuracy 1 mm) and recorded. Distances from the division of the SN to the popliteal fossa crease were compared between the left and right legs by paired t-test. In order to determine whether the distances differed by sex, mean distances for each cadaver were calculated; these mean distances were then compared by Students t-tests. P values 0.05 were considered statistically significant. All analyses were performed with the Statistical Package for the Social Sciences (SPSS Version 5.02 for Windows; SPSS Inc, Chicago, IL).

Results
We examined 14 right and 14 left legs from seven female and eight male adult cadavers. The left leg of one male cadaver and the right leg of one female cadaver could not be examined because of prior disruptions in the area to be measured. The connective tissue sheaths surrounding the TN, the CPN, and the division of the SN were easily identified in all dissected specimens (Figure 1). Distances from the division of the SN to the popliteal fossa crease did not differ between the left and right legs of the 13 cadavers whose legs could both be measured (57.5 25.7 mm and 66.2 23.9 mm, respectively). There was no difference in the measured distance by sex (55.1 25.5 and 64.2 27.5 mm for male and female legs, respectively). The fascial sheath enveloping the SN divided into two distinct sheaths surrounding the CPN and TN as they separated in the distal popliteal fossa. The mean distance above the popliteal fossa crease at which the SN divided was 60.5 27.0 mm. The measured distances varied widely, ranging from 0 to 115 mm (Figure 2). The coefficient of variation was 44.6%. In 57% of the specimens, the SN divided within 70 mm of the popliteal crease; in 75% of the legs, the SN divided within 81 mm; and in all legs studied, the SN divided within 115 mm.

Figure 1. Division of the sciatic nerve in the popliteal fossa. Popliteal fossa is shown dissected. The sciatic nerve (SN) divides into tibial (TN) and common peroneal nerves (CPN) above the popliteal fossa crease (crease). The TN and CPN depart the common epineural sheath (ES) of the SN and descend into the popliteal fossa enveloped by their respective sheaths.

Discussion
Anatomical variations in the level at which the SN divides into the TN and CPN have been suspected as a possible cause for incomplete block of the SN in the popliteal fossa (5,8,9). Our data indicate that the division of the SN does occur at highly variable distances from the popliteal crease. These distances did not differ by sex or by laterality. Indeed, according to our anatomical model, if a needle is inserted at 50 mm (1) or 70 mm above the popliteal fossa crease, the chance

Figure 2. Division of the sciatic nerve above the popliteal fossa crease. Horizontal bars indicate distances above the popliteal fossa crease at which the sciatic nerves divided in individual leg specimens.

that the tip of the needle will be proximal to the division of the SN is only 46% and 57%, respectively. In contrast, insertion of the needle at 100 mm above the popliteal fossa crease, as suggested by Singelyn et al. (4), virtually ensures placement of the needle in the vicinity of or proximal to the division of the SN.

ANESTH ANALG 2001;92:2157

REGIONAL ANESTHESIA AND PAIN MEDICINE VLOKA ET AL. DIVISION OF THE SCIATIC NERVE IN THE POPLITEAL FOSSA

217

To improve the success rate of popliteal blocks, some investigators have suggested a double-injection technique, in which both branches are separately identified and anesthetized (10). Others have suggested injecting a larger volume of local anesthetic to increase the spread within the epineural sheath to reach both the TN and CPN (5). It is important to keep in mind that the SN is composed of independent medial and lateral divisions that are physically but not functionally joined by a common connective tissue sheath. These nerve trunks (TN and CPN) are bundled together with multiple layers of connective tissue, but they do not exchange nerve fibers (5). This is important in popliteal nerve blockade, because TN and CPN remain enveloped in their respective sheaths as they diverge from the epineural sheath of the SN. This, in turn, may limit exposure of one of these branches to the local anesthetic when the injection is made distal to the division of the SN. It is possible that distortion in the anatomy caused by the embalming process and dissection may limit the applicability of these data to clinical practice. However, we exercised great care in selecting undistorted specimens and performing dissections of the popliteal fossae. In conclusion, the SN divides into the TN and CPN at highly variable distances above the popliteal fossa crease. If our findings are applicable to clinical practice, when the needle is inserted at commonly suggested insertion sites in performing the popliteal block (50 70 mm), local anesthetic may be deposited in the vicinity of the TN or the CPN, but not both. However, insertion of the needle at 100 mm above the popliteal crease virtually ensures placement of the needle in the vicinity of or proximal to the division of the SN. Although these findings may not be of importance in

the double-injection technique (10), in which the TN and CPN are identified and anesthetized by separate injections of local anesthetic, they may have implications for the more commonly used single-injection technique.
The authors thank Drs. Kayser Enneking and Alan Santos for their editorial comments.

References
1. Rorie DK, Byer DE, Nelson DO, et al. Assessment of block of the sciatic nerve in the popliteal fossa. Anesth Analg 1980;59:371 6. 2. Brown DL. Popliteal block. In: Brown DL, ed. Atlas of regional anesthesia. Philadelphia: WB Saunders Co, 1992:109 13. 3. Hadz ic A, Vloka JD. A comparison of the posterior versus lateral approaches to the block of the sciatic nerve in the popliteal fossa. Anesthesiology 1998;88:1480 6. 4. Singelyn FJ, Gouverneur JM, Gribomont BF. Popliteal sciatic nerve block aided by a nerve stimulator: a reliable technique for foot and ankle surgery. Reg Anesth 1991;16:278 81. 5. Sunderland S, ed. The sciatic nerve and its tibial and common peroneal divisions: anatomical and physiological features. In: Nerves and nerve injuries. New York: Churchill Livingstone, 1978:925 66. 6. Vloka JD, Hadz ic A, Lesser JB, et al. A common epineural sheath for the nerves in the popliteal fossa and its possible implications for sciatic nerve block. Anesth Analg 1997;84:38790. 7. Birch R, Bonney G, Wynn Parr CB. The peripheral nervous system. In: Birch R, Bonney G, Wynn Parr CB, eds. Surgical disorders of the peripheral nerves. London: Churchill Livingstone, 1998:1137. 8. Benzon HT, Kim C, Benzon HP, et al. Correlation between evoked motor response of the sciatic nerve and sensory blockade. Anesthesiology 1997;87:548 52. 9. Hadz ic A, Vloka JD, Kitain E, et al. Division of the sciatic nerve and its possible implications in popliteal nerve blockade [abstract]. Anesthesiology 1996;85:A733. 10. Paqueron X, Bouaziz H, Macalou D, et al. The lateral approach to the sciatic nerve at the popliteal fossa: one or two injections? Anesth Analg 1999;89:12215.

ABA Regional Anesthesia and Pain Management Questions


1993 A
5. The elimination half-life of an amide local anesthetic is prolonged in which of the following conditions? (1) Liver disease (2) Term pregnancy (3) Heart failure (4) Kidney disease 19. Trigeminal neuralgia is characterized by (1) unilateral, intense, paroxysmal pain of sudden onset (2) diminished sensation in the distribution of the maxillary division of the trigeminal nerve (3) normal function of the glossopharyngeal nerve (4) resolution of symptoms by injection of local anesthetic at trigger points 27. Landmarks used in performing a superior laryngeal nerve block include the (1) transverse process of C6 (2) cricoid cartilage (3) angle of the mandible (4) greater cornu of the hyoid cartilage 30. Landmarks for caudal block include the (1) sciatic notch (2) posterior-superior iliac spines (3) iliac crests (4) sacral cornu 37. Landmarks for the sciatic nerve via a posterior approach include the (1) posterior superior iliac spine (2) coccyx (3) greater trochanter of the femur (4) iliac crest 50. Stellate ganglion block is associated with ipsilateral (1) mydriasis (2) diaphoresis (3) exophthalmos (4) scleral hyperemia 58. An asymptomatic 32-year-old man with asthma undergoes hemiorrhaphy under 1.5% lidocaine epidural anesthesia to a sensory level of T2-3. Which of the following will occur with this level of anesthesia? (1) The ability to cough will be normal (2) Vital capacity will be unchanged (3) Intraoperative bronchospasm will be prevented (4) Tidal volume will be unchanged

59. True statements concerning epidurally administered morphine include: (1) The long duration of analgesia results from high lipid solubility (2) Pruritus is completely reversed by naloxone (3) Plasma morphine levels are lower than those seen after intramuscular administration (4) Analgesia is inadequate for the pain of labor

69. A 24year-old man has constant burning pain three months after sustaining a crush injury to the arm. The injured muscles and joints are healed. Findings consistent with a diagnosis of causalgia include (1) beads of perspiration on the skin (2) skin discoloration (3) hypersensitivity to touch (4) warm extremity 79. A 30-year-old woman has difficulty talking 15 minutes after initiation of interscalene block for closed reduction of a dislocated shoulder. The most likely cause is (A) cervical sympathetic block (B) delayed systemic toxic reaction (C) phrenic nerve paralysis (D) pneumothorax (E) recurrent laryngeal nerve block

105. Myofascial pain is an example of (A) a central pain state (B) neuropathic pain (C) psychogenic pain (D) somatic pain (E) visceral pain 116. Surgery is cancelled 10 minutes after initiation of intravenous regional anesthesia with 50 ml of lidocaine 0.5%. To ter-minate anesthesia safely, what is the most appropriate timing for deflating the tourniquet? (A) Immediately if benzodiazepines have been administered (B) Immediately after intravenous administration of ephedrine 10 mg (C) Immediately, followed by repeated reinflation and deflation (D) In no less than 20 minutes after initial injection (E) In no less than 45 minutes after initial injection 120. Eight hours after abdominal surgery, a 51-year-old patient becomes increasingly somnolent. Epidural morphine 5 mg was administered immediately following the procedure. Postoperatively, respiratory rate has not decreased below 12lmin and SpO, has remained greater than 92%. Arterial blood gas analysis shows PaO, 80 mmHg, PaCO, 82mmHg, and pH 7.1. Which of the following is the mo st appropriate conclusion? (A) Analysis of the blood sample was delayed (B) The blood sample was venous rather than arterial (C) The patient is receiving supplemental oxygen (D) The pulse oximeter readings are falsely high (E) No treatment is required at this time 138. Characteristics of postdural puncture headache include (A) incidence unrelated to the timing of ambulation (B) increased severity with addition of vasoconstrictors to the anesthetic (C) less frequent occurrence if the needle bevel is perpendicular to the direction of dural fibers (D) more frequent occurrence in men (E) prevention by prophylactic epidural blood patch 139. Which of the following statements concerning the superior laryngeal nerve is true? (A) It provides sensory innervation to the subglottic surface of the vocal cord (B) It provides sensory innervation to the inferior surface of the epiglottis (C) It is a branch of the glossopharyngeal nerve (D) It is blocked by injection of anesthetic near the lateral portion of the cricothyroid membrane (E) It is the most commonly injured nerve during thyroid surgery

145. Which of the following is the most likely sequela of interscalene brachial plexus block? (A) Cervical epidural block (B) Hemidiaphragmatic paralysis (C) Pneumothorax (D) Seizure (E) Vocal cord paralysis 150. After an axillary brachial plexus block, the patient feels pain when the surgeon clips the skin over the thenar eminence. The most likely cause is inadequate anesthesia in the distribution of the (A) intercostobrachial nerve (B) median nerve (C) musculocutaneous nerve (D) radial nerve (E) ulnar nerve

156. A 3%year-old man has acute onset of low back pain, lower extremity weakness, and bladder dysfunction. He had a lumbar laminectomy two years ago. A myelogram shows disk herniation at L4-5. The most appropriate management i s (A) bed rest (B) administration of a nonsteroidal anti-inflammatory agent (C) epidural administration of a corticosteroid (D) epidural administration of a local anesthetic (E) surgical decompression 162. Local anesthetics block nerve conduction by (A) closing calcium channels (B) decreasing intracellular calcium concentration (C) decreasing potassium conductance (D) causing extrusion of intracellular potassium (E) inhibiting cellular influx of sodium 170. A successful ankle block for transmetatarsal amputation of the first and second toes should include each of the follow-ing nerves EXCEPT the (A) saphenous (B) deep peroneal (C) superficial peroneal (D) sural (E) tibia1

1993 B
6. Which of the following increases the cephalad spread of hyperbaric intrathecal local anesthetics? (A) Cephalad-directed needle bevel (B) Coughing (C) Lithotomy position (D) Obesity (E) Rapid injection

8. Intrathecally administered opioids exert their analgesic effects primarily in the (A) brain stem (B) fourth ventricle (C) spinal nerve roots (D) spinothalamic tracts (E) substantia gelatinosa 11. Which of the following is a cardiorespiratory effect of epidural block to a T4 sensory level? (A) Decreased expiratory reserve volume (B) Decreased tidal volume (C) increased circulating catecholamine concentrations (D) Increased heart rate (E) Unchanged vital capacity 13. A 67-year-old man undergoes spinal anesthesia with hyperbaric tetracaine 10 mg for transurethral resection of the prostate. At the end of the 50-minute procedure, the level of anesthesia is T6 and blood pressure is 120/70 mmHg. Within two minutes after transfer to a stretcher, the patient has nausea and blood pressure decreases to 76/42 mmHg. Which of the following is the most likely cause of the acute hypotension? (A) Acute congestive heart failure (B) Decreased venous return (C) Dilutional hyponatremia (D) Progression of sympathetic block (E) Unrecognized bladder perforation 41. If administered epidurally in equipotent doses, which of the following opioids will produce analgesia over the greatest number of dermatomes? (A) Fentanyl (B) Hydromorphone (C) Meperidine (D) Morphine (E) Sufentanil

42. Which of the following is decreased by alkalinization of a 1.5% lidocaine solution? (A) Concentration of free base (B) Dose required for anesthesia (C) Duration of anesthesia (D) Intracellular concentration of ionized lidocaine (E) Time to onset of anesthesia

44. Twelve hours after an uneventful hysterectomy with lidocaine epidural anesthesia, a 70-year-old woman has partial paralysis of the lower extremities. She is receiving morphine 0.5 mg/hr through an epidural catheter and is pain free. On examination, definite motor loss is noted in the lower extremities, but no other deficits are apparent. The most appropriate action at this time is to (A) administer naloxone (B) substitute fentanyl for morphine infusion (C) remove the epidural catheter (D) obtain an MRI of the lumbar spine (E) reassure the patient 47. Which of the following characteristics of local anesthetics is associated with long duration of action?

(A) High degree of lipid solubility (B) High degree of protein binding (C) High molecular weight CD) High pK, (E) Presence of ester linkage 54. Which of the following is the most likely cause of apnea occurring after a retrobulbar block? (A) Epidural injection (B) Increased intracranial pressure (C) Oculopontine reflex (D) Ophthalmic artery injection (E) Subarachnoid injection 75. A combined epidural and general anesthetic is used for aortofemoral bypass surgery. Just prior to extubation, the patient received morphine 5 mg through the epidural catheter. Eleven hours later, he is unresponsive while breathing 40% oxygen from a face mask. Respiratory rate is 6/ min and SpO, is 92%. Arterial blood gas analysis shows PaO, 80mmHg, PaCO, 84 mmHg, and pH 7.16. Which of the following statements concerning this patient is true? (A) Hypercarbia is contributing to the decreased level of consciousness (B) Naloxone is ineffective for reversing the respiratory depression (C) The oxygen saturation is higher than expected because of the pH (D) The risk for respiratory depression would have been lower with subarachnoid administration of 0.5 mg mor-phine (E) Residual local anesthetic is contributing to the respiratory depression

78. Five minutes after intrathecal administration of tetracaine 12 mg in hyperbaric solution, a 60-year-old man has a weak hand grasp. Respirations are normal, heart rate has decreased from 80 to 45 bpm, and blood pressure has decreased from 150/80 to 90150 mmHg. The most appropriate management at this time is (A) administration of atropine (B) administration of ephedrine (C) administration of phenylephrine (D) placement of the patient in the head-down position (E) observation 81. Which of the following nerves should be blocked for an operation at the medial aspect of the lower leg? (A) Femoral (B) Sciatic (C) Obturator (D) Common peroneal (E) Tibia1

90. A 40-year-old woman has continuous nondermatomal burning pain of the distal foot four weeks after sustaining a meta-tarsal fracture. On examination, the foot is mildly swollen, tender, and cool. Which of the following statements concerning this condition is true? (A) A radiograph of the distal bones of the painful foot will show severe osteoporosis (B) A technetium scan of the distal joints of the painful foot will show decreased uptake (C) Early use of opioid analgesia will prevent progression of the symptoms (D) Intravenous phentolamine will relieve the pain (E) The chance of spontaneous recovery within eight weeks is greater than 80%

91. Evaluation of a postoperative neurologic deficit discloses inability to oppose the thumb and little finger, weakness of abduction of the thumb, and loss of flexion of the distal phalanx of the index finger. This problem is most likely related to (A) paresthesia occurring during an interscalene brachial plexus block (B) attempted radial artery cannulation at the wrist (C) inadequate padding under the elbow (D) attempted venipuncture in the antecubital fossa (E) abduction of the upper humerus against an ether screen 104. Before awake nasal intubation in a patient who has been NPO, areas to be anesthetized are supplied by which of the following nerves? (1) Glossopharyngeal (2) Superior laryngeal (3) Recurrent laryngeal (4) Hypoglossal

105. A 40-year-old patient is referred to a pain clinic for evaluation of right upper quadrant pain six months after cholecystectomy performed through a subcostal incision. Which of the following procedures would provide diagnostic information? (1) Intercostal nerve blocks (2) Celiac plexus block (3) Differential spinal block (4) Lumbar sympathetic block 114. Advantages of performing spinal anesthesia via the lateral approach include (1) larger opening for needle insertion than for the midline approach (2) avoidance of the calcified interspinous ligament in the elderly (3) less flexion of the spine required than for the midline approach (4) less likelihood of peridural vein puncture than for the midline approach 115. In meralgia paresthetica (1) there is pain in the anterolateral aspect of the thigh (2) the obturator nerve is involved (3) obesity is an associated factor (4) neurolytic alcohol block is the treatment of choice

118. The duration of an epidural block can be increased clinically by (I) use of a local anesthetic with low protein binding (2) use of a local anesthetic with low pK, (3) addition of sodium bicarbonate to the local anesthetic (4) increasing the total dose of the local anesthetic 125. True statements concerning regional anesthesia with peripheral nerve blocks for an operation on the knee using a tourniquet include: (1) The inguinal perivascular block includes the obturator nerve (2) Paresthesias are required during sciatic block (3) The lateral femoral cutaneous nerve must be blocked (4) Block of the lumbar plexus in the psoas compartment provides adequate anesthesia 126. Proximal spread of a local anesthetic solution placed in the axillary perivascular space is promo ted by (1) increased volume of the local anesthetic agent (2) digital pressure distal to the injection site (3) cephalad direction of the needle (4) adduction of the shoulder after the injection

127. Indications for neurolytic celiac plexus ablation include pain due to carcinoma of the (1) sigmoid colon (2) kidney (3) ovary (4) pancreas 34. Which of the following peripheral nerves must be blocked for removal of a glass splinter from the plantar surface of the heel? (1) Tibia1 (2) Saphenous (3) Sural (4) Superficial peroneal 142. Neural fibers that transmit pain include (1) B fibers (2) C fibers (3) A-alpha fibers (4) A-delta fibers 174. Complications of stellate ganglion block include (1) elevation of the ipsilateral hemidiaphragm (2) total spinal anesthesia (3) seizures (4) hoarseness

1992 A
8. Anatomic landmarks for a deep cervical plexus block include (1) the mastoid process (2) the superior ala of the thyroid cartilage (3) Chassaignacs tubercle of the sternocleidomastoid muscle (4) the midpoint of the posterior border of the sternocleidomastoid muscle 15. Anatomic landmarks for sciatic nerve block in the lateral position include the (1) greater trochanter of the femur (2) tip of the coccyx (3) posterior superior iliac spine (4) ischial tuberosity 18. A compression injury of the ulnar nerve at the elbow (1) causes weak adduction of the thumb (2) is more likely after intraoperative abduction than after positioning the arm at the side (3) causes sensory loss in the fifth digit (4) causes wrist drop 20. Relative contraindications to administration of thoracic epidural anesthesia include (1) bilateral rib fractures (2) postherpetic neuralgia at T6 (3) angina pectoris (4) aortic stenosis

30. Compared with ethanol for celiac plexus ablation, phenol is associated with (1) less neural regeneration (2) greater neurotoxicity (3) more profound hypotension on injection (4) less pain on injection 44. True statements concerning epidural administration of morphine include: (1) The incidence of severe respiratory depression requiring antagonist therapy is 10% to 15% (2) The primary cause of analgesia is stimulation of descending inhibitory tracts (3) Pruritus is caused by central neurogenic histamine release (4) Nausea is caused by stimulation of supraspinal opioid receptors 73. Which of the following additives accelerates the onset of lidocaine axillary block without shortening duration? (A) Carbon dioxide (B) Dextran (C) Dextrose (D) Epinephrine (E) Hyaluronidase 78. The physiologic function most likely to be spared when a local anesthetic differential nerve block is administered is (A) sweating (B) temperature sensation (C) proprioception (D) touch sensation (E) pain sensation 80. A patient has palpitations, flushing, and light-headedness after gingival injection of a local anesthetic. This reaction is most likely caused by (A) epinephrine in the local anesthetic (B) local anesthetic allergy (C) para -aminobenzoic acid allergy (D) methylparaben reaction (E) vasovagal reaction

98. Which of the following will most closely mimic the effects of stellate ganglion block? (A) Axillary perivascular block with 25 ml of 1.5% lidocaine (B) Cervical nerve block at C2-C5 with 2 ml of 1.5% lidocaine (C) Supraclavicular block at the level of the first rib with 25 ml of 1.5% lidocaine (D) Block of the median, radial, ulnar, musculocutaneous, and intercostobrachial nerves (E) Excision of thoracic sympathetic ganglia Tl-T4 101. A 28-year-old woman receives a lumbar epidural anesthetic for uncomplicated labor and delivery. During removal of the catheter, 1 cm breaks off and remains in her back. After informing the patient, the most appropriate management i s (A) no intervention unless symptoms occur (B) prophylactic antibiotics (C) epidural corticosteroids (D) dye contrast study of the epidural space (E) neurosurgical exploration 115. Following paravertebral blocks of the second through fifth intercostal nerves, numbness is noted on the medial aspect of the ipsilateral arm. The most likely cause is

(A) anesthesia of the intercostobrachial nerve (B) anxiety-induced hyperventilation (C) injection into a dural cuff (D) intravascular injection of local anesthetic (E) partial block of the brachial plexus 119. Percutanous cordotomy is being considered for a patient with severe pain that has persisted for three months after amputation of the arm for osteogenic sarcoma. Which of the following statements is true? (A) An effective cordotomy will produce motor block (B) A series of stellate ganglion blocks will provide permanent relief (C) Cordotomy must be performed with the patient awake (D) Cordotomy will effectively relieve phantom limb pain (E) Spinal opioids are an alternative treatment of this pain 122. Nonsteroidal analgesics such as aspirin decrease pain by (A) decreasing axonal transmission of pain stimuli through a local anesthetic mechanism (B) directly competing with substance P for receptor occupancy (C) inhibiting enzymatic breakdown of met-enkephalin (D) inhibiting production of prostaglandin E, (E) potentiating the action of endorphin in the substantia gelantinosa of the spinal cord 134. A 31-year-old man underwent uneventful epidural anesthesia for arthroscopy of the knee and meniscectomy. Twenty-four hours later, he still has flaccid paralysis in both legs and pain in the lower back. The most likely cause i s (A) adhesive arachnoiditis (B) anterior spinal artery thrombosis (C) epidural abscess (D) epidural hematoma (E) transverse myelitis 139. In an infant, spinal anesthesia to a sensory level of T8 is achieved with tetracaine administered at the L2-3 interspace. Compared with spinal anesthesia to the same sensory level in an adult, this anesthetic is associated with a (A) higher risk for neurotoxicity (B) higher risk for systemic toxicity (C) lower risk for spinal cord injury (D) more significant decrease in blood pressure (E) shorter duration of action 141. Which of the effects of epidural morphine is most resistant to intravenous naloxone? (A) Analgesia (B) Nausea (C) Pruritus (D) Respiratory depression (E) Urinary retention 150. To eliminate all pain during the second stage of labor, a lumbar epidural block must extend from (A) T6 to Ll (B) TlO to Ll (C) TlO to Sl (D) TlO to S4 (E) Ll to S5 155. A 60-year-old man undergoes transurethral resection of a bladder tumo r in the lithotomy position with spinal anesthe-sia.

During the procedure the surgeon reports that the patients right leg is jumping. This movement is most likely caused by stimulation of which of the following nerves? (A) Femoral (B) Lateral femoral cutaneous (0 Obturator (D) Pudendal (E) Sciatic 157. Which of the following is the most likely finding in the affected extremity during the acute stage of reflex sympathetic dystrophy? (A) Dermatomal distribution of pain (B) Hypesthesia (C) Increased growth of hair and nails (D) Localized edema (E) Radiographic evidence of osteoporosis 164. The primary determinant of the duration of an intravenous regional block sufficient to provide surgical anesthesia of an upper extremity is (A) duration of tourniquet inflation (B) capacitance of the venous system of the extremity (C) local anesthetic agent injected (D) volume of anesthetic solution injected (E) method of exsanguinating the arm 165. Compared with morphine 5 mg administered epidurally at T12, morphine 5 mg administered intravenously is associ-ated with (A) greater incidence of urinary retention (B) less intense analgesia (C) less nausea and vomiting (D) longer duration of analgesia (E) longer time to maximum analgesia 166. An obese, 75-year-old woman is scheduled for open reduction of a left forearm fracture. Thirty minutes after successful interscalene block using 20 ml of 2% lidocaine, she becomes dyspneic. The dyspnea is most likely related to (A) cervical epidural block (B) cervical sympathetic block with bronchospasm (C) chylothorax (D) elevation of the left hemidiaphragm (E) recurrent laryngeal nerve block

You might also like