You are on page 1of 20

Peter Nagele and Michael Hpfl

73
Key Points

Anesthesia and Prehospital Emergency and Trauma Care

1. Anesthesiology was instrumental in creating critical care and emergency medicine. In many countries, worldwide anesthesiology departments are responsible for prehospital emergency and trauma care. Physician-based emergency medical services (EMS) systems exist in many European countries, whereas paramedics are the mainstay of EMS in the United States. 2. The basics of prehospital care follow the general form of basic life support (BLS) and advanced life support (ALS), depending on the nature of the EMS system in place for a given region or country. 3. In major trauma, prehospital care must aim at limiting the time spent on the scene, controlling hemorrhage, and expediting transport to a trauma center, often ideally via rescue helicopter. Prehospital fluid resuscitation for major trauma is controversial and patients with penetrating torso injuries and hemorrhagic shock might benefit from limited fluid resuscitation, in particular in urban settings. Prevention of the lethal triad in trauma of hypothermia, acidosis, and coagulopathy is paramount.

4. In acute coronary syndrome (ACS), achieving rapid reperfusion of the ischemic myocardium is the main goal. Opioids (morphine), oxygen, nitrates, and aspirin (MONA) are the main components of prehospital therapy. Fibrinolysis for ACS has been used with much success in the prehospital setting but requires very close supervision by EMS physicians. 5. Prehospital endotracheal intubation and rapid-sequence induction (RSI) have been shown in more than 15 studies to be associated with increased mortality and poorer neurologic outcomes, in particular after traumatic brain injury. It appears that what anesthesiologists consider the standard of care in the operating room may not be the ideal approach for less experienced EMS providers. Perhaps reappraising intubation and RSI in favor of alternate airways such as the laryngeal mask airway or laryngeal tube can improve outcomes. 6. Medical simulation is rapidly becoming the future standard for teaching and training prehospital EMS personnel in the management of complex emergency scenarios.

Unbeknown to many current practitioners, anesthesiologists not only serve as the backbone of physician-staffed prehospital emergency medical services (EMS) in many countries1-3 but also were instrumental in the creation of modern prehospital emergency medicine.4,5 In several countries, emergency medicine is considered the fourth pillar of our specialty besides anesthesiology, critical care, and pain therapy. After World War II, the inadequacy of the civilian ambulance service became all too apparent for wartime experienced physicians, as motor vehiclerelated injuries and deaths increased sharply and ambulances provided only transport to the hospital. In Germany, this led to a call to bring the doctor to the patient rather than the more traditional converse arrangement.1,5 As a consequence, the first surgeon-staffed ambulance, the

Klinomobil, was implemented in 1957 in Heidelberg and functioned as a mobile operating room,5 a concept that was soon abandoned in favor of physician-staffed ambulances, sometimes referred to as mobile ICUs.6 Similar developments occurred throughout Europe, for example, in Prague, Mainz, Munich, Moscow, and several Scandinavian countries where physicians began to staff ambulances to respond to motor vehicle crashes and other accidents.3,4 In Belfast, Ireland, Frank Partridge was the first to implement a mobile coronary care unit (CCU) that brought a defibrillator to prehospital patients in cardiac arrest due to ventricular fibrillation.7 Over the next decades, several European countries, such as France, Germany, Austria, and Norway, implemented nationwide physician-staffed ambulance services that were often complemented by physician-staffed helicopter
2313

2314 Adult Subspecialty Management

emergency medical services (HEMS).1-3,8,9 Anesthesiology is currently the dominant medical specialty providing EMS physicians in Europe, and many anesthesiology departments are in charge of staffing ambulances and helicopters with physicians.3,10 In the United States, a different development took place. For various reasons, physician-staffed ambulances were not considered an option, so specifically trained emergency medical technicians (EMTs) and paramedics became the mainstay of EMS in North America. A landmark event for the creation of modern EMS systems was the publication of the report Accidental Death and Disability: The Neglected Disease of Modern Society, commonly referred to as The White Paper, published by the National Academy of Sciences in 1966.11 Among several remarkable physicians, two anesthesiologists, Peter Safar, in Pittsburgh, the father of modern cardiopulmonary resuscitation (CPR), and Eugene Nagel, in Miami, were most influential in creating the first paramedic-based EMS system in the United States in the 1960s. Before emergency medicine became its own medical specialty, anesthesiologists were often involved in prehospital emergency care based on their unique expertise in resuscitation, airway management, and critical care. This involvement has dwindled over the past 2 decades, although it appears that anesthesiology could still play an important role in prehospital emergency care in the United States based on what the recent, widely cited Institute of Medicine report Emergency Medical Services: At the Crossroads identified as the major problems in prehospital care, such as prehospital airway management, fluid resuscitation, rapid-sequence induction (RSI), and pain therapy.12 In this chapter, we provide a concise overview of prehospital emergency and trauma care with a focus on the important role of anesthesiology.

Table 73-1 Skill Set for BLS and ALS Providers and Prehospital Emergency Physicians BLS Basic patient evaluation Vital signs and basic monitoring (noninvasive blood pressure, pulse oximetry, heart rate) Hemorrhage control via direct pressure or pressure dressing Cardiopulmonary resuscitation (BLS) with use of automatic external defibrillator Bag-valve-mask ventilation Oral and nasal airway Oral suctioning Administration of oxygen Basic patient rescue, positioning, and transport skills Cervical spine stabilization Fracture splinting Assisting a patient taking prescribed medication (e.g., albuterol inhaler) ALS Venous access and fluid therapy Administration of select drugs* Advanced monitoring (ECG, capnography) Cardiopulmonary resuscitation (ACLS) Manual defibrillation Basic 12-lead ECG interpretation External pacing Endotracheal intubation* (without drugs; some EMS systems allow rapid-sequence induction) Cricothyroidotomy Sedation* Needle decompression of tension pneumothorax Knowledge of mass casualty incident management and triage Knowledge of biological, chemical, and nuclear agents Prehospital Emergency Physician Unlimited administration of emergency drugs Use of narcotics and controlled substances (e.g., fentanyl) Prehospital rapid-sequence induction and anesthesia Advanced airway management Chest tube placement Advanced 12-lead ECG interpretation Advanced techniques (ultrasound, noninvasive ventilation) Emergency surgery Can assume role of medical incident commander in mass casualty event
*Not in all ALS-level EMS systems allowed. ALS, advanced life support; BLS, basic life support.

Organizational Models of Prehospital Emergency Care


Levels of Care
Originally, the terms basic and advanced life support were only used to describe the two different levels of care for CPR (see Chapter 97) but were later expanded to all emergency situations, including trauma and pediatric emergencies. Currently, they denote the level of care in an organized EMS system and are usually divided into BLS and ALS. First responders, such as police officers and firefighters, are not part of an EMS system and they provide first aid, operating below the formal BLS level. Basic Life Support Prehospital care at the BLS level is usually provided by EMTs and does not include advanced and invasive techniques and skills.13 BLS is focused on the classic ABCs (Airway, Breathing, Circulation) and includes simple airway maneuver, such as chin lift or jaw thrust, oral and nasal airways, bag-valve-mask ventilation, hemorrhage control, and administration of oxygen (Table 73-1). During CPR, the use of an automated external defibrillator is considered a BLS skill. BLS-only ambulances are often found

Anesthesia and Prehospital Emergency and Trauma Care 2315

73

in rural areas where the incidence of emergencies is low and volunteer EMTs frequently staff ambulances. Advanced Life Support ALS is usually provided by paramedics and flight nurses with a rigorous educational background and includes the use of advanced and invasive techniques (see Table 73-1). In many countries, ALSlevel providers are allowed to work semi-independently with often only indirect physician control through the use of standing orders and protocols issued by the EMS medical director. The use of drugs is tightly regulated, and physician permission from the emergency department must be sought in cases where drugs outside the approved list are to be administered. Most urban EMS systems use ALS units to provide prehospital care. In EMS systems without EMS physicians responding directly to the scene, the scope of practice of ALS-level providers is usually greater and they are allowed to work more independently. Combat medics in the military frequently operate at the ALS level. EMS physicians, flight nurses, and critical care EMTs provide a level of care above and beyond ALS.

Models
Worldwide, a high degree of variability exists as to how prehospital care is organized and staffed. It would be beyond the scope of this chapter to describe all the nuances of the different EMS systems worldwide; rather, we will outline the differences between the two main models, which primarily differ by their level of direct physician involvement in prehospital emergency care: EMT-based models versus physician-based models. These two models are also known as the U.S. model and the Franco-German model.14 Although controversy exists which model provides superior patient care and outcomes, the design of the EMS system is dictated by tradition and fiscal constraints. Single-Tiered Versus Multi-Tiered EMS Systems If an EMS system is composed only of units providing the same level of careBLS or ALSit is referred to as a single-tiered system. The advantage of a single-tiered EMS system is that it eliminates overtriage or undertriage by emergency medical dispatch. In rural areas with a low frequency of emergency calls, a single-tiered EMS systemoften composed of volunteer EMTs at the BLS levelis the only possible model; and in dense urban areas with a high frequency of emergency calls, an all-ALS EMS system might be advantageous. The disadvantage of single-tiered all-ALS EMS systems is that the highly trained paramedics will not be exposed to the same number of high priority situations, making it difficult to maintain their skills and experience. Multi-tiered EMS systems use priority dispatch to determine the adequate level of response; it would dispatch an ALS unit to a patient with myocardial infarction but a BLS unit to a patient with a broken wrist. EMT- and Paramedic-Based EMS Systems In North America, EMTs provide organized prehospital care.15 Only rarely will physicians be directly involved in out-of-hospital patient care on the scene.16 By U.S. law, physicians are required in most states to oversee the medical operations of EMS systems, either online through direct communication or offline as EMS medical directors responsible for developing and implementing

standardized treatment protocols. Besides developing standard operating procedures and approving new medical devices, EMS medical directors are responsible for quality control and improvement and are an integral part of EMT education.17 In the United States, EMS medical directors are often emergency physicians working in the emergency department of the local hospital. In general, three levels of EMTs can be distinguished by their educational background and experience in the United States: EMT-basic (EMT-B), EMT-intermediate (EMT-I), and EMTparamedic (EMT-P). EMT-basic providers usually spend between 80 and 120 hours of training and are only allowed to perform basic life support (BLS) skills. The basic level is the minimum level required to work in an organized ambulance service. Often, EMT-basic providers staff nonemergent ambulance transport but are also used to respond to nonlife-threatening situations. EMTI providers are regulated differently among the individual states, and their skills range between BLS and ALS. Often, EMT-I providers can be found in rural areas where it would be too expensive to employ paramedics only but where more advanced skills than those of an EMT-B level are needed. Most states do not allow EMT-I providers to perform endotracheal intubation or administer drugs, except very few deemed safe, such as albuterol for asthma, but allow the placement of an intravenous (IV) line. Paramedics (EMT-P), on the other hand, are the most advanced EMTs. They undergo intensive 1- to 2-year training in all different aspects of advanced prehospital emergency care. Their scope of practice is ALS, and they are allowed to perform advanced techniques such as endotracheal intubation, administration of selected drugs, and manual defibrillation. Most developed countries have training requirements for their EMT work force similar to the U.S. model. Physician-Based EMS Systems Germany and France represent the two most well-known examples of countries where doctors are an integral part of prehospital emergency care.14 These integrated EMS systems combine BLSlevel ambulances with physician-staffed ALS ambulances and helicopters in a noncompetitive but complementary fashion. Physician-based EMS systems are always part of a multi-tiered system and often use what is referred to as a rendezvous system to dispatch two units simultaneously to the scene, one unit staffed with EMTs (BLS or ALS level) and one ambulance carrying the doctor. Often, the physician-staffed vehicle will be a nontransport vehicle aimed to provide quick response to emergencies without having to move a large and slow truck through traffic. EMS physicians are licensed physicians from various medical specialties who undergo a specialized training in prehospital emergency care. The level of mandatory education varies from country to country. In many countries with direct physician involvement in prehospital care, anesthesiologists compose the largest fraction of physicians and often assume the role of chief emergency physician/medical director.18,19 Helicopter EMS In many countries worldwide, helicopter EMS (HEMS) continues to expand as an integral component of EMS. Although being mostly privately organized and funded in the United States, several European countries have implemented a nationwide HEMS network that is strategically distributed to be able to respond to emergency calls from every location in the country. HEMS have historically been mostly associated with rapid

Section V Adult Subspecialty Management

2316 Adult Subspecialty Management

patcher but also to the hospital and other emergency response teams such as fire department, police, and helicopters.

Basic Techniques and Skills


Patient Examination and Initial Assessment on the Scene
Emergency patients show a wide range of symptoms, challenging the prehospital care team to obtain a tentative diagnosis quickly and start a specific treatment. In contrast to the emergency department, fewer diagnostic tools are available in the prehospital setting. The classic approach of look, listen, and feel is still the most important aspect of the patient examination. On approaching the patient, a rapid appraisal of the scene should assess potential hazards that could endanger patient and prehospital personnel. The initial assessment should take into account the surrounding circumstances such as environmental factors and should focus on the vital signs of the patient using a systematic approach.23 Visual impressions from the scene help to understand the circumstances of an accident or medical emergency, such as intoxication. The medical history of the patient can guide the presumptive diagnosis. In case the patient is unresponsive, relatives or bystanders must be queried. The most common chief complaints encountered in the prehospital setting are dyspnea/respiratory distress, cardiac/circulatory dysfunction, altered level of consciousness, and trauma.

Figure 73-1 Train accident in Wampersdorf, Austria. Mass-casualty incident with 6 fatalities and 17 injured patients. In addition to more than 50 EMS personnel and EMS physicians on the ground, 5 EMS-physicianstaffed helicopters provide invaluable EMS-physician level backup and transport capacity.

response to and fast transport from the scene of an accident, particularly from remote rural accident scenes. However, HEMS is now considered more and more as a vital ALS-level backup for all emergency situations in rural areas lacking local ALS-level response.20 EMS helicopters are usually staffed with experienced personnel, either paramedic and flight nurse or EMS physician.9,21 In mass-casualty incidents, all helicopters from a large area can be concentrated at the scene of the accident within a short period of time and provide several highly trained ALS teams to support local rescue operations and transport the victims quickly to trauma centers (Fig. 73-1). Another area where HEMS have become the most important means of evacuation is in mountain rescue operations. Pioneered in the Alps, helicopter evacuation by ALS teams specially trained in mountain rescue techniques is now considered the standard of care in countries such as Switzerland, Austria, Germany, and France (Fig. 73-2).

Basic Evaluation
Vital Signs The first quick survey employs the A-B-C-D scheme well known by ALS.24 The first step is to use acoustic and tactile stimuli to rule out unconsciousness in the emergency patient. The next step

Communication
Starting with the emergency call, communications is a vital aspect of prehospital emergency care. The implementation of a universal emergency telephone number (911 in the United States and Canada, 112 in several European countries, 000 in Australia, and 999 in the United Kingdom)17 was a milestone in prehospital care, and nowadays the location of the caller can be automatically identified by the system except if calling from a cellular phone. Emergency medical dispatchers will determine the type and priority of the response based on standardized questionnaires and will provide invaluable prearrival instructions to the caller to start potentially lifesaving interventions before the ambulance arrives. For example, dispatcher-assisted CPR has been shown to improve survival after cardiac arrest.22 Two-way radio communication has long been an indispensable part of EMS communications, allowing not only fast communication between ambulance and dis-

Figure 73-2 In mountain rescue, helicopter EMS systems have become the standard of care for evacuating patients from alpine accidents and emergencies. Here, a rescue helicopter supports a complex mock rescue operation from a ski lift in alpine Austria.

Anesthesia and Prehospital Emergency and Trauma Care 2317

73

is to identify airway (A) obstruction and, if necessary, to secure the airway. In trauma patients, manual in-line stabilization of the cervical spine should be performed at this point. While maintaining an open airway, breathing (B) is assessed by look, listen, and feel.23 Respiratory rate, efficacy, work and pattern of breathing are also assessed. If indicated, supplemental oxygen or assisted ventilation will be administered to support oxygenation and ventilation. Circulation (C) is initially checked by feeling the pulse centrally at the carotid artery and can be extended to feel pulses proximally and distally on the extremities. The pulse should be assessed for rate, quality, and regularity. To complete C, the capillary refill time should be measured. The level of responsiveness and consciousness (D, disability) should be determined according to the Glasgow Coma Scale (GCS) and a quick pupil check performed to determine size and reaction to light. The GCS is a reliable prognostic tool for head trauma25,26 and other forms of nontraumatic altered level of consciousness.27 For children younger than the age of 3 years, the verbal component of the GCS score must be adjusted. In case of trauma, exposure and environment (E) are assessed to complete the primary survey. At this point, the patient should be checked from head to toe, and his or her back should be evaluated after a log-roll maneuver.28 AuscultationPalpationPercussion Auscultation is a versatile and useful tool in most prehospital settings, limited primarily by ambient noise in the environment. Palpation, also an important tool, may be more useful in such settings. In trauma, the entire body can be examined by palpation only. During the primary and secondary survey in the field, the rescuer has to rely solely on palpation. Percussion is specifically helpful to detect abnormal air or gas collections, such as tension pneumothorax and large amounts of free intra-abdominal air. Capillary Refill Time Capillary refill time is a quick and reliable parameter to detect hypovolemia; 5% dehydration can be detected by capillary refill time with 87% sensitivity and 82% specificity.29 After compression of the skin for at least 5 seconds, the time is measured until normalization of skin color occurs. Normal findings are less than 2 seconds. Capillary refill time may be especially useful to detect significant dehydration in children.30

Arterial Blood Pressure Noninvasive blood pressure measurement is the most common form of blood pressure monitoring in the prehospital setting. Automatic or manual cuffs are mostly used and use an oscillo metric technique that has known limitations.36 The precision and accuracy of noninvasive oscillometric blood pressure monitoring is reduced by movement during transport. During interhospital transfer of a critically ill patient, intensive care unit (ICU) monitoring with invasive blood pressure monitoring is frequently used. Electrocardiography ECG monitoring has two major goals in the prehospital setting. A quick 3- or 4-lead ECG or an ECG rhythm obtained from defibrillator paddles is sufficient for a rapid ECG rhythm diagnosis. This is of importance for treatment decisions during ALS in cardiac arrest to decide if a shockable rhythm is present. On the other hand, to identify myocardial ischemia or infarction, a 12lead ECG is needed. It is more cumbersome to apply and requires more experience for interpretation. Prehospital ECG allows for a focused prehospital therapy in acute coronary syndrome37 and has been shown to shorten the door-to-reperfusion time.38-40 This benefit can be further improved by transmission of the ECG to medical oversight.41 Temperature Monitoring Oral, esophageal, and some tympanic thermometers42-45 are reliable tools in the field given the more extreme weather conditions in which they must operate. The most important aspect of measuring temperature on the scene is to identify hypothermia. Special temperature probes are required to assess severe hypothermia (<28C) because most commercially available instruments become unreliable at these low temperatures. Trauma patients, patients at the extremes of age, disabled patients (stroke, coma), and intoxicated patients are especially vulnerable to hypothermia, particularly during winter months. Even relatively mild outside temperatures can result in hypothermia with long exposure times.46,47 Capnography Colorimetric, main- or side-stream capnographs are primarily used in the prehospital setting for identification of the correct positioning of the endotracheal tube.48-50 With a single measurement the correct tube position can be identified. Capnography became the standard for identifying the correct tube position in addition to auscultation and identifying chest movements. Continuous monitoring of end-tidal CO2 allows the monitoring of hemodynamics and the success of resuscitation.51 In cardiac arrest, low CO2 levels are a predictor for fatal outcome.52 Hyperventilation may also contribute to poor outcome53 and may be preventable by controlling ventilation with the aid of end-tidal CO2 monitoring.54,55 Significant cardiorespiratory failure may limit the usefulness of capnography owing to significantly reduced pulmonary perfusion. Prehospital Blood Tests Despite proposals for prehospital point-of-care testing,56 only blood glucose testing is routinely used. Troponin and blood gas analysis are examples of tests that may have potential utility.53,57

Section V Adult Subspecialty Management

Monitoring and Equipment


In a similar development compared with standard monitoring in operating room and intensive care, monitoring in the prehospital setting is now considerably standardized with noninvasive blood pressure monitoring, electrocardiography (ECG), and pulse oximetry regarded as the classic trifecta. Capnography and temperature monitoring have begun to supplement the diagnostic and monitoring armamentarium in the field. Pulse Oximetry After the first prehospital uses in the 1980s,31 pulse oximeters have become the universal, easy to attach and use monitor for the prehospital patient. The improvement of the devices led not only to smaller machines but also to pulse oximeters with better quality even in motion and the cold.32,33 Additionally, new features such as detection of methemoglobin and carboxyhemoglobin34,35 can help diagnose conditions such as carbon monoxide poisoning.

2318 Adult Subspecialty Management

Ultrasonography Small and portable ultrasound machines have been developed that are applicable in the prehospital setting. Diagnostic ultrasound requires a physician on the scene (or online through medical oversight if adequate transmission of the image can be achieved) but offers very interesting possibilities. Focused abdominal sonography for trauma (FAST) is a rapid and reliable technique even in inexperienced hands in the emergency department and could be used in the field.58 It has the potential to improve the triage and management of trauma patients.59 During resuscitation from cardiac arrest, cardiac activity is a predictor for return of spontaneous circulation (ROSC).60 Nevertheless, diagnostic prehospital ultrasonography is still considered experimental and may interfere with management priorities.

which is rapidly becoming the standard alternate airway in the field.75 Ventilation in the prehospital setting is most commonly achieved by bag-mask-valve ventilation. Although fairly common in Europe, only a few EMS services in the United States currently use portable field ventilators to ventilate intubated patients. The major disadvantage of relying on bag-mask-valve ventilation is that minute ventilation and tidal volumes are uncontrolled, often leading to significant hyperventilation that has been shown to be detrimental in patients in cardiac arrest53,80,81 and with head trauma.82-88 Airway Management at the ALS Level Endotracheal intubation is considered the standard of care in prehospital airway management and a standard intervention for paramedics and EMS physicians because it provides a definitive airway and protection from aspiration of gastric contents. Compared with standard endotracheal intubation in the operating room, several differences exist. First, most prehospital intubations (50%-70%) are performed in patients with cardiac arrest, obviating the need for any anesthetic induction or muscle relaxation.89 Most other intubations in spontaneously breathing patients require sedation or induction of general anesthesia, most commonly performed as RSI. Particular emphasis should be placed on adequate preoxygenation because these critically ill patients have little to no oxygen reserves,90 as well as on maintaining cervical in-line stabilization in trauma patients. Oral, direct laryngoscopic, and tracheal intubation while maintaining cricoid pressure (Sellicks maneuver) is the recommended approach. Endotracheal intubation in the field requires confirmation by auscultation and end-tidal CO2 monitoring. Adverse intubating conditions in the field lead to more frequent incidences of difficult endotracheal intubations,70 including higher incidences of multiple attempts91 and of unrecognized esophageal or endobronchial intubation. This occurs even by experienced EMS physicians and anesthesiologists and is associated with a high mortality rate.71 However, experienced physician airway managers and, in particular, anesthesiologists have more frequent success rates and lower complication rates after prehospital endotracheal intubation.92-94 A significant portion of failed field intubations results from inadequate operator training or experience. Less than one half of field intubation failures were remedied in the emergency department by the use of neuromuscular blocking agents.95 Because of the higher complication rate, every prehospital provider performing advanced airway management must be trained also in alternate airway management, such as Combitube, King-LT, or laryngeal masks. A special training for managing the unanticipated difficult airway according to adapted international guidelines should be required.96,97 Surgical airway access such as cricothyroidotomy is considered the last resort but can be lifesaving in patients when every other airway technique fails.98

Prehospital Airway Management and Ventilation


Early and adequate airway management and ventilation in the out-of-hospital setting are lifesaving interventions that define prehospital emergency and trauma care (see Chapter 50). For anesthesiologists who perform advanced airway management on a daily basis it is often difficult to comprehend why airway management is such a contentious issue in prehospital emergency medicine, particularly in the United States.61-67 The reasons are many. The major difference from the operating room environment, as an EMS saying goes, is that there is no easy airway in the field. Challenges include nonstandard environments; difficult access to the patients head; bad weather and lighting conditions; and blood, vomitus, secretions, and foreign bodies obstructing the airway. In a patient who is critically ill with limited oxygen reserves and a higher baseline oxygen consumption all of these factors contribute to a significantly more difficult prehospital airway management compared with the operating room, even for an experienced anesthesiologist.68-72 Airway Management at the BLS Level Supplemental oxygen is the most important component in prehospital airway management, hence the saying The patient does not die because of a lack of an endotracheal tube but because of lack of oxygen. EMTs are trained and qualified to administer supplemental oxygen, often via a non-rebreather facemask allowing for a significantly higher Fio2. Furthermore, EMTs are allowed to use nasopharyngeal or oropharyngeal airways to maintain a patent airway and to ventilate a patient by bag-mask-valve ventilation. The use of the Combitube (or Easytube), a double-lumen tube with a proximal and a distal balloon that was initially developed for ventilation during CPR, is considered a BLS skill.73,74 The use of other alternate supraglottic airways, such as the King-LT (laryngeal tube) or laryngeal masks, is increasingly recommended at the BLS level if adequate training and experience are provided.75 These alternate airways have been shown to be safe and easy to apply and have excellent success rates when used by EMTs and paramedics.76-79 The limitations of these alternate airways are a disability to ventilate if misplaced in the larynx and lack of airway protection compared with a cuffed endotracheal tube. It is of note that many anesthesiologists who do not work in EMS medicine are currently not familiar with many new alternate airway devices used in prehospital care, such as the King-LT,

Sedation, Anesthesia, and Pain Control in the Prehospital Setting


In this section, the focus will be on generally accepted strategies for providing prehospital sedation, induction, and maintenance of anesthesia and pain control (see Chapters 26 and 27). Drugs

Anesthesia and Prehospital Emergency and Trauma Care 2319

73

commonly used in the field will be discussed with an emphasis on the differences from use in the hospital. There is hardly any area of greater controversy than RSI in the field. Drugs Commonly Used for Prehospital Sedation, Anesthesia, and Pain Control In general, not all drugs routinely used in the operating room for general anesthesia are safe in the prehospital setting.99 Nonanesthesia personnel such as paramedics and nurses are much less experienced than anesthesiologists in the use of these drugs, and critical patients often require a more cautious approach and dosage titration. Therefore, drugs used for analgesia, sedation, and anesthesia in the field should have the following desirable properties100: Wide safety margineven for inexperienced providers Hemodynamic stability Minimal respiratory depression Ease of administration via different routes (IV, IM, IO, rectal) Rapid onset Minimal side effects Benzodiazepines, such as midazolam (0.05-0.2mg/kg IV) and diazepam (3-10mg IV), are well suited for sedation in adult and pediatric patients and can be administered via several different routes; however, the complete lack of analgesic properties usually requires the concurrent administration of a potent analgesic in patients with severe pain. Ketamine, on the other hand, is closer to an ideal agent for sedation and analgesia (0.5-2mg/kg IV) and in higher doses for induction of anesthesia (2-5mg/kg IV) in the field.101 It provides potent analgesia and sedation, maintains hemodynamic stability in patients with hypovolemic shock, is a potent bronchodilator, and causes minimal respiratory depression. The use of ketamine in severe head trauma is controversial: some make a case for its use if ventilation is controlled.102 Owing to its hallucinogenic properties, a benzodiazepine should be coadministered. The new isoform, S(+) ketamine, is twice as potent but has otherwise a similar profile as a racemic drug. Ketamine can be administered intramuscularly (IM) and is ideally suited to provide analgesia and sedation for entrapped trauma patients during extrication (4-12mg/kg IM). Etomidate is rarely used for sedation but is often used as an anesthetic induction agent for RSI in the field (0.2-0.3mg/kg IV). It maintains hemodynamic stability but frequently produces myoclonus, rendering intubating conditions less favorable. Furthermore, etomidate may cause adrenocortical suppression, but this is controversial and has not been sufficiently studied in the EMS setting.103,104 Propofol has become the anesthetic induction agent of choice in the operating room but has several disadvantages for use in the field. It causes significant hemodynamic compromise and respiratory depression, both highly undesirable properties, especially for inexperienced nonanesthesiologists. Barbiturates, such as thiopental, share similar pharmacologic properties and side effects with propofol and should be used only by experienced providers and only in selected patients, for example, in those with status epilepticus. The prehospital muscle relaxant of choice is still succinylcholine (1-1.5mg/kg IV) during RSI,105 although rocuronium has been successfully used by anesthesiologists in EMS.106 For pain control in the prehospital setting, opioids are useful owing to the rapid onset and the ability to control severe pain.107 Morphine is con-

Box 73-1 Prehospital Rapid-Sequence Induction 1. Preparation of necessary equipment (suction, oxygen, laryngoscopes, endotracheal tube, Ambu bag plus masks, stylet) 2. Preoxygenation with 100% oxygen (at least 3 minutes) 3. Pretreatment (optional) with small defasciculation dose of a nondepolarizing neuromuscular blocking agent, such as vecuronium (0.01-0.015mg/kg IV) 4. Pretreatment (optional) with sedative (midazolam) and/or opioid analgesic (fentanyl) 5. Induction agent (etomidate, 0.3mg/kg IV, or ketamine, 2-3mg/kg IV) plus succinylcholine (1.0-1.5mg/kg IV) 6. Cricoid pressure (Sellicks maneuver), plus manual in-line stabilization for trauma patients 7. Direct laryngoscopy with oral intubation (use of stylet is recommended) 8. Confirmation of correct endotracheal tube placement: bilateral auscultation plus end-tidal CO2 confirmation

Section V Adult Subspecialty Management

sidered the standard of care in acute coronary syndrome (0.10.2mg/kg IV), but in trauma other opioids, such as fentanyl (1-3 g/kg IV), piritramide (0.1-0.2mg/kg IV, currently not available in the United States but widely used in Europe), and sometimes tramadol (50-100mg bolus IV) are more frequently used. Nonopioid analgesics are infrequently used in the prehospital setting, although they offer a potential additive analgesic effect with opioids. Induction of Anesthesia The management of prehospital induction of anesthesia will nearly always be performed as RSI, also known as drug-assisted intubation (DAI) in EMS literature (see Chapter 50). The goal is twofold: (1) the provision of a definitive airway and ventilation control and (2) the provision of potent analgesia and sedation. The actual principles of RSI do not differ from RSI in the emergency department, operating room, or ICU, only in the environment in which it is applied, as discussed earlier.63 A standard RSI sequence consists of the steps presented in Box 73-1. Extremely important in the prehospital setting is to have a backup plan, if intubation fails. Bag-mask-valve ventilation, placement of an alternative airway such as a laryngeal mask airway or intubating laryngeal mask airway, or Combitube are acceptable options. EMS providers must understand that prehospital RSI is not trivial and is associated with significant morbidity and mortality if performed improperly.108 Life-threatening hypoxemia will ensue if intubation is prolonged and ventilation is difficult (Fig. 73-3).

Prehospital Trauma Care


The management of a severely injured patient is one of the biggest challenges in prehospital care. Trauma represents a leading cause of death and disability among all ages but disproportionally affects the younger age groups. The potential for saving lives is probably nowhere more prevalent in EMS than in major trauma, which was also historically the driving force behind the creation of modern EMS systems. The challenges for EMS providers

2320 Adult Subspecialty Management 100 80 60 40 20 0 0 112 224 336 448 560 672 Seconds 784 896 1,008 1,120 1,232 1,344 (2) (1)
SpO2 % (1) ETCO2 mm Hg (2) Pulse rate, beats/min (3)

(3)

Figure 73-3 Potentially life-threatening desaturation and bradycardia after out-of-hospital rapid-sequence induction and endotracheal intubation. (From Dunford JV, etal: Incidence of transient hypoxia and pulse rate reactivity during paramedic rapid-sequence intubation. Ann Emerg Med 42:721-728, 2003.)

responding to major trauma begin with assessment of safety at the scene, scene assessment (number of injured patients, mechanism of injury, injury severity, difficult weather and lighting conditions), and coordination with other first-responder services (police, fire department). Significant pressure exists to minimize on-scene and transport times, based on the golden hour and platinum 10 minutes concept.109 The most commonly used guideline for prehospital trauma care is Prehospital Trauma Life Support (PHTLS) developed by the National Association of Emergency Medical Technicians (NAEMT) in cooperation with the Committee on Trauma, American College of Surgeons.28 PHTLS is based on the Advanced Trauma Life Support (ATLS) concept that is used as standard for in-hospital trauma care.110

Patient Evaluation on the Scene and Initial Management


Before turning the attention to the trauma patient, the EMS crew must ensure that the accident scene is safe. Crew protection is paramount, in particular in hazardous accidents involving fire, chemical agents, or electricity or on busy highways. Motor vehicle accidents often involve more than one patient; therefore, the EMS team must perform a quick but thorough scan of the entire accident scene and triage all victims. The overriding goal of the initial patient assessment on the scene is to identify life-threatening injuries. This usually follows the well-known A(irway), B(reathing), C(irculation), D(isability), and E(xposure/Environment) algorithm. Treatment is immediately initiated if a life-threatening injury is identified (Treat what kills first). This often involves emergent airway management to protect the airway or hemorrhage control for major bleeding. Supplemental oxygen via a nonrebreathing facemask should be started on the scene, as well as standard monitoring (ECG, pulse oximetry, blood pressure, capnography) applied. The goal is to maintain Spo2 above 90%. If the patients respiratory status deteriorates or if he or she is unable to protect the airway (due to unconsciousness, intoxication, hypo perfusion, or direct trauma), establishing a definitive airway is indicated. Endotracheal intubation in the field with or without RSI by paramedics has been intensely scrutinized over the past decade,62,68 especially in traumatic brain injury,111 and it is currently unclear if field airway management with RSI by paramedics decreases or contributes to trauma mortality.67 It appears that experience of the provider in endotracheal intubation and RSI is a strong predictor for successful prehospital airway management,112 and therefore it should not come as a surprise that success

rates are higher and complication rates lower if anesthesiologists perform airway management in the prehospital setting.93 Hemorrhage control is probably the most critical aspect of prehospital trauma care, particularly in combat casualty care. A large fraction of traumatic deaths within the first hour can be attributed to exsanguination.113 The recommended first step in controlling the bleeding is direct pressure to the wound, including application of an appropriate pressure dressing. Compressing the major artery proximal to the wound at pressure points is the next step, and tourniquets should only be used as a last resort in civilian trauma.114 In military trauma, the use of tourniquets has recently surged again in Iraq and Afghanistan and has been accredited with having saved many lives.115,116 Several different hemostatic wound dressings have been advocated and used in combat casualties suffering from severe hemorrhage117; their use in civilian EMS is currently not well defined.118 In patients with hemorrhagic shock, it is currently recommended to start two large-bore IV lines and give a fluid bolus of 2L of a warm crystalloid solution.28 This practice, however, has been repeatedly questioned, particularly for penetrating trauma in an urban setting.119,120 Newer concepts, such as small-volume resuscitation with either hypertonic saline or hypertonic-hyperoncotic hy droxyethyl starch solution,121,122 permissive hypotension/hypotensive resuscitation,123 or hemoglobin-based oxygen carriers,124 are promising but currently still under investigation. The controversy regarding optimal fluid therapy goes hand in hand with the ongoing discussion of if a scoop-and-run (evacuate) or stayand-play (resuscitate on site) or BLS versus ALS approach is superior in major trauma (see Controversies in Prehospital Emergency and Trauma Care).125 There is evidence supporting a minimalistic on-the-scene management (BLS only) for penetrating trauma to the torso in an urban setting because transport times are very short, usually a younger age group is affected that will tolerate a period of hypotension well, and the injury can only be treated by emergent surgery. On the other hand, evidence is scant to support a scoop-and-run approach for blunt trauma, especially in the rural setting where transport times are much longer than in an urban setting. Irrespective of the current controversy regarding fluid resuscitation in major trauma, the goal is to minimize on-scene time to less than 10 to 15 minutes (or less than 5 minutes in a hypotensive penetrating trauma victim) as well as transport times. Helicopter transport has been advocated for major trauma patients because it was shown to reduce mortality as well as facilitate treatment in a trauma center.126 Current criteria that patients should be transported to a trauma center can be found in Table 73-2. In major head trauma (see Chapter 63), a modified approach is required in the field: adequate oxygenation

Anesthesia and Prehospital Emergency and Trauma Care 2321


Table 73-2 On-Scene Triage Criteria for Major Trauma Necessitating Transport to Trauma Center Mechanism of Injury Extrication time >20 minutes High-speed crash with: Intrusion into passenger compartment >20cm Speed >40mph Ejection Death of another passenger in same car Rollover Pedestrian or bicyclist hit by motor vehicle Motorcycle crash >20mph or with separation of rider from motorcycle Falls >3m (9.8ft) Physiologic Criteria Systolic blood pressure <90mmHg Respiratory rate <10 or >29 breaths/min Glasgow Coma Scale score <14 Pregnancy Anatomic Criteria Flail chest Two or more proximal long bone fractures Pelvic fracture Penetrating injury to head, neck, torso, or extremities proximal to elbow or knee Open or depressed skull fracture Amputation proximal to wrist or ankle Paralysis Discretion of EMT Per discretion of EMT, patients can be declared a major trauma patient. In addition, preexisting conditions of the patient can lead to an upgraded status, such as coronary artery disease, congestive heart failure, severe COPD, morbid obesity, or bleeding disorders.
COPD, chronic obstructive pulmonary disease. From OConnor RE: Trauma triage: Concepts in prehospital trauma care. Prehospital Emerg Care 10:307-310, 2006.

73

Dyspnea/Respiratory Distress
Dyspnea is one of the most common presenting symptoms in prehospital care, carrying a significant mortality rate of 12% to 18% in adults.128 Each year in the United States alone, 2 million patients with respiratory distress are transported by EMS. As shown in a large Canadian outcome study (OPALS), prehospital ALS decreases mortality in these patients.128 In adults, acute shortness of breath is most commonly caused by either broncho spasm or pulmonary edema secondary to CHF, but other differential diagnoses such as pulmonary embolism, pneumonia, and foreign body obstruction must be considered (Table 73-3).129 Identifying the correct diagnosis on the scene is not always as straightforward as it may seem130-132 because the etiology of dyspnea can be multifactorial, for example, in a patient with severe COPD and CHF who develops pneumonia. Patient Evaluation The most important aspect of the initial evaluation in a patient with acute dyspnea is to identify if the condition is life threatening with imminent respiratory failure and requires immediate treatment. Physical signs and symptoms indicating a life-threatening condition are cyanosis, tachypnea (respiratory rate <6 breaths/ min or >30 breaths/min), stridor, tachycardia, inability to speak in full sentences (<5 to 6 words), and the use of accessory respiratory muscles.133 The use of pulse oximetry has quickly become the standard of care in diagnosing hypoxemia and adds an objective measurement of the severity of respiratory distress. Irrespective of the underlying cause, all patients with respiratory distress should receive high-flow oxygen, ideally through a non-rebreathing facemask. Owing to space constraints, we will focus on the two most common causes of respiratory distress in adults: bronchospasm due to asthma and COPD and pulmonary edema due to CHF. Prehospital Management Bronchospasm: COPD and Asthma Bronchospasm as a cause for dyspnea is often suggested by the patients medical history and current medications. Asthmatics and COPD patients often use inhaled -agonists and corticosteroids and, in severe cases, also systemic agents. These patients tend to have frequent encounters with EMS and emergency departments. In addition to the clinical signs of dyspnea, wheezing and coughing are classic symptoms of bronchospasm. In extreme forms of bronchospasm, no breath sounds can be auscultated,
Table 73-3 Common Nontraumatic Causes of Acute Shortness of Breath in Adults Exacerbated chronic obstructive pulmonary disease Asthma Cardiogenic pulmonary edema (congestive heart failure) Noncardiogenic pulmonary edema (inhalational, sepsis-related) Pneumonia Pulmonary embolism Foreign body obstruction Aspiration
Adapted from Fowlkes TD: Shortness of breath. In National Association of EMS Physicians and Kuehl AE (eds): Prehospital Systems and Medical Oversight, Dubuque, IA, Kendall and Hunt, 2002, p 665.

Section V Adult Subspecialty Management

and avoiding hypoxemia (SpO2 >90%), maintaining an adequate cerebral perfusion pressure >70mmHg with a systolic blood pressure of at least 90mmHg with isotonic or hypertonic fluids, endotracheal intubation for patients with a GCS less than 9, maintaining normoventilation (end-tidal CO2 of 35-40mmHg) and avoiding hyperventilation or hypoventilation are considered the mainstays of prehospital therapy in traumatic brain injury.127 However, if cerebral herniation is suspected, moderate hyperventilation (end-tidal CO2 of 30-35mmHg) and hyperosmolar therapy with mannitol or hypertonic saline should be instituted.

Adult Medical Emergencies


Most textbooks on prehospital emergency medicine discuss adult medical emergencies under the patients final diagnosis, such as congestive heart failure (CHF) or chronic obstructive pulmonary disease (COPD). This, however, is not how patients present to the EMS team on the scene; instead, patients usually present with a chief complaint (or lead symptom), such as dyspnea or chest pain. Most often, adult medical emergencies fall into only three separate chief complaints: dyspnea/respiratory distress, cardiac/circulatory emergencies, and altered level of consciousness. Therefore, we will discuss adult medical emergencies in this chapter by their respective chief complaint.

2322 Adult Subspecialty Management

resulting in a silent chest. It has been a long-standing myth in EMS that patients with severe COPD should not receive highflow oxygen to prevent respiratory arrest.129 However, these patients often die as a result of hypoxemia and thus should be given the same oxygen therapy as any other patient in respiratory distress. On-the-scene management for patients with asthma attacks or acute COPD exacerbations is very similar and follows the recommended guidelines.134 Besides providing oxygen to prevent hypoxemia, the first line of treatment is inhaled -agonists plus an anticholinergic agent such as albuterol or salbutamol plus ipratropium, either via nebulized aerosol or metered-dose inhaler. Often these patients will already have made extensive use of their inhalers before calling EMS, so in moderate to severe exacerbations the use of systemic corticosteroids, for example, an IV bolus of methylprednisolone, 60 to 125mg, is clearly indicated. In lifethreatening asthma and COPD exacerbations, IV -agonists (epinephrine, 100-200 g IV bolusequals one tenth of the recommended resuscitation dose) are particularly helpful. In case of impending or actual respiratory arrest, endotracheal intubation is warranted, often after RSI with ketamine as the preferred induction agent owing to its favorable bronchodilating properties. The EMS crew must be vigilant for additional conditions that add to the severity of hypoxemia, such as pneumonia, pneumothorax, anaphylaxis, and pulmonary embolism. Pulmonary Edema: Congestive Heart Failure In addition to standard monitoring, a 12-lead ECG should be obtained to guide prehospital therapy for patients with acute pulmonary edema and rule out acute myocardial infarction. As a biomarker, brain natriuretic peptide (BNP) has become the mainstay in the clinical diagnosis of cardiogenic pulmonary edema, and some EMS systems now use point-of-care BNP testing on the scene to determine the cause of dyspnea.132 Noninvasive positivepressure ventilation (NIPPV) such as continuous positive airway pressure (CPAP) has become a standard therapy over the past several years in the treatment of acute pulmonary edema.135 Recently published recommendations suggest NIPPV for every patient with acute pulmonary edema.136,137 The subsequent therapy should be guided by the patients systolic blood pressure (SBP): SBP greater than 140mmHg: NIPPV plus nitrates (nitroglycerin spray or continuous IV infusion; recommended starting dose: 5-20 g/min) SBP 140 to 100mmHg: NIPPV plus nitrates plus diuretics (furosemide) if signs of systemic fluid retention SBP less than 100mmHg: hypoperfusion is predominant and signs of cardiogenic shock are present. These patients require inotropic agents such as dopamine and dobutamine and a carefully titrated volume challenge. Pulmonary edema with signs of acute coronary syndrome (ACS): these patients require ACS specific management in addition to NIPPV plus nitrates.136 The treatment goals are to stabilize the patient, improve oxygenation and perfusion, and reduce dyspnea.

60). Much emphasis has been put on the role of EMS in the management of ST-segment myocardial infarction (STEMI), but other conditions such as pulmonary embolism, hypertensive emergencies, or dysrhythmias require a similar high quality of prehospital diagnostic and therapeutic skills. In this overview, we will focus on the most common emergencies resulting in chest painacute coronary syndrome (ACS) and pulmonary embolismand also on hypertensive emergencies. Patient Evaluation The most important goal in evaluating a patient with acute chest pain or circulatory problems is to rule out a life-threatening condition such as ACS or pulmonary embolism. Because chest pain is a common complaint, identifying high-risk patients among all the patients with back spasm, esophageal acid reflux, or other gastrointestinal disorders is difficult. History and physical examination as well as a 12-lead ECG often guide the diagnosis. Prehospital 12-lead ECG can be quickly performed by the EMS crew and transmitted to physician oversight if required and prolongs on-scene time only 3 to 4 minutes while offering very high sensitivity and specificity for detecting myocardial ischemia and in particular for detecting patients with STEMI. Assessing the type and quality of chest pain (stabbing versus burning, radiation, severity, and provocation) is critical in determining the differential diagnosis (Table 73-4). However, some patients with ACS have atypical pain symptoms (epigastric discomfort, jaw pain) or no pain at all in case of silent myocardial ischemia in diabetic patients. Prehospital Management Acute Coronary Syndrome Patients with acute coronary syndrome should immediately be put on continuous ECG monitoring in addition to blood pressure and oxygen saturation monitoring. Rapid access to a defibrillator is mandatory, especially for patients with STEMI. The most recent guidelines recommend starting supportive therapy on the scene consisting of morphine (pain control), high-flow oxygen, nitrates (if patient is not hypotensive), and oral aspirin, easily memorized as the well-known MONA acronym.138,139 -Blockers can be considered if the patient is tachycardic and/or hypertensive. The goal is to reduce myocardial oxygen demand while improving oxygen supply. The 12-lead ECG is a vital component of prehospital care for ACS because it is the only tool that can reliably identify patients who may benefit from prehospital reperfusion therapy (Fig. 73-4). The overarching goal is to limit myocardial ischemia time to less than 120 minutes (ideally <60 minutes). Early out-ofTable 73-4 Differential Diagnosis of Acute Chest Pain in Adults Acute coronary syndrome (unstable angina, myocardial infarction) Pulmonary embolism Gastroesophageal reflux Peptic ulcer disease Gastritis or esophagitis Aortic dissection Pericarditis Pneumonia Pneumothorax Pleurisy Boerhaaves syndrome (esophageal rupture) Pancreatitis Musculoskeletal pain

Cardiac and Circulatory Emergencies


Chest pain and circulatory problems (e.g., syncope, hypertension) are encountered frequently in the prehospital setting (see Chapter

Anesthesia and Prehospital Emergency and Trauma Care 2323 Call 9-1-1 Call fast Hospital fibrinolysis: Door-to-needle within 30 min Onset of symptoms of STEMI 9-1-1 EMS Dispatch EMS on-scene EMS Encourage 12-lead ECGs Triage Consider prehospital fibrinolytic if capable Plan and EMS-to-needle within 30 min Not PCI capable Inter-hospital transfer PCI capable

73

Section V Adult Subspecialty Management

Goals Patient 5 min after symptom onset Dispatch 1 min EMS on scene Within 8 min EMS transport Prehospital fibrinolysis: EMS-to-needle within 30 min EMS transport: EMS-to-balloon within 90 min

Patient self-transport: Hospital door-to-balloon within 90 min

Total ischemic time: within 120 min*


*Golden hour = first 60 minutes

Figure 73-4 Prehospital management priorities in patients with ST-segment elevation myocardial infarction. ECG, electrocardiogram; EMS, emergency medical service; PCI, percutaneous coronary intervention; STEMI, ST-segment myocardial infarction. (Modified from Antman EM, Anbe DT, Armstrong PW, etal: ACC/ AHA guidelines for the management of patients with ST-elevation myocardial infarction: A report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines [Committee to Revise the 1999 Guidelines for the Management of patients with acute myocardial infarction]. J Am Coll Cardiol 44:e1-e211, 2004; and Armstrong PW, Collen D, Antman E: Fibrinolysis for acute myocardial infarction: The future is here and now. Circulation 107:2533-2537, 2003.)

hospital fibrinolysis reduces mortality in patients with STEMI or new-onset left bundle branch block140 and is the recommended standard for fibrinolysis-capable EMS systems.138 Several fibrinolytic agents are available, and the choice is often based on the institution. Streptokinase was the first fibrinolytic agent but is now rarely used in prehospital care. Most EMS systems now use recombinant tissue plasminogen activator (rt-PA or t-PA) or modified forms of t-PA, such as reteplase or tenecteplase (TNKase). The latter offer convenient single- or double-bolus dosing, making them preferable fibrinolytic agents in the prehospital setting.141 It is important to note, however, that fibrinolytic therapy has several contraindications (Table 73-5) and also carries the risk of intracranial hemorrhage, resulting in major disability or death in 1% to 2% of the patients. If fibrinolysis cannot be performed on the scene, rapid transport to a facility with percutaneous coronary intervention is warranted with the goal of EMS arrival-to-balloon time of less than 90 minutes. Pulmonary Embolism A definitive diagnosis of pulmonary embolism is difficult to establish on the scene because signs and symptoms are nonspecific and no rapid objective measure, such as the 12-lead ECG in STEMI, is available for on-site diagnosis. Often, the patients medical history is telling; for example, an immobilized leg and signs of a preexisting deep venous thrombosis can be found, such as unilateral, tender swelling in one extremity. The severity of pulmonary embolism directly relates to the size of the clots and ranges from mild symptoms of dyspnea and coughing to acutely life-threatening states of cardiogenic shock due to acute rightsided heart failure. Classic symptoms of pulmonary embolism include distended neck veins, hypoxemia, tachycardia, tachypnea, and chest pain. More objective, but nevertheless unspecific find-

Table 73-5 Contraindications of Fibrinolytic Therapy in ST-Segment Elevation Myocardial Infarction Absolute Contraindications Any prior intracranial hemorrhage Known structural cerebrovascular lesion (e.g., AVM) Known malignant intracranial neoplasm (primary or metastatic) Ischemic stroke within 3 months EXCEPT acute ischemic stroke within 3 hours Suspected aortic dissection Active bleeding or bleeding diathesis (excluding menses) Significant closed-head trauma or facial trauma within 3 months Relative Contraindications History of chronic, severe, poorly controlled hypertension Severe uncontrolled hypertension on presentation (SBP > 180mmHg or DBP > 110mmHg)* History of prior ischemic stroke < 3 months, dementia, or known intracranial pathologic process not covered in contraindications Traumatic or prolonged (>10 minutes) CPR or major surgery (<3 weeks) Recent (within 2 to 4 weeks) internal bleeding Noncompressible vascular punctures For streptokinase/anistreplase: prior exposure (<5 days ago) or prior allergic reaction to these agents Pregnancy Active peptic ulcer Current use of anticoagulants: the higher the INR, the higher the risk of bleeding
AVM, arteriovenous malformation; SBP, systolic blood pressure; DBP, diastolic blood pressure; INR, international normalized ratio. *Could be an absolute contraindication in low-risk patients with myocardial infarction. Adapted from American Heart Association Guidelines for Cardiopulmonary Resuscitation and Emergency Cardiovascular Care. Part 8. Stabilization of the Patient with Acute Coronary Syndromes. Circulation 112(24 Suppl):IV-89-110, 2005.

2324 Adult Subspecialty Management

ings are new right bundle branch block, symmetric T-wave inversion in the anterior leads (V1-V4), and SIQIII pattern in the 12-lead ECG.142 Often the 12-lead ECG is more useful to rule out other conditions as myocardial infarction than ruling in pulmonary embolism. In absence of an objective means to reliably diagnose pulmonary embolism in the prehospital setting, the management is largely supportive. Applying standard monitoring, administration of high-flow oxygen, establishing an IV line, and pain control with opioid analgesics are considered standard therapy. In severe pulmonary embolism that is usually accompanied with systemic hypotension (SBP < 90mmHg), in other words with signs of shock, circulatory support with inotropic agents is often required. The prehospital use of heparin in pulmonary embolism is not universally recommended and must be determined on a case-bycase basis, taking into account the likelihood of pulmonary embolism and the severity of symptoms. Prehospital fibrinolysis in a patient with pulmonary embolism is not recommended, although it has been successfully used in extremis.143 Hypertensive Emergencies EMS often responds to patients with acute hypertensive episodes, but these are rarely true hypertensive emergencies. The classification of hypertensive episodes is very helpful in determining the urgency of a prehospital intervention (Table 73-6). Acute endorgan damage in conjunction with a hypertensive episode, such as hypertensive encephalopathy, stroke, pulmonary edema, myocardial infarction, aortic dissection, acute renal failure, and preeclampsia must be immediately treated and progression must be stopped. The goal is to lower the blood pressure by 20% to 25% within 30 to 60 minutes.144 However, in the setting of cerebral infarction, great care must be exercised in reducing blood pressure as needed because the ischemic territory may be critically dependent on collateral perfusion pressure; this is an area of controversy in need of further study. Evidence for benefits of antihypertensive treatment is scant; some argue that only excessively high SBP or diastolic blood pressure (DBP) should be treated.145 On the other hand, patients with an aortic dissection require swift and aggressive treatment with a target blood pressure of SBP less than 120mmHg and DBP less than 80mmHg.146 The choice of antihypertensive should take into account the underlying pathophysiology. For example, patients with acute hypertension plus acute coronary syndrome and tachycardia
Table 73-6 Classification of Acute Hypertensive Episodes Hypertensive emergency Also termed hypertensive crisis; hypertension with acute end-organ dysfunction Hypertensive episode with high risk of imminent end-organ damage that has not yet occurred SBP > 180mmHg or DBP > 110mmHg without signs of evolving or imminent end-organ dysfunction Related to anxiety or primary complaint

might benefit from a -blocker and nitroglycerin, whereas in intracranial hemorrhage labetalol is the preferred agent. The antihypertensive agent of choice in many European EMS systems, urapidil, an -adrenoceptor antagonist, is currently not available in the United States.

Altered Level of Consciousness


Altered level of consciousness is often encountered in prehospital care. The differential diagnoses are vast; therefore, most patients will receive only supportive care in the field. The overarching goal is to stabilize the patients vital signs, quickly identify and treat potentially life-threatening conditions, and determine a likely cause.147 For example, rapid detection of severe hypoglycemia in an unconscious patient can be assessed by point-of-care testing and reversed by administration of an IV glucose bolus. In this section, we focus on general principles but briefly mention a few specific conditions, such as stroke or seizure. Patients with an altered level of consciousness should immediately receive high-flow oxygen and IV access, be undressed, and undergo standard monitoring. The addition of continuous capnography, when available, is helpful to assess adequate ventilation. Level of unconsciousness and determination of the GCS score should be performed (see Chapter 46). In the prehospital setting, it is often sufficient to distinguish if the patient is alert (responsive to verbal commands), only responsive to painful stimuli, or is completely unresponsive (unconscious).148 Because patients with impaired consciousness are at risk for aspiration, the need for airway protection needs to be assessed. Recommended diagnostic steps on the scene include determining the pupil size and reaction to light, measuring core body temperature, and obtaining point-of-care blood glucose levels. During the initial neurologic examination, EMS providers must determine if the main neurologic symptoms are global or if there is a focal neurologic deficit. Nasal or oral airways equipment may be inserted to keep the airway open and to test for a gag reflex. Once these basic steps have been accomplished, the EMS team should try to obtain as much information about the patient and the circumstances as possible to narrow the possible causes for the patients unconsciousness (Table 73-7). Often, this can only be accomplished by involving bystanders and family members. Important aspects are onset and progression of unconsciousness (quick or over a few hours), fever, the patients status before the onset of symptoms, headache, medical history and medications (e.g., hypertension, diabetes, depression), signs of intoxication (open drug containers, needles, drug paraphernalia), and signs of trauma. For patients with hypoglycemia, the recommended treatment plan is to immediately administer dextrose (D50W, 50mL = 1 ampule). If an IV cannot be established, IM glucagon (1-2mg) is a viable alternative. On average, blood glucose levels rise after this treatment between 100-150mg/dL and patients are fully awake within a few minutes. In cases where opioid overdose is suspected (miosis, respiratory rate <10 breaths/min), the opioid antagonist naloxone should be administered. The recommended starting dose is 0.4mg IV, but it should be titrated to clinical effect, which in this case is restoration of spontaneous ventilation and not complete awakening. A wide range of required dosages may be required, sometimes requiring as little as 0.2mg and sometimes up to 10mg.148 If too much naloxone is administered,

Hypertensive urgency

Acute hypertensive episode

Transient hypertension

SBP, systolic blood pressure; DBP, diastolic blood pressure. Adapted from Chobanian AV, etal: Seventh Report of the Joint National Committee on Prevention, Detection, Evaluation, and Treatment of High Blood Pressure. Hypertension 42:1206-1252, 2003.

Anesthesia and Prehospital Emergency and Trauma Care 2325


Table 73-7 Causes of Altered Level of Consciousness in Adults: AEIOU-TIPS Alcohol and Airway (hypoxia) Epilepsy, Electrolytes, Encephalopathy Insulinhypoglycemia Opioids/Overdose Urea (metabolic) Trauma, Tumor Infection Psychiatric Shock, Subarachnoid hemorrhage, Snake bite
Adapted from Wolfe RE, Brown DFM: Coma and depressed level of consciousness. In Marx JA (ed): Rosens Emergency Medicine. Philadelphia, Mosby, 2006.

73

ondary complications such as hypoxia. IV access is difficult to establish during a generalized seizure. High-flow oxygen should be administered via a non-rebreathing facemask. Status epilepticus is a state of ongoing, refractory general seizure activity and is a true life-threatening emergency; sometimes treatment requires induction of general anesthesia.154

Section V Adult Subspecialty Management

Pediatric Emergencies
Pediatric emergencies are one of the biggest challenges for a prehospital provider (see Chapter 82). Between 5% and 8% of emergency calls to an EMS are pediatric emergencies.155,156 In the first years of age, common pediatric emergencies are respiratory problems and febrile seizures. Trauma increases in incidence in higher age groups.155 Some of the skills, such as securing the airway and vascular access, are more difficult to perform for EMS providers in children than in adults. In this section, the main differences in etiology, diagnostic procedures, and treatment are stressed and the chief symptoms and their relevance in the pediatric population are discussed.

severe opioid withdrawal symptoms can result. Similarly, if patients with benzodiazepine overdose (sleeping pills) are treated aggressively with the antagonist flumazenil, generalized seizures can result that may be unresponsive to further benzodiazepine treatment. Patients with hypoglycemia or opioid intoxication comprise probably the only category of patients with an altered level of consciousness that can be successfully identified and treated in the field. Only over the past decade, stroke patients have received similar attention in prehospital care as patients with acute myocardial infarction. It is now clear that a similar window of opportunity exists in stroke patients as in patients with acute myocardial infarction.149 The goal is to quickly identify stroke in the field and rapidly transport the patient to a facility that can perform computed tomography (CT) to aid in diagnosis and rule out the presence of intracranial hemorrhage. Ischemic stroke without hemorrhage may be treated with thrombolysis. Classic symptoms of stroke are hemiparesis, aphasia, and altered level of consciousness. Formal prehospital screening tools such as the Cincinnati Prehospital Stroke Scale are useful and should be used.150 Stroke is a leading cause of death in the United States and is the number one cause for disability in adults. Rapid identification and treatment of stroke limits brain damage and improves survival after stroke.145,151 Because of the inability to clinically distinguish between ischemic stroke (85% of cases) and hemorrhagic stroke (15%), prehospital thrombolysis cannot be performed as in patients with STEMI. Thus, supportive care to stabilize the patients vital signs and to avoid risk factors that negatively influence outcome, such as hypoxia, hypotension, and hyperglycemia are main treatment goals in stroke in the prehospital setting.152 Therefore, stroke patients require supplemental oxygen, IV access, continuous ECG monitoring, as well blood pressure monitoring. Hyperglycemia should be corrected to normal levels. Blood pressure management in stroke patients is controversial; it is clear that any aggressive lowering of SBP worsens outcome and that a certain degree of hypertension in acute stroke might be beneficial. If other end organ dysfunction is apparent, such as myocardial ischemia, or SBP greater than 220mmHg, cautious antihypertensive treatment can be started on the scene. One of the most important aspects of prehospital care for stroke patients is the expeditious transport to a facility that specializes in stroke treatment, ideally a stroke center.153 Seizure is a common disorder seen in EMS practice. Often, the seizure is over when EMS arrives and the patient will be found in a post-seizure state that seldom requires intervention. If the seizure is ongoing, however, the seizure may be usually treated by the IV administration of benzodiazepine and prevention of sec-

Dyspnea/Respiratory Distress
Respiratory failure is the leading cause for cardiac arrest in children. The anatomy of the airway and the smaller respiratory reserve due to a smaller residual capacity as well as a higher basal metabolism make children more vulnerable to hypoxic events. Infections or an obstructed airways by foreign bodies dominates the respiratory problems in younger children; in older children the airway problems are usually due to direct trauma or altered level of consciousness after head trauma or shock. Patient Evaluation Assessing the child may be more difficult if the child is frightened. An agitated child interferes with many diagnostic procedures, and the resultant stress level increases oxygen consumption, potentially leading to a decompensation of the situation. After assessing the airway, the work of breathing is estimated. Physical signs and symptoms indicating respiratory distress are tachypnea and intercostal, sternal, and subcostal retractions. Children use accessory muscles to support breathing, which may manifest as head bobbing and nasal flaring. An early sign of distress is use of chest wall muscles. Paradoxical breathing movements of chest and abdomen (see-saw respiration) indicate decompensated respiration. Abnormal sounds from the airway pinpoint the level of airway obstruction. Inspiratory stridor indicates an extrathoracic airway problem. In contrast, expiratory noises are signs of lower tracheal problems, whereas wheezing leads to a problem at the bronchial level. Hypoxia can cause vasoconstriction, and cyanosis can be masked by skin pallor. Pulse oximetry should be mandatory in respiratory distress. Prehospital Management Pediatric respiratory support may include facemask with supplemental oxygen, assisted bag-mask-valve ventilation, or controlled ventilation with a secured airway. Because of the different anatomy dependent on the age of the child and the limited skills in most nonanesthesiologist prehospital providers, difficulties in airway management occur more often than in adults.

2326 Adult Subspecialty Management

Infection: Asthma
Laryngotracheobronchitis (croup) is a typical emergency of the 6-month- to 3-year-old child. Because of the dramatic onset, usually during the night with sudden loud inspiratory stridor and sometimes respiratory distress, it frightens both child and parents. Except for a minor infection of the upper respiratory tract, the medical history is usually noncontributory. On-the-scene management of mild croup (without retractions) consists of psychological support by the EMS team, humidified oxygen via facemask in a nonthreatening manner, and transport to a pediatric emergency department. More severe cases with retractions and/or neurologic signs such as apathy or agitation require oral dexamethasone (0.15-0.6mg/kg) and, as a next step, nebulized epinephrine (0.5mL of 2.25% racemic epinephrine or 5.0mL of epinephrine 1:1000, irrespective of the weight of the child).157 The need for prehospital intubation is rare. Another pediatric infection of the upper airway is much more dangerous but fortunately very rare: epiglottitis. Epiglottitis usually occurs in patients age 2 to 7 years and is characterized by inspiratory stridor, hypersalivation, high fever, and inability to swallow. The voice sounds muffled. Any anxiety-provoking maneuvers such as examination or IV access should be avoided. The child should be attended at all times by someone capable of establishing an airway and brought to a facility where endoscopy can be done. Cricothyroidotomy may be necessary in very severe cases.158 Bronchospasm is a frequent symptom in pediatric respiratory infections. Therapy ranges from supportive therapy only in mild cases to inhalational -adrenergic agonists and corticosteroids and intravenous drug therapy sometimes requiring anesthesia and intubation in status asthmaticus. Ketamine is the anesthetic drug of choice in exacerbated bronchospasm.

Patient Evaluation After assessing consciousness, airway, and breathing, the circulatory status is determined. Heart rate, pulse quality, and capillary refill time are quick tools for evaluation. Blood pressure and preload assessment complete the picture. To determine if the child is already decompensated, additional information of the neurologic status (apathetic or agitated) is taken into account. Prehospital Management Arrhythmia Bradyarrhythmia is usually the result of hypoxia or poisoning. If the heart rate is below 60 beats/min and the child shows signs of inadequate perfusion (unconsciousness, apnea), resuscitation is started and the underlying cause can then be identified and reversed. Most tachyarrhythmias are supraventricular and need only be treated if decompensated. After exclusion of causes for a reflex tachyarrhythmia, specific therapy starts with vagal maneuvers, adenosine, and, if the child is severely decompensated, cardioversion. During resuscitation, epinephrine (10 g/kg), atropine (0.02mg/kg), and amiodarone (5mg/kg) are given according to international guidelines (see Chapter 97).160 Recurrent ventricular tachycardia/ventricular fibrillation in an otherwise healthy child or a torsades de pointes pattern on an ECG suggests a long-QT syndrome, which follows a different path in the universal algorithm of resuscitation. In a long-QT syndrome, amiodarone is contraindicated and magnesium is given as a bolus (3-12mg/kg).161 Shock Hypovolemic shock due to acute blood loss or acute diarrheal disease is the most common type, followed by septic, cardiogenic, and distributive shock.162,163 Obtaining quick vascular access in a decompensated situation is essential, but difficult. Venous access is the first choice; but if it cannot be rapidly established, intraosseous access is the next alternative. Intraosseous access provides a quick and reliable venous access, adequate for all IV medications. An initial fluid bolus of 20mL/kg of fluid should be given. Crystalloids are preferred.164 After reassessing the child, a second bolus can be given. If shock persists after the second reassessment, packed red blood cells should be administered, especially in trauma victims, which usually can only be done in the emergency department.159 This requires rapid transport to the nearest hospital and minimizing on-scene times.

Foreign Body Airway Obstruction


Foreign body airway obstruction is characterized by the sudden onset of respiratory distress associated with coughing, gagging, or stridor in a child with no other signs of illness. If the child is coughing effectively, then no immediate action to remove the foreign body by the EMS is necessary. The child is monitored and brought to the hospital and advised to try to expel the foreign body by coughing. If the coughing becomes ineffective, intervention is required. With a series of back blows (5) in turn with thrusts (abdominal in children >1 year and chest thrusts in infants <1 year of age) attempts are made to remove the foreign body. If the child becomes unconscious, standard BLS must be started159 and the airway must be rapidly checked with direct laryngoscopy. If the foreign body cannot be removed, an effort can be made to advance the foreign body into the bronchus and ventilate the remaining lung.

Altered Level of Consciousness


Altered level of consciousness is one of the first signs for a decompensation of a respiratory or circulatory failure. Besides that, disorders of the brain (infection, ischemia, trauma, intoxication, metabolic disorders like diabetic coma, hepatic or nephrologic problems and tumors) will alter the level of alertness. Patient Evaluation After evaluation of the respiratory and circulatory status, a quick neurologic assessment should be made. The GCS examination is adapted to the age-related abilities of the child.165 Especially, the

Cardiac Emergencies and Shock


Life-threatening cardiac rhythm disturbances in children are more frequently the result, rather than the cause, of acute emergencies. Nevertheless some arrhythmias are seen, especially in children with congenital problems.

Anesthesia and Prehospital Emergency and Trauma Care 2327

73

motor quality of the GCS has a good predictive value for the outcome after traumatic brain injury.166 In unconscious children, the repeated assessments of the pupils can give information about intracranial injury and/or elevated intracranial pressure. Blood glucose measurement is mandatory. Prehospital Management Infection If signs of meningeal irritation are seen, bacterial meningitis accounts for roughly 30%, whereas viral or aseptic meningitis is the cause in approximately 13% of cases. Other diagnoses are pneumonia (8%), other serious bacterial infections (2%), upper respiratory tract infections or other self-limiting diseases (46%).167 Independent predictors of the presence of bacterial meningitis obtained from a patients history and physical examination are the duration of the main complaint, a history of vomiting, and meningeal irritation, cyanosis, petechial bleeding, and altered consciousness at physical examination.168 Identifying this potentially life-threatening disease is an important challenge for the prehospital provider. Airway management and fluid resuscitation may be necessary before transferring the child to a pediatric emergency department. If the diagnosis of bacterial meningitis is confirmed, antibiotic prophylaxis of all persons having had direct contact with the patient is necessary. Seizures Febrile seizures are the most common seizure disorder in childhood, occurring in 2% to 5% of children, typically between the age of 6 months to 5 years.169 Usually the seizure has already terminated when the EMS team arrives and the patient is postictal. An antipyretic therapy may prevent a recurrence of the seizure. To rule out other underlying conditions for a seizure, such as hypoxia, intoxication, infectious reasons, metabolic disorders, trauma, or sunstroke, the child should be hospitalized after stabilization of the vital signs. If seizures recur or are still ongoing (status epilepticus) the initial assessment should follow the ABC principle. High-flow oxygen should be given and blood glucose should be tested. For those children in whom intravenous access can immediately be established, lorazepam, 0.1mg/kg, should be given intravenously. If a rapid IV access is not possible, diazepam (0.5mg/kg) is given rectally. If after 10 minutes the initial seizure has not stopped or another seizure has begun, a second dose of lorazepam (0.1mg/kg) should be given.170 The child has to be repeatedly reassessed to verify that an unrecognized underlying cause has not been missed. During transport, epileptogenic maneuvers should be avoided, such as overzealous papillary examination with a strong light.

Prehospital Management Carrying out a quick primary survey, treating the life-threatening conditions according to the ABCDE approach, and organizing the communication and quick transport to an adequate facility are the main issues of trauma treatment of children.

Section V Adult Subspecialty Management

The Abused Child


Child abuse is common and a major challenge for prehospital providers (see Chapter 82).171 Typical injuries include evidence of injuries of different ages, atypical burns, abrasions and hematomas reflecting the form of the beating hand or instrument, shaken baby syndrome, and injuries at unusual locations and on the top of the head. A delayed or inconclusive patient history should arouse suspicion. A management goal is to guarantee the wellbeing of the child. In suspected cases of child abuse, the appropriate authorities are informed, depending on the local system, and the patient is transferred to a hospital or another safe facility.

Prehospital Mass Casualty Incident Management and Disaster Medicine


Prehospital providers are often faced with situations such as multiple vehicle accidents where the number of simultaneous patients initially exceeds the capabilities of the rescue squad or local rescue forces. This creates an inadequate ratio of rescuers to victims. In these mass casualty situations (Table 73-8; see Chapter 74), the concept of triage must be employed: trying to save as many patients as possible without being able to provide individual one-on-one care. During the mass casualty event, the boundaries between an emergency, a major incident (mass casualty situation), and a disaster are constantly in a state of flux. The first and most important duty of the first prehospital team at the scene is to communicate the dimension of the accident/disaster and the ongoing hazards to the dispatcher. A mass casualty situation creates many problems for EMS providers. Frequent errors included having multiple communicators on site (38%), misidentifying the number of victims (56%), and having unclear information for the resource physician (43%).172 It has been estimated that prehospital triage information was deemed appropriate in only one third of mass casualty events. After alerting local backup resources, patients must be triaged. In man-made and natural disasters, prehospital providers and their EMS systems have to shift their triage methodology from a daily operational framework of treating the most severely injured patient first and providing the highest level of care for each patient to the concept of providing the greatest good for the greatest number of casualties. This shift is necessary to identify those critically injured patients who can benefit from immediate, lifesaving interventions and to conserve resources.173 In disasters such as hurricane Katrina, coordination of the rescue and relief effort on a national level is required. National disaster resources such as DMAT (disaster medical assistance teams) from unaffected states will be deployed as will be military forces (National Guard).

Trauma
Trauma is a leading cause of death from the ages 1 to 18. Boys are more at risk than girls are. Patient Evaluation Although the pattern of injury differs from that of the adult patient (more pelvis, chest, abdominal, and head trauma) the evaluation and the treatment principles are in accordance with the adult management guidelines.

2328 Adult Subspecialty Management


Table 73-8 Definitions in Disaster Medicine Level of Prehospital Medical Care Possible Individual medicine Triage; increasing level of medical treatment dependent on resources Triage; first aid; significant delay until professional medical help available usually only after temporary infrastructure was set up

Term Emergency Mass casualty

Definition Incident that demands urgent action; impact: minutes to hours Large number of casualties in a short period that initially exceed local logistic support capabilities; impact: hours to days Large number of casualties and destroyed infrastructure; exceeds local response capabilities. Supraregional (national) forces are needed to solve the problem; damage and interruption last long-term (weeks to years)

Examples Patient with bicycle accident, stroke, angina Airplane crash, train collision, mass poisoning Hurricane, earthquake, tsunami, nuclear disaster (e.g., Chernobyl)

Disaster

Different triage strategies (Simple Triage and Rapid Treatment-START, CareFlight Triage, modified START) show comparable results in identifying the severely injured (Table 73-9). Triage implies constant reevaluation of victims because conditions of the victims and of available resources change continuously.174 Patients whose injuries are so grave (i.e., traumatic cardiac arrests) that their prognosis is dismal should not be classified as critical in the setting of a large mass casuality incident (MCI). They are likely to consume large quantities of resources in a futile resuscitative effort.175 The motor score of the GCS and systolic blood pressure show good association with severe injury.175 The emergency physician on the scene has to guarantee that the treatment priorities are followed by the EMTs rather than treating patients by himself or herself. In the first phase, only BLS, such as opening the airway, recovery position, tourniquets, and positioning, should be done. Advanced procedures such as IV access, oxygen therapy, anesthesia, intubation, and ventilation can be performed as soon as resources are available. With incoming transport capacity, patients are stabilized and transported according to urgency priorities. For
Table 73-9 Triage Groups in Mass Casualty Incidents With the aid of triage tags or colored flags the patients are sorted into four groups: Green/minor Minor injuries with no immediate need of treatment; psychological support needed; patients can walk Urgent treatment, no vital threat Immediate vital threat Wounds, minor fractures

example, patients with intra-abdominal bleeding requiring emergent surgery have the highest transport priority. For disposition of the patients to adequate hospitals, a quick registration and documentation stage is necessary.

Controversies in Prehospital Emergency and Trauma Care


The EMS literature is full of conflicting opinions and recommendations about what the ideal EMS system and the ideal level of prehospital care are. The plethora of differing opinions is likely due to the vast differences in how EMS systems are organized around the world and probably also to a lack of reliable evidence from high-quality research. In the United States, one of the most important documents ever written on the current problems in EMS, The Future of Emergency Care: Emergency Medical Services at the Crossroads, was published in 2007 by the Institute of Medicine of the National Academy of Sciences.12 It is the third part of a comprehensive study on the state of emergency medicine in the United States that also includes in-hospital emergency medicine176 and pediatric emergency care.177 It is likely that this report will have a similar effect on EMS as did The White Paper published in 1966.11 Here, we focus on those areas in prehospital care engendering the fiercest debate and where anesthesiologists could, in our opinion, make the biggest impact.

Yellow/delayed

Fractures, joint injuries, amputation, major blood loss, burns Respiratory insufficiency, shock, brain trauma, burns with immediate vital threat, abdominal trauma Respiratory arrest, cardiac arrest, head injuries with dismal prognosis

The Optimal Level of Prehospital Care: BLS Versus ALS or Scoop-and-Run Versus Stay-and-Play
Numerous articles have been published over the past decades favoring either a minimalistic approach to prehospital care, basically BLS level only, or the opposite, an aggressive strategy trying to bring the level of care in an emergency department to the scene. In our opinion this discussion is misguided. It is unlikely that a country that currently employs a paramedic-level EMS system will abandon this system in favor of a BLS-only system or vice versa (see Chapter 97). More importantly, this discussion leads away from the most important determinant of the overall efficacy and efficiency of any EMS system: the outcome of the

Red/immediate

Black/expectant

Dead or moribund

Anesthesia and Prehospital Emergency and Trauma Care 2329

73

patient. Whereas it makes perfect sense to argue in favor of a minimalistic prehospital approach in young patients in severe hemorrhagic shock after penetrating torso trauma in an urban area with transport times of only a few minutes,119,125 the same approach might be detrimental in rural areas with long transport times, an older patient population with significant comorbidities, and a predominantly blunt trauma mechanism.178-180 Therefore, in our view, it is overly simplistic to strongly argue for one or the other system. The condition of the patient, the environment, and the setting should determine the adequate EMS response and level of treatment. However, a well-trained physician or paramedic can easily downgrade his or her level of response on the scene and rush the patient to the hospital, whereas a basic level EMT cannot upgrade his or her response if required by the nature of the emergency. A similar discussion is going on between advocates of paramedic-based EMS models and physician-based models.14,94 Again, this is more a theoretical discussion about advantages and disadvantages of the two different systems rather than a realistic option for change. The discussion shows, however, the impact well-trained physicians and, in particular, anesthesiologists, can have on patient outcomes when directly involved in prehospital care.8,19,181 Airway management immediately comes to mind.

Prehospital Endotracheal Intubation and Rapid-Sequence Induction


This area might be the one where the clearest difference between an anesthesiologist-run and a paramedic EMS system materializes. Success rates for out-of-hospital endotracheal intubation performed by paramedics are unacceptably low compared with standards in the operating room and are associated with a high complication rate.182,183 More than 15 studies have shown that prehospital endotracheal intubation and RSI by paramedics are associated with an increased mortality rate or poorer neurologic outcomes compared with patients who were not intubated.66,67

Most of the studies focused on prehospital airway management in patients with severe traumatic brain injury,111,184,185 but also studies investigating prehospital pediatric intubation delivered sobering results.186-188 These results should serve as a wake-up call among anesthesiologists, especially in the United States, but so far they have failed to do so. The central question is: why are endotracheal intubation and RSI, procedures that many consider the standard of care and lifesaving, associated with worse outcomes when performed in the out-of-hospital setting? The reasons are manifold, but over the past several years research has shown that not only is airway management in the field much more challenging and difficult than in the hospital70 but also that paramedics and many nonanesthesiologist physicians are inexperienced in airway management and fail to maintain their airway skills learned during training because the actual numbers of intubations performed per year are very small.61,65,67,189,190 Less experience equals more intubation attempts, a more frequent complication rate, including severe hypoxemia and bradycardia during field RSI and airway management, and a higher rate of failed intubations.67,68,84,111,191 Unbeknown to many anesthesiologists, paramedics require only five successful endotracheal intubations in a clinical setting to graduate and not a single pediatric intubation, numbers that are significantly lower than any medical specialty training such as anesthesiology or emergency medicine.66,67 Even in the landmark paper on out-of-hospital pediatric intubation by Gausche and colleagues,188 paramedics who were previously not trained in advanced pediatric airway management received only 6 hours of classroom training including manikin intubations but not a single live intubation in a pediatric patient! Taken together, it appears that there is currently a great unmet need for our specialty to reach out to nonanesthesiologist airway providers and help define the ideal strategy for prehospital airway management focusing on the infrequent, occasional airway manager. It may be the case that endotracheal intubation and RSI as a standard of care need to be reassessed with a clearer vision of to whom this standard applies. Perhaps, alternate airways are safer and more efficient for EMS personnel who lack the necessary training or who cannot maintain their skills.

Section V Adult Subspecialty Management

References
1. Roessler M, Zuzan O: EMS systems in Germany. Resuscitation 68:45-49, 2006. 2. Adnet F, Lapostolle F: International EMS systems: France. Resuscitation 63:7-9, 2004. 3. Langhelle A, et al: International EMS Systems: The Nordic countries. Resuscitation 61:9-21, 2004. 4. Safar P: On the history of modern resuscitation. Crit Care Med 24(2 Suppl):S3-S11, 1996. 5. Sefrin P: Geschichte der Notfallmedizin in Deutschlandunter besonderer Bercksichtigung des Notarztdienstes [History of the rescue service in Germanyespecially in regard to emergency medicine]. Anasthesiol Intensivmed Notfallmed Schmerzther 38:623-629, 2003. 6. Frey R, Nagel E, Safar P: Mobile intensive care units: Advanced emergency care delivery systems. Symposium 1973 in Anesthesiology and Resuscitation, Mainz, Germany. Heidelberg, Springer-Verlag, 1976. 7. Pantridge JF, Geddes JS: A mobile intensive-care unit in the management of myocardial infarction. Lancet 2:271-273, 1967. 8. Einav SA, et al: Anesthesiologists on ambulances: Where do we stand? Curr Opin Anaesthesiol 16:585-591, 2003. 9. Gisvold SE: Helicopter emergency medical service with specially trained physiciansdoes it make a difference? Acta Anaesthesiol Scand 46:757-758, 2002. 10. Christenszen EF, et al: Anesthesiologists in prehospital care make a difference to certain groups of patients. Acta Anaesthesiol Scand 47:146-152, 2003. 11. Committee on Trauma and Committee on Shock; Division of Medical Sciences; National Academy of Sciences/National Research Council (US), Accidental Death and Disability: The Neglected Disease of Modern Society. Washington, DC, National Academy of Sciences, 1966. 12. Committee on the Future of Emergency Care in the United States Health SystemInstitute of Medicine of the National Academies, Future of Emergency Care: Emergency Medical Services at the Crossroads. Washington, DC, The National Academies Press, 2007. 13. Blackwell TH: Emergency Medical Service: Overview and Ground Transport. In Adams J, et al (eds): Rosens Emergency Medicine. Philadelphia, Mosby, 2006. 14. Dick WF: Anglo-American vs. Franco-German emergency medical services system. Prehosp Disaster Med 18:29-35; discussion 35-37, 2003. 15. National Association of EMS Physicians and Kuehl SE (eds): Prehospital Systems and Medical Oversight, 3rd ed. Dubuque, IA: Kendall and Hunt, 2002. 16. Pozner CN, et al: International EMS Systems: The United States: Past, present, and future. Resuscitation 60:239-244, 2004. 17. Lilja GP: Emergency Medical Services. In Tintinalli JE (ed): Emergency Medicine: A Comprehensive Study Guide. New York, McGraw-Hill, 2004, pp 1-5. 18. Gries A, et al: Realistic assessment of the physicianstaffed emergency services in Germany. Anaesthetist 55:1080-1086, 2006. 19. Moecke H, von Knobelsdorff G: The anesthesiologist in prehospital and hospital emergency

2330 Adult Subspecialty Management


medicine. Curr Opin Anaesthesiol 21:228-232, 2008. 20. McGinnis KK, et al: Air Medical Services: Future Development as an Integrated Component of the Emergency Medical Services (EMS) System. Prehospital Emerg Care 11:353-368, 2007. 21. Garner AA: The role of physician staffing of helicopter emergency medical services in prehospital trauma response. Emerg Med Australas 16:318-323, 2004. 22. Rea TD, et al: Dispatcher-assisted cardiopulmonary resuscitation and survival in cardiac arrest. Circulation 104:2513-2516, 2001. 23. Part 2: Adult Basic Life Support. Circulation 112(22 Suppl):III-5-III-16, 2005. 24. 2005 American Heart Association Guidelines for Cardiopulmonary Resuscitation and Emergency Cardiovascular Care. Circulation 112(24 Suppl): IV-1-IV-203, 2005. 25. Moore L, et al: Statistical validation of the Glasgow Coma Scale score. J Trauma 60:1238-1243; discussion 1243-1244, 2006. 26. McNett M: A review of the predictive ability of Glasgow Coma Scale scores in head-injured patients. J Neurosci Nurs 39:68-75, 2007. 27. Heard K, Bebarta VS: Reliability of the Glasgow Coma Scale for the emergency department evaluation of poisoned patients. Hum Exp Toxicol 23:197200, 2004. 28. National Association of Emergency Medical Technicians, PHTLSPrehospital Trauma Life Support, 6th ed. St. Louis, Mosby/JEMS, 2007. 29. Popov T: Review: Capillary refill time, abnormal skin turgor, and abnormal respiratory pattern are useful signs for detecting dehydration in children. Evid Based Nurs 8:57, 2005. 30. Shavit I, et al: A novel imaging technique to measure capillary-refill time: Improving diagnostic accuracy for dehydration in young children with gastroenteritis. Pediatrics 118:2402-2408, 2006. 31. McGuire TJ, Pointer JE: Evaluation of a pulse oxim eter in the prehospital setting. Ann Emerg Med 17:1058-1062, 1988. 32. Petterson MT, Begnoche VL, Graybeal JM: The effect of motion on pulse oximetry and its clinical significance. Anesth Analg 105(6 Suppl):S78-S84, 2007. 33. Nishiyama T: Pulse oximeters demonstrate different responses during hypothermia and changes in perfusion. Can J Anaesth 53:136-138, 2006. 34. Barker SJ, et al: Measurement of carboxyhemoglobin and methemoglobin by pulse oximetry: A human volunteer study. Anesthesiology 105:892897, 2006. 35. Hampson NB, Scott KL: Use of a noninvasive pulse CO-oximeter to measure blood carboxyhemoglobin levels in bingo players. Respir Care 51:758760, 2006. 36. Davis JW, et al: Are automated blood pressure measurements accurate in trauma patients? J Trauma 55:860-863, 2003. 37. Smalling RW, et al: Pre-hospital reduced-dose fibrinolysis coupled with urgent percutaneous coronary intervention reduces time to reperfusion and improves angiographic perfusion score compared with prehospital fibrinolysis alone or primary percutaneous coronary intervention: Results of the PATCAR Pilot Trial. J Am Coll Cardiol 50:16121614, 2007. 38. Brown JP, et al: Effect of prehospital 12-lead electrocardiogram on activation of the cardiac catheterization laboratory and door-to-balloon time in ST-segment elevation acute myocardial infarction. Am J Cardiol 101:158-161, 2008. 39. Ortolani P, et al: Usefulness of prehospital triage in patients with cardiogenic shock complicating ST-elevation myocardial infarction treated with primary percutaneous coronary intervention. Am J Cardiol 100:787-792, 2007. 40. Afolabi BA, et al: Use of the prehospital ECG improves door-to-balloon times in ST segment elevation myocardial infarction irrespective of time of day or day of week. Emerg Med J 24:588-591, 2007. 41. Adams GL, et al: Effectiveness of prehospital wireless transmission of electrocardiograms to a cardiologist via hand-held device for patients with acute myocardial infarction (from the Timely Intervention in Myocardial Emergency, NorthEast Experience [TIME-NE]). Am J Cardiol 98:1160-1164, 2006. 42. Harasawa K, et al: Comparison of tympanic, esophageal and blood temperatures during mild hypothermic cardiopulmonary bypass: A study using an infrared emission detection tympanic thermometer. J Clin Monit 13:19-24, 1997. 43. Rogers IR, et al: Tympanic thermometry is unsuitable as a screening tool for hypothermia after open water swimming. Wilderness Environ Med 18:218221, 2007. 44. Moran JL, et al: Tympanic temperature measurements: Are they reliable in the critically ill? A clinical study of measures of agreement. Crit Care Med 35:155-164, 2007. 45. Woodrow P, et al: Comparing no-touch and tympanic thermometer temperature recordings. Br J Nurs 15:1012-1016, 2006. 46. Incagnoli P, et al: [Resuscitation from accidental hypothermia of 22 degrees C with circulatory arrest: importance of prehospital management]. Ann Fr Anesth Reanim 25:535-538, 2006. 47. Giesbrecht GG: Prehospital treatment of hypothermia. Wilderness Environ Med 12:24-31, 2001. 48. Donald MJ, Paterson B: End tidal carbon dioxide monitoring in prehospital and retrieval medicine: A review. Emerg Med J 23:728-730, 2006. 49. Bhende MS, LaCovey DC: End-tidal carbon dioxide monitoring in the prehospital setting. Prehosp Emerg Care 5:208-213, 2001. 50. Grmec S, Mally S: Prehospital determination of tracheal tube placement in severe head injury. Emerg Med J 21:518-520, 2004. 51. Morley PT: Monitoring the quality of cardiopulmonary resuscitation. Curr Opin Crit Care 13:261-267, 2007. 52. Grmec S, Klemen P: Does the end-tidal carbon dioxide (Etco 2) concentration have prognostic value during out-of-hospital cardiac arrest? Eur J Emerg Med 8:263-269, 2001. 53. Aufderheide TP, Lurie KG: Death by hyperventilation: A common and life-threatening problem during cardiopulmonary resuscitation. Crit Care Med 32(9 Suppl):S345-S351, 2004. 54. Lal D, et al: Prehospital hyperventilation after brain injury: A prospective analysis of prehospital and early hospital hyperventilation of the braininjured patient. Prehospital Disaster Med 18:20-23, 2003. 55. Helm M, et al: Tight control of prehospital ventilation by capnography in major trauma victims. Br J Anaesth 90:327-332, 2003. 56. Hetz H, et al: [Preclinical blood gas analysis. Technical descriptioninitial experiencesindications]. Anaesthesist 45:750-754, 1996. 57. Svensson L, et al: Prehospital identification of acute coronary syndrome/myocardial infarction in relation to ST elevation. Int J Cardiol 98:237-244, 2005. 58. Corbett SW, et al: ED evaluation of the pediatric trauma patient by ultrasonography. Am J Emerg Med 18:244-249, 2000. 59. Walcher F, et al: Prehospital ultrasound imaging improves management of abdominal trauma. Br J Surg 93:238-242, 2006. 60. Salen P, et al: Does the presence or absence of sonographically identified cardiac activity predict resuscitation outcomes of cardiac arrest patients? Am J Emerg Med 23:459-462, 2005. 61. Bochicchio GV, Scalea TM: Is field intubation useful? Curr Opin Crit Care 9:524-529, 2003. 62. Davis BD, et al: Role of rapid sequence induction for intubation in the prehospital setting: helpful or harmful? Curr Opin Crit Care 8:571-577, 2002. 63. Davis DP, et al: Paramedic rapid sequence intubation for severe traumatic brain injury: Perspectives from an expert panel. Prehosp Emerg Care 11:1-8, 2007. 64. Gausche-Hill M: Ensuring quality in prehospital airway management. Curr Opin Anaesthesiol 16:173-181, 2003. 65. Nolan JD: Prehospital and resuscitative airway care: Should the gold standard be reassessed? Curr Opin Crit Care 7:413-421, 2001. 66. Wang HE, et al: Prehospital rapid-sequence intubationwhat does the evidence show? Proceedings from the 2004 National Association of EMS Physicians Annual Meeting. Prehospital Emerg Care 8:366-377, 2004. 67. Wang HE, Yealy DM: Out-of-hospital endotracheal intubation: Where are we? Ann Emerg Med 47:532541, 2006. 68. Dunford JV, et al: Incidence of transient hypoxia and pulse rate reactivity during paramedic rapid sequence intubation. Ann Emerg Med 42:721-728, 2003. 69. Jones JH, et al: Emergency physician-verified outof-hospital intubation: Miss rates by paramedics. Acad Emerg Med 11:707-709, 2004. 70. Timmermann A, et al: Prehospital airway management: A prospective evaluation of anaesthesia trained emergency physicians. Resuscitation 70:179-185, 2006. 71. Timmermann A, et al: The out-of-hospital esophageal and endobronchial intubations performed by emergency physicians. Anesth Analg 104:619-623, 2007. 72. Wirtz DD, et al: Unrecognized misplacement of endotracheal tubes by ground prehospital providers. Prehosp Emerg Care 11:213-218, 2007. 73. Calkins TR, Miller K, Langdorf MI: Success and complication rates with prehospital placement of an esophageal-tracheal Combitube as a rescue airway. Prehosp Disaster Med 21(2 Suppl 2):97-100, 2006. 74. Russi CS, Miller L, Hartley MJ: A comparison of the King-LT to endotracheal intubation and Combitube in a simulated difficult airway. Prehosp Emerg Care 12:35-41, 2008. 75. Guyette FX, et al: Alternate airways in the prehospital setting (resource document to NAEMSP position statement). Prehosp Emerg Care 11:56-61, 2007. 76. Davis DP, et al: The Combitube as a salvage airway device for paramedic rapid sequence intubation. Ann Emerg Med 42:697-704, 2003. 77. Tanigawa K, Shigematsu A: Choice of airway devices for 12,020 cases of nontraumatic cardiac arrest in Japan. Prehosp Emerg Care 2:96-100, 1998. 78. Russi CS, Wilcox CL, House HR: The laryngeal tube device: A simple and timely adjunct to airway management. Am J Emerg Med 25:263-267, 2007.

Anesthesia and Prehospital Emergency and Trauma Care 2331


79. Kette F, et al: The use of laryngeal tube by nurses in out-of-hospital emergencies: Preliminary experience. Resuscitation 66:21-25, 2005. 80. Aufderheide TP, et al: Hyperventilation-induced hypotension during cardiopulmonary resuscitation. Circulation 109:1960-1965, 2004. 81. ONeill JF, Deakin CD: Do we hyperventilate cardiac arrest patients? Resuscitation 73:82-85, 2007. 82. Biros MH, Heegaard W: Prehospital and resuscitative care of the head-injured patient. Curr Opin Crit Care 7:444-449, 2001. 83. Davis DP: Early ventilation in traumatic brain injury. Resuscitation 76:333-340, 2008. 84. Davis DP, et al: The impact of hypoxia and hyperventilation on outcome after paramedic rapid sequence intubation of severely head-injured patients. J Trauma 57:1-8; discussion 8-10, 2004. 85. Davis DP, et al: Ventilation patterns in patients with severe traumatic brain injury following paramedic rapid-sequence intubation. Neurocrit Care 2:165171, 2005. 86. Davis DP, et al: Early ventilation and outcome in patients with moderate to severe traumatic brain injury. Crit Care Med 34:1202-1208, 2006. 87. Thomas SH, et al: Hyperventilation in traumatic brain injury patients: Inconsistency between consensus guidelines and clinical practice. J Trauma 52:47-52; discussion 52-53, 2002. 88. Warner KJ, et al: The impact of prehospital ventilation on outcome after severe traumatic brain injury. J Trauma 62:1330-1336; discussion 1336-1338, 2007. 89. Wang HE, et al: Preliminary experience with a prospective, multi-centered evaluation of out-ofhospital endotracheal intubation. Resuscitation 58:49-58, 2003. 90. Mort TC: Preoxygenation in critically ill patients requiring emergency tracheal intubation. Crit Care Med 33:2672-2675, 2005. 91. Krisanda TJ, et al: An analysis of invasive airway management in a suburban emergency medical services system. Prehosp Disaster Med 7:121-126, 1992. 92. Timmermann A, et al: [Out-of-hospital airway management in northern Germany. Physicianspecific knowledge, procedures and equipment]. Anaesthetist 56:328-334, 2007. 93. Helm M, et al: Factors influencing emergency intubation in the pre-hospital settinga multicentre study in the German Helicopter Emergency Medical Service. Br J Anaesth 96:67-71, 2006. 94. Timmermann A, Russo SG, Hollmann MW: Paramedic versus emergency physician emergency medical service: Role of the anaesthesiologist and the European versus the Anglo-American concept. Curr Opin Anaesthesiol 21:222-227, 2008. 95. Wang HE, et al: Failed prehospital intubations: An analysis of emergency department courses and outcomes. Prehosp Emerg Care 5:134-141, 2001. 96. Henderson JJ, et al: Difficult Airway Society guidelines for management of the unanticipated difficult intubation. Anaesthesia 59:675-694, 2004. 97. Practice guidelines for management of the difficult airway: An updated report by the American Society of Anesthesiologists Task Force on Management of the Difficult Airway. Anesthesiology 98:1269-1277, 2003. 98. Robinson KJ, Katz R, Jacobs LM: A 12-year experience with prehospital cricothyrotomies. Air Med J 20:27-30, 2001. 99. Ricard-Hibon A, et al: Epidemiology of adverse effects of prehospital sedation analgesia. Am J Emerg Med 21:461-466, 2003. 100. Easby J, Dodds C: Emergency induction of anaes thesia in the prehospital setting: A review of the anaesthetic induction agents. Trauma 6:217-224, 2004. 101. Svenson JE, Abernathy MK: Ketamine for prehospi tal use: New look at an old drug. Am J Emerg Med 25:977-980, 2007. 102. Sehdev RS, Symmons D, Kindl K: Ketamine for rapid sequence induction in patients with head injury in the emergency department. Emerg Med Australas 18:37-44, 2006. 103. Cotton BA, et al: Increased risk of adrenal insuffi ciency following etomidate exposure in critically injured patients. Arch Surg 143:62-67; discussion 67, 2008. 104. Fengler BT: Should etomidate be used for rapid sequence intubation induction in critically ill septic patients? Am J Emerg Med 26:229-232, 2008. 105. Davis DP, et al: Paramedic-administered neuromus cular blockade improves prehospital intubation success in severely head-injured patients. J Trauma 55:713-719, 2003. 106. Kill C, et al: [Rapid sequence induction in prehos pital emergency medicine: Is it safe?]. Anasthesiol Intensivmed Notfallmed Schmerzther 39:668-671, 2004. 107. Alonso-Serra HM, Wesley K: Prehospital pain man agement. Prehosp Emerg Care 7:482-488, 2003. 108. Wang HE, et al: Drug-assisted intubation in the prehospital setting (resource document to NAEMSP position statement). Prehosp Emerg Care 10:261271, 2006. 109. Lerner EB, Moscati RM: The golden hour: scientific fact or medical urban legend? Acad Emerg Med 8:758-760, 2001. 110. American College of Surgeons: ATLS Advanced Trauma Life Support Program for Doctors, 7th ed. Chicago, 2004. 111. Davis DP, et al: The effect of paramedic rapid sequence intubation on outcome in patients with severe traumatic brain injury. J Trauma 54:444-453, 2003. 112. Wang HE, et al: How would minimum experience standards affect the distribution of out-of-hospital endotracheal intubations? Ann Emerg Med 50:246252, 2007. 113. Kelly JF, et al: Injury severity and causes of death from Operation Iraqi Freedom and Operation Enduring Freedom: 2003-2004 versus 2006. J Trauma 64:S21-S27, 2008. 114. Lee C, Porter KM, Hodgetts TJ: Tourniquet use in the civilian prehospital setting. Emerg Med J 24:584-587, 2007. 115. Beekley AC, et al: Prehospital tourniquet use in Operation Iraqi Freedom: Effect on hemorrhage control and outcomes. J Trauma 64:S28-S37, 2008. 116. Kragh JF, et al: Practical use of emergency tourni quets to stop bleeding in major limb trauma. J Trauma 64:S38-S50, 2008. 117. Pusateri AE, et al: Making sense of the preclinical literature on advanced hemostatic products. J Trauma 60:674-682, 2006. 118. Brown MA, Daya MR, Worley JA: Experience with chitosan dressings in a civilian EMS system. J Emerg Med 2007; Nov 14 [Epub ahead of print]. 119. Bickell WH, et al: Immediate versus delayed fluid resuscitation for hypotensive patients with penetrating torso injuries. N Engl J Med 331:1105-1109, 1994. 120. Pepe PE, Mosesso VN Jr, Falk JL: Prehospital fluid resuscitation of the patient with major trauma. Prehosp Emerg Care 6:81-91, 2002. 121. Cooper DJ, et al: Prehospital hypertonic saline resuscitation of patients with hypotension and severe traumatic brain injury: A randomized controlled trial. JAMA 291:1350-1357, 2004. 122. Kreimeier U, Messmer K: Small-volume resuscita tion: From experimental evidence to clinical routine: Advantages and disadvantages of hypertonic solutions. Acta Anaesthesiol Scand 46:625638, 2002. 123. Soreide E, Deakin CD: Pre-hospital fluid therapy in the critically injured patienta clinical update. Injury 36:1001-1010, 2005. 124. Moore EE, et al: Hemoglobin-based oxygen carriers in trauma care: Scientific rationale for the US multicenter prehospital trial. World J Surg 30:12471257, 2006. 125. Liberman M, et al: Multicenter Canadian study of prehospital trauma care. Ann Surg 237:153-160, 2003. 126. MacKenzie EJ, et al: A national evaluation of the effect of trauma-center care on mortality. N Engl J Med 354:366-378, 2006. 127. Badjatia N, et al: Guidelines for Prehospital Man agement of Traumatic Brain Injury, 2nd ed. Prehosp Emerg Care 12(1 Suppl 1):1-52, 2008. 128. Stiell IG, et al: Advanced life support for out-of hospital respiratory distress. N Engl J Med 356: 2156-2164, 2007. 129. Fowlkes TD: Shortness of breath. In National Association of EMS Physicians and Kuehl WE (eds): Prehospital Systems and Medical Oversight. Dubuque, IA, Kendall and Hunt, 2002, pp 665-671. 130. Ackerman R, Waldron RL: Difficulty breathing: Agreement of paramedic and emergency physician diagnoses. Prehosp Emerg Care 10:77-80, 2006. 131. Pozner CN, et al: Concordance of field and emer gency department assessment in the prehospital management of patients with dyspnea. Prehosp Emerg Care 7:440-444, 2003. 132. Teboul A, et al: Management of acute dyspnoea: Use and feasibility of brain natriuretic peptide (BNP) assay in the prehospital setting. Resuscitation 61:9196, 2004. 133. Shapiro SE, et al: Determining severe respiratory distress in older out-of-hospital patients. Prehosp Emerg Care 9:310-317, 2005. 134. Expert Panel Report 3 (EPR-3): Guidelines for the Diagnosis and Management of Asthma-Summary Report 2007. J Allergy Clin Immunol 120(5 Suppl 1):S94-S138, 2007. 135. Peter JV, et al: Effect of non-invasive positive pres sure ventilation (NIPPV) on mortality in patients with acute cardiogenic pulmonary oedema: A metaanalysis. Lancet 367:1155-1163, 2006. 136. Mebazaa A, et al: Practical recommendations for prehospital and early in-hospital management of patients presenting with acute heart failure syndromes. Crit Care Med 36(1 Suppl):S129-S139, 2008. 137. Mosesso VN Jr, et al: Prehospital therapy for acute congestive heart failure: State of the art. Prehosp Emerg Care 7:13-23, 2003. 138. 2007 Focused Update of the ACC/AHA 2004 Guidelines for the Management of Patients with STElevation Myocardial Infarction: A report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines. J Am Coll Cardiol 51:210-247, 2008. 139. American Heart Association: Guidelines for cardi opulmonary resuscitation and emergency cardiovascular care. Part 8. Stabilization of the patient with acute coronary syndromes. Circulation 112(24 Suppl):IV-89-IV-110, 2005. 140. The European Myocardial Infarction Project Group: Prehospital thrombolytic therapy in patients with

73

Section V Adult Subspecialty Management

2332 Adult Subspecialty Management


suspected acute myocardial infarction. N Engl J Med 329:383-389, 1993. 141. Brady WJ, Harrigan RA, Chan T: Acute coronary syndromes. In Marx JA (ed): Rosens Emergency Medicine. Philadelphia, Mosby, 2006. 142. Piazza G, Goldhaber SZ: Acute pulmonary embo lism. Part II. Treatment and prophylaxis. Circulation 114:e42-e47, 2006. 143. Janata K, et al: Major bleeding complications in car diopulmonary resuscitation: The place of thrombolytic therapy in cardiac arrest due to massive pulmonary embolism. Resuscitation 57:49-55, 2003. 144. Marik PE, Varon J: Hypertensive crises: Challenges and management. Chest 131:1949-1962, 2007. 145. Broderick J, et al: Guidelines for the management of spontaneous intracerebral hemorrhage in adults: 2007 update: A guideline from the American Heart Association/American Stroke Association Stroke Council, High Blood Pressure Research Council, and the Quality of Care and Outcomes in Research Interdisciplinary Working Group. Circulation 116:e391-e413, 2007. 146. Gray RO: Hypertension. In Marx JA (ed): Rosens Emergency Medicine. Philadelphia, Mosby, 2004. 147. Wolfe RE, Brown DFM: Coma and depressed level of consciousness. In Marx JA (ed): Rosens Emergency Medicine. Philadelphia, Mosby, 2006. 148. Beck PW, Davis E: Altered level of consciousness. In National Association of EMS Physicians and Kuehl AE (eds): Prehospital Systems and Medical Oversight. Dubuque, IA, Kendall and Hunt, 2002, pp 774-782. 149. Millin MG, Gullett T, Daya MR: EMS management of acute strokeout-of-hospital treatment and stroke system development (resource document to NAEMSP position statement). Prehosp Emerg Care 11:318-325, 2007. 150. Kothari RU, et al: Cincinnati Prehospital Stroke Scale: Reproducibility and validity. Ann Emerg Med 33:373-378, 1999. 151. Adams HP Jr, et al: Guidelines for the early manage ment of adults with ischemic stroke: A guideline from the American Heart Association/American Stroke Association Stroke Council, Clinical Cardiology Council, Cardiovascular Radiology and Intervention Council, and the Atherosclerotic Peripheral Vascular Disease and Quality of Care Outcomes in Research Interdisciplinary Working Groups. Circulation 115:e478-e534, 2007. 152. Sayre MR: Damage control: Past, present, and future of prehospital stroke management. Emerg Med Clin North Am 20:877-886, 2002. 153. Crocco TJ, et al: EMS management of acute stroke prehospital triage (resource Document to NAEMSP position statement). Prehosp Emerg Care 11:313317, 2007. 154. Costello DJ, Cole AJ: Treatment of acute seizures and status epilepticus. J Intensive Care Med 22:319347, 2007. 155. Nagele P, Kroesen G: [Pediatric emergencies. An epidemiologic study of mobile care units in Innsbruck]. Anaesthetist 49:725-731, 2000. 156. Suominen P, et al: Pediatric prehospital care pro vided by a physician-staffed emergency medical helicopter unit in Finland. Pediatr Emerg Care 12:169-172, 1996. 157. Bjornson CL, Johnson DW: Croup. Lancet 371:329 339, 2008. 158. Stroud RH, Friedman NR: An update on inflamma tory disorders of the pediatric airway: Epiglottitis, croup, and tracheitis. Am J Otolaryngol 22:268-275, 2001. 159. Biarent D, et al: European Paediatric Life Support Provider Manual, 3rd ed. Antwerp, European Resuscitation Council Secretariat, 2006. 160. Pediatric Advanced Life Support. Part 12. Circula tion 112(24 Suppl):IV-167-IV-187, 2005. 161. Hoshino K, et al: Successful uses of magnesium sulfate for torsades de pointes in children with long QT syndrome. Pediatr Int 48:112-117, 2006. 162. Singh D, et al: A clinical profile of shock in children in Punjab, India. Indian Pediatr 43:619-623, 2006. 163. Chang P, et al: Shock in the pediatric emergency service: Five years experience. Acta Paediatr Taiwan 40:9-12, 1999. 164. Boluyt N, et al: Fluid resuscitation in neonatal and pediatric hypovolemic shock: A Dutch Pediatric Society evidence-based clinical practice guideline. Intensive Care Med 32:995-1003, 2006. 165. Kirkham FJ, Newton CR, Whitehouse W: Paediatric coma scales. Dev Med Child Neurol 50:267-274, 2008. 166. Van de Voorde P, et al: Assessing the level of con sciousness in children: A plea for the Glasgow Coma Motor subscore. Resuscitation 76:175-179, 2008. 167. Oostenbrink R, et al: Signs of meningeal irritation at the emergency department: How often bacterial meningitis? Pediatr Emerg Care 17:161-164, 2001. 168. Oostenbrink R, et al: Prediction of bacterial menin gitis in children with meningeal signs: Reduction of lumbar punctures. Acta Paediatr 90:611-617, 2001. 169. Baumann RJ, Duffner PK: Treatment of children with simple febrile seizures: The AAP practice parameter. American Academy of Pediatrics. Pediatr Neurol 23:11-17, 2000. 170. Appleton R, et al: The treatment of convulsive status epilepticus in children. The Status Epilepticus Working Party, Members of the Status Epilepticus Working Party. Arch Dis Child 83:415-419, 2000. 171. Markenson D, et al: A national assessment of knowledge, attitudes, and confidence of prehospital providers in the assessment and management of child maltreatment. Pediatrics 119:e103e108, 2007. 172. Johnson DR, Dunlap A, Hauswald M, Doezema D: A new approach to teaching prehospital trauma care to paramedic students. Ann Emerg Med 33:51, 1999. 173. Sasser S: Field triage in disasters. Prehosp Emerg Care 10:322-323, 2006. 174. Domres B, et al: Ethics and triage. Prehosp Disaster Med 16:53-58, 2001. 175. Garner A, et al: Comparative analysis of multiple casualty incident triage algorithms. Ann Emerg Med 38:541-548, 2001. 176. Committee on the Future of Emergency Care in the United States Health SystemInstitute of Medicine of the National Academies, Future of Emergency Care: Hospital based emergency careat the breaking point. Washington, DC, The National Academies Press, 2007. 177. Committee on the Future of Emergency Care in the United States Health SystemInstitute of Medicine of the National Academies, Future of Emergency Care: Emergency Care for childrengrowing pains. Washington, DC, The National Academies Press, 2007. 178. Davis DP, et al: The impact of aeromedical response to patients with moderate to severe traumatic brain injury. Ann Emerg Med 46:115-122, 2005. 179. Osterwalder JJ: Can the golden hour of shock safely be extended in blunt polytrauma patients? Prospective cohort study at a level I hospital in eastern Switzerland. Prehosp Disaster Med 17:7580, 2002. 180. Osterwalder JJ: Mortality of blunt polytrauma: A comparison between emergency physicians and emergency medical techniciansprospective cohort study at a level I hospital in eastern Switzerland. J Trauma 55:355-361, 2003. 181. Lossius HM, et al: Prehospital advanced life support provided by specially trained physicians: Is there a benefit in terms of life years gained? Acta Anaesthesiol Scand 46:771-778, 2002. 182. Katz SH, Falk JL: Misplaced endotracheal tubes by paramedics in an urban emergency medical services system. Ann Emerg Med 37:32-37, 2001. 183. Wang HE, et al: Paramedic intubation errors: Isolated events or symptoms of larger problems? Health Aff (Millwood) 25:501-509, 2006. 184. Chi JH, et al: Prehospital hypoxia affects outcome in patients with traumatic brain injury: A prospective multicenter study. J Trauma 61:1134-1141, 2006. 185. Davis DP, et al: The impact of prehospital endotra cheal intubation on outcome in moderate to severe traumatic brain injury. J Trauma 58:933-939, 2005. 186. Cooper A, et al: Prehospital endotracheal intubation for severe head injury in children: A reappraisal. Semin Pediatr Surg 10:3-6, 2001. 187. Ehrlich PF, et al: Endotracheal intubations in rural pediatric trauma patients. J Pediatr Surg 39:13761380, 2004. 188. Gausche M, et al: Effect of out-of-hospital pediatric endotracheal intubation on survival and neurological outcome: A controlled clinical trial. JAMA 283:783-790, 2000. 189. Johnston BD, Seitz SR, Wang HE: Limited opportu nities for paramedic student endotracheal intubation training in the operating room. Acad Emerg Med 13:1051-1055, 2006. 190. Wang HE, et al: Defining the learning curve for paramedic student endotracheal intubation. Prehosp Emerg Care 9:156-162, 2005. 191. Davis DP, Hwang JQ, Dunford JV: Rate of decline in oxygen saturation at various pulse oximetry values with prehospital rapid sequence intubation. Prehosp Emerg Care 12:46-51, 2008.

You might also like