You are on page 1of 10

Orthop Clin N Am 34 (2003) 245 254

Core stability exercise in chronic low back pain


Paul W. Hodges, BPhty (Hons), PhD
Department of Physiotherapy, The University of Queensland, Brisbane, Qld 4072, Australia

Exercise is commonly used in the management of chronic musculoskeletal conditions, including chronic low back pain (CLBP). The focus of exercise is varied and may include parameters ranging from strength and endurance training, to specific training of muscle coordination and control. The assumption underpinning these approaches is that improved neuromuscular function will restore or augment the control and support of the spine and pelvis. In a biomechanical model of CLBP, which assumes that pain recurrence is caused by repeated mechanical irritation of pain sensitive structures [1], it is proposed that this improved control and stability would reduce mechanical irritation and lead to pain relief [1]. Although this model provides explanation for the chronicity of LBP, perpetuation of pain is more complex, and contemporary neuroscience holds the view that chronic pain is mediated by a range of changes including both peripheral (eg, peripheral sensitization) and central neuroplastic changes [2]. Although this does not exclude the role of improved control of the lumbar spine and pelvis in management of CLBP, particularly when there is peripheral sensitization, it highlights the need to look beyond outdated simplistic models. One factor that this information highlights is that the refinement of control and coordination may be more important than simple strength and endurance training for the trunk muscles. The objective of this article is to discuss the rationale for core stability exercise in the management of CLBP, to consider critical factors for

its implementation, and to review evidence for efficacy of the approach.

The rationale for core stability exercise Core stability exercise can be defined loosely as the restoration or augmentation of the ability of the neuromuscular system to control and protect the spine from injury or reinjury. The term is variously used to describe a spectrum of exercise approaches that have the common goal to improve lumbopelvic control, with varied rationales. In general, strategies can be divided into two main groups: those that aim to restore the coordination and control of the trunk muscles to improve control of the lumbar spine and pelvis [3,4], and those that aim to restore the capacity (strength and endurance) of the trunk muscles to meet the demands of control [5,6]. Although both components are necessary, it is useful to consider the rationales for each. The muscle capacity model of core stability exercise is based on the well-established premise that stability of the spine is dependent on the contribution of muscle. The contemporary view of spinal stability is based on the Euler model that considers the control of buckling forces [7 9]. This is based on the understanding that buckling failure of the lumbar spine, devoid of muscle, occurs with compressive loading of as little as 90 N. This model argues that muscle activity is required to act like guy wires to stiffen intervertebral joints that they span to maintain the lumbar spine in a mechanically stable equilibrium [7 9]. This definition is relatively static and suggests the maintenance of a set position of the spine. Few studies have considered this model in more dynamic terms [10]. Consistent with the proposal that muscle

Financial support was provided by the National Health and Medical Research Council of Australia and The Swedish Research Council. E-mail address: p.hodges@shrs.uq.edu.au

0030-5898/03/$ see front matter D 2003, Elsevier Inc. All rights reserved. doi:10.1016/S0030-5898(03)00003-8

246

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254

activity is required to control stability and restrict trunk motion; studies have identified increased activity [11,12] and coactivity of antagonist trunk muscles [13], presumably to augment the control of the spine in individuals with CLBP. On the basis of this philosophy, a range of exercise strategies has been developed to overcome the buckling force and augment the stability of the spine. In general, these strategies involve exercises that encourage maintenance of a stable lumbopelvic position, while forces are applied with limb movements and loads. The control model of core stability exercise is based on the premise that lumbopelvic function and health are dependent on the accurate interplay of the trunk muscles. Stability and control of the spine is dependent not only on the muscles but also on the central nervous system (CNS) which must determine the requirements of stability to plan and implement strategies that meet these demands [1]. When the CNS can predict that the control of the spine will be challenged by an internal or external force (eg, as a result of the reactive forces from a moving limb), it has the capacity to plan strategies of muscle activity in advance of the movement to prepare the spine (ie, feedforward control). In other cases, when the perturbation is unexpected, the CNS can initiate a response of the trunk muscles [14,15]. Thus, accurate control of the lumbar spine and pelvis is dependent not only on the capacity of muscle, but also on the sensory system that provides information about the status of stability, recognition of perturbation, and the development of the internal model of body dynamics that provides the CNS with the capacity to predict the outcome of mechanical events and predict the interaction between body segments and the environment. Accurate control is also dependent on the capacity of the motor system to plan the appropriate strategy to meet the demands of stability. This system is not only responsible for the control of buckling forces, but also for the control of the overall orientation of the spine and pelvis, and intervertebral translation and rotation during trunk movement. Although the control of these two elements cannot be completely separated, particular attention must be paid to control of translations and rotations. For instance, during an arc of movement, it is important to control the coordination between translation and rotation at the intervertebral levels [16]. It has been shown that, if the spine is modeled with one segment with no muscle attachment, the spine is as stable as having no muscle at all [8]. Thus, segmental control is an essential component for spinal stability. There is an increasing body of literature that has outlined the normal strategies for control

of the spine, as well as changes in this control when people have LBP. As a result, many contemporary approaches to core stability exercise focus on rehabilitation of the control of the trunk muscles. In many cases, this focus has been on retraining the function of the deep intrinsic muscles of the lumbar spine and pelvis (eg, transversus abdominis and lumbar multifidus), and then on integration of the activity of the deep and superficial trunk muscles in functional tasks. The hypothesis here is that the coordinated activity of the deep spinal muscles plays an important role in the fine-tuning of intersegmental motion of the spine and pelvis (eg, sacroiliac joints). Though these muscles may not provide the largest contribution to stability, their function is essential for optimal spinal health. It is impossible to separate these two rationales completely, but there are differences in focus that have lead to variations in the exercises incorporated into a core stability regime. The following section addresses the neuromuscular control of stability to provide greater insight into the clinical decisions for implementation of core stability exercise in the management of CLBP.

Normal and abnormal control of the lumbar spine and pelvis Many muscles span the lumbopelvic region and contribute to its control and stability. In fact, all muscles are required for this function. There is considerable biomechanical and motor control evidence, however, to argue that different muscles and different control strategies may be involved in the control of different elements of stability, ie, intervertebral control, and the control or buckling forces and orientation of the lumbar spine and pelvis. The motor system has considerable redundancy, with many muscles capable of performing similar functions. There is variation, however, in the architectural properties of the trunk muscles; this has led to the proposal of functional differentiation in the muscle system. It has been argued that muscles are suited biomechanically to either motion or stability [3,4,17,18]. Though this oversimplifies the complex control of spinal stability and motion, there is biomechanical evidence to argue that the contribution of muscles differs. Bergmark [17] presented a model for the trunk that considered differentiation in the contribution of muscle to stability. This model identified muscles as either local or global, based on anatomic characteristics. The local muscles are those that cross one

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254

247

or a few segments and that have a limited moment arm to move the joint, but an ideal anatomy to control intervertebral motion. Bergmark included muscles such as the lumbar multifidus in this group; however, other muscles that satisfy these criteria are transversus abdominis (TrA), intertransversarii and interspinales, posterior fibers of psoas [19], and medial fibers of quadratus lumborum [20]. Obliquus internus has fibers that parallel the TrA and may contribute to the force closure of the sacroiliac joint (SIJ) [21]. By contrast, the global muscles cross several joints with attachments to the pelvis and the thorax. These muscles have a larger moment arm (ie, greater torque-generating capacity) and are suited to the control of orientation and external forces. Examples of the global muscles include: rectus abdominis, obliquus externus and internus abdominis, thoracic portions of the longissimus and iliocostalis, lateral fibers of quadratus lumborum, anterior fibers of psoas, and latissimus dorsi. There is overlap between these systems, and some muscles share features of both, such as the lumbar longissimus and iliocostalis, and the superficial fibers of multifidus having one attachment to the lumbar vertebrae and sharing some features of the local system. Two local or deep intrinsic lumbopelvic muscles have received attention specifically in the literature: the TrA and the lumbar multifidus. TrA is a broad sheet-like muscle with extensive attachments to the lumbar vertebrae via the thoracolumbar fascia, and to the pelvis and rib cage. The muscle fibers have a relatively horizontal orientation and, therefore, minimal ability to move the spine, although they may contribute to rotation [22,23]. Contribution to spinal control is likely to involve modulation of intraabdominal pressure (IAP) and tensioning the thoracolumbar fascia. TrA has been shown to be associated closely with control of IAP [22,24], and recent data confirm that spinal stiffness is increased by IAP [25,26]. Fascial tension may restrict intervertebral motion directly or provide gentle segmental compression via the posterior layer of the thoracolumbar fascia [27]. Recent porcine studies confirm that the combined effect of IAP and fascial tension is required for TrA to increase intervertebral stiffness, and the mechanical effect of its contraction is reduced if the fascial attachments are cut [25]. For sacroiliac support, TrA acts via the ilia to compress the SIJ anteriorly [21], and this finding has been confirmed in vivo [28]. The lumbar multifidus has five fascicles that arise from the spinous process and lamina of each lumbar vertebra and descend in a caudo-lateral direction [29]. The most superficial fibers cross up to five segments

and attach caudally to the ilia and sacrum, whereas the deep fibers attach from the inferior border of a lamina and cross a minimum of two segments to attach on the mamillary process and facet joint capsule [30]. The superficial fibers are distant from the centers of rotation of the lumbar vertebrae, have an extension moment arm, and can control the lumbar lordosis [29]. By contrast, the deep fibers have a limited moment arm and a minor ability to extend the spine [31]. Multifidus can control intervertebral motion by generation of intervertebral compression [32]. The proximity of deep multifidus to the center of rotation results in compression with minimal extension moment. In addition, multifidus may contribute to the control of intervertebral motion by control of anterior rotation and translation of the vertebrae [29], or by tensioning the thoracolumbar fascia as it expands on contraction [33]. Several studies have provided in vitro and in vivo evidence of the ability of multifidus to control intervertebral motion [32,34]. The contribution of the superficial global muscles to lumbopelvic movement and stability generally is predictable based on the moment arm and the direction of force provided by the muscles [17,35]. That is, flexors generate flexion torque and oppose extension. In addition, it generally has been considered that antagonist trunk muscles are coactivated to stiffen the spine and prevent buckling [36]. For example, greater activity of the superficial abdominal muscles has been demonstrated during isometric trunk tasks than predicted by a biomechanic model [35], and antagonistic activity of obliquus externus (OE) and internus abdominus (OI) has been recorded during lifting and isometric trunk efforts [35]. Optimal function of both systems is required to maintain spinal function. The local system has only a limited ability to influence the control of orientation, and the global system has only a limited ability to control intervertebral motion. It is important to consider how these muscles are coordinated in function. Studies of feedforward and feedback control of the trunk suggest that the CNS uses separate strategies for the control of intervertebral motion and orientation of the spine. Initial studies of the deep muscles investigated locomotion and trunk movements [22,37]. In these tasks, the superficial abdominal and extensor muscles were active in a manner that was specific to direction. TrA was tonically active, however, and not influenced by the direction of movement [22,37]. Precise evidence of this differential control has come from investigation of predictable challenges to spinal stability caused by rapid limb movements. Early studies of arm movements showed activity of leg muscles and erector spinae before flexion of the arm

248

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254

[38]. More recently, studies have confirmed that TrA and the deep fibers multifidus are activated as a component of this anticipatory response [39 41]. Though the response of the superficial muscles is linked to the direction of force [40,42], the activity of TrA and deep multifidus is independent of the direction of force [40 42]. That is, the superficial muscles control the orientation of the spine, whereas the deep muscles provide control of intersegmental motion that is not specific to the direction of force. An important additional finding of these studies is that the CNS does not make the spine rigid by coactivation of the superficial muscles but, instead, uses controlled movement to help counteract the applied forces [42]. The differential control of the deep and superficial muscles has been confirmed in experiments that have manipulated task expectation [43]. Thus, feedforward studies indicate that the CNS uses specialized strategies for each element of spinal control (ie, nondirection specific, early, tonic coactivation of the deep muscles and direction specific, phasic activation of the superficial muscles). This is supported by studies of feedback-mediated control. Many studies have investigated activity of the superficial muscles in response to external perturbations such as a load dropped into a box held in the hands in front of the body [44], translation of the support surface [45], or trunk loading [24]. These studies report direction-specific activity of the superficial muscles to maintain spinal orientation. By contrast, studies of TrA report activity irrespective of the direction of force [24]. Recent data indicate that the deep and superficial fibers of multifidus are differentially activity in response to trunk loading [15]. This was only apparent, however, when the loading was predictable. Many studies report changes in motor control in people with acute and chronic LBP. Though there is considerable variability in results, differential changes in activity of deep and superficial muscles have been relatively consistent. In terms of deep muscle activity, there is evidence of delayed activity of TrA in association with rapid limb movements in CLBP [46]. The changes in TrA have been replicated when pain is induced by intramuscular injection of hypertonic saline into the longissimus muscle [47]. Notably in studies of CLBP the changes were identified in people who had a history of LBP but were in remission from their symptoms. Although these studies have reported a delay in activation, it is likely that the change is not confined to this parameter but, instead, may be reflective of a general change in control. For example, tonic activity of TrA, which is normally observed during repetitive trunk [22] and limb movements [48], is

reduced during experimental pain [47], and relative EMG activity of the rectus abdominis and the inferolateral abdominal wall is altered in CLBP [40]. There is preliminary evidence that the deep paraspinal muscles show similar changes in activity. During functional tasks, there is reduced amplitude of activity of the multifidus in CLBP [50], and altered responses have been observed during loading of the trunk. For example, when a load is unexpectedly dropped into the hands, there is normally a shortlatency response of the paraspinal muscles [15,51]; in healthy control subjects, when the loading can be anticipated, the response of the deep fibers of multifidus occurs earlier [15]. When people with sciatica catch a load that is predictable, however, the earlier response of the paraspinal muscles does not occur [51]. Others, using an unexpected loading paradigm, report both delayed response [44] and no change [12] in paraspinal muscle activity. These changes in activity are consistent with changes in the morphology and fatigability of the muscle, which could be explained by altered use of the muscle. For example, studies report changes in muscle fiber composition [52], increased fatigability [53], and a reduced cross-sectional area that occurs as little as 24 hours after the onset of acute, unilateral LBP [54]. In CLBP, there is evidence of fatty infiltration into the paraspinal muscles [55]. There is a large body of literature that investigates changes in the superficial trunk muscles in LBP. Although the changes have been studied extensively, there is considerable disagreement as to whether the superficial muscles have reduced endurance and strength capacity [56,57]. It has been suggested that these changes may be more related to inactivity than to pain [57]. Likewise, studies of superficial paraspinal muscle activity have had variable results, with some reporting increased activity [58], others reporting decreased [50] or asymmetrical activity [59], and still others reporting no change in activity [60]. One finding that has been observed consistently in people with LBP is sustained activity of the erector spinae muscles at the end of the range of spinal flexion a point at which the erector spinae muscles are normally inactive (the so-called flexion-relaxation response) [61]. This finding has been replicated by experimental pain [12] and has been shown to limit intervertebral motion [62]. The normal periods of silence in erector spinae activity between heel contacts are reduced, furthermore, in LBP patients and in otherwise asymptomatic participants given experimentally induced LBP [11], which may serve to splint the area during this period. Others have identified delayed reduction of activity and increased coactivation of superficial muscles in response to

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254

249

removal of a load from the trunk in a subset of CLBP [13]. Experimentally induced pain caused variable responses of the superficial trunk muscles in association with rapid limb movements [47]. Importantly, however, although there was considerable intersubject variability in the pattern of superficial trunk muscle activity, at least one superficial muscle was augmented during pain in every subject. Notably, despite the variation in superficial muscle activity, hypoactivity of the intrinsic spinal muscle, TrA, was a consistent finding across the group in that study. In addition to changes in muscle recruitment, impairment of other elements of the sensorimotor system has been identified in CLBP that may affect the normal control of the lumbar spine and pelvis. For example, studies have reported reduced acuity [63] and impaired ability to perform repositioning tasks [64]. Because both feedforward- and feedback-mediated components of motor control are dependent on sensory input, any change in sensory input is likely to be important. In summary, the evidence suggests that, with LBP, the activity of the deep and superficial muscles are affected differentially. Although the results are variable, there is evidence for augmented activity of at least one of the superficial muscles in CLBP, whereas the control of the deep intrinsic spinal muscles is invariably impaired irrespective of the pathology. Although the increase in superficial trunk muscle activity can prevent lumbar spine buckling and increase spinal stiffness [9,17], it is also associated with increased compressive loading of the spine [36], which has long been considered a risk factor for spinal degeneration and pain. It is known, moreover, that the CNS uses movement rather than stiffening of the spine to overcome challenges to stability [42] and reduce energy expenditure. A strategy of trunk stiffening, although requiring less complex neural control, may compromise optimal spinal loading. Finally, the response to pain of stiffening the spine appears to be at the cost of a loss of the fine tuning of intersegmental motion of the spine provided by the deep muscles. The mechanism of these changes is not completely understood, nor whether the changes precede the pain or are a result of the painful episode. Though it is not possible to exclude the possibility that motor control changes may predispose an individual to pain, recent evidence argues that pain may replicate the changes seen in clinical populations [36]. The mechanism for the effect of pain may involve factors such as changes in excitability along the motor pathway or more complex changes caused by fear-avoidance behaviors. This topic is reviewed elsewhere [65].

Clinical factors in the implementation of core stability exercise From the preceding discussion, it is clear that the specific nature of the core stability exercise program can vary depending on whether the focus is on control or muscle capacity. In reality all components require consideration, and a specific intervention must be tailored to the individual patients presentation. This section addresses factors to consider in the implementation of the spectrum of exercise interventions that fall under the umbrella of core stability exercise. In terms of retraining the control of core stability, evidence thus far suggests that there are specific changes in the strategies used by the CNS to control the spine, and that this consistently involves impaired activity of the deep muscle system, often in association with overactivity of one or more superficial muscles. It is unlikely that general exercise for the trunk such as sit-ups and back extension exercises would restore the coordination between the trunk muscles, and a strategy based on the principles of motor relearning and skill acquisition is required. This approach aims to train the skilled activation of the deep muscles, to train the integration of the deep and superficial systems, and to progress through a program of tailored functional exercises in varying environments and contexts to ensure transfer to normal activity [4]. The nervous system has considerable potential for plasticity and learning. A key strategy described in the literature to retrain motor function involves practice of parts of movement rather than the whole movement [66]. When a skill is trained in parts, the attention demand is reduced to allow attention to be focused on a single element. Several techniques have been described for part-task training. These include segmentation and simplification [66]. For segmentation, the task is divided into smaller parts, to be practiced as an independent unit, and then the practiced elements are integrated together progressively to practice the complete skill. A key feature of this strategy is selection of the specific features of a movement that are impaired or dysfunctional (ie, essential components), and then implementation of strategies that optimize the performance of that component using cognitive strategies. At a later stage, it is important to perform the interdependent parts of a task together, and to integrate the trained components in a functional context [67]. From the evidence presented above, the component of movement that is impaired in normal function is the activity of the deep muscle system (the activity of this system is delayed, more phasic, and no longer independent of the

250

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254

superficial muscles). Thus, the focus of initial stages of rehabilitation is to train this component, independently from the superficial muscles. Simplification refers to strategies to increase the ease of movement performance and may be achieved by changing parameters such as postural load (eg, training in supported positions), reduction of attention demands, reduction of speed, or additional strategies to augment performance (eg, improved accuracy of feedback with ultrasound imaging). Motor learning occurs in three main phases: cognitive, associative, and autonomous [68]. In the cognitive phase, the focus is on cognitively oriented problems. All elements of the movement performance are organized consciously with attention to feedback, movement sequence, performance, and instruction during repetition and practice. In motor relearning for LBP, the goal of the initial phase of motor relearning is to contract the deep muscles cognitively to increase the precision and skill of the contraction of the local muscles. Once mastered, the goal shifts to increased precision, increased number of repetitions and holding time, and decreased feedback. In the associative phase, the fundamentals of the skill have been acquired, and the cognitive demands reduced. The focus moves from simple elements of task performance to consistency of performance, success, and refinement. In this stage, many repetitions are required in a variety of contexts to reduce the cognitive demand of the task. In terms of trunk muscle training, this phase involves performance of the task in increasingly challenging positions (eg, sitting and standing) and integration of deep and superficial muscle function, eg, using leg-loading tasks [18], proprioceptive neuromuscular facilitation techniques [69], and postural challenges [3]. The final stage of motor learning, the autonomous phase, is achieved after considerable practice and experience. The task becomes habitual or automatic, and the requirement for conscious intervention is reduced. Importantly, there is preliminary evidence that training in this manner can result in a change in the automatic control of the spine [70,71] and peripheral joints [72]. An important element of motor learning is the provision of augmented feedback. Feedback can be divided generally into two main types: knowledge of performance and knowledge of results [73]. Put simply, feedback that provides knowledge of performance relates to ongoing information provided during movement, whereas knowledge of results is feedback of the outcome of the task. In the cognitive phase, it is critical to provide accurate feedback of contraction quality. This feedback may be intrinsic

(naturally occurring), or augmented in some way, and may involve any of the senses including tactile (palpation) visual (ultrasound imaging) [74], and auditory (EMG biofeedback) information. A critical component of skill learning is that the performance of the skill can be transferred to different conditions in which the environment, personal characteristics, or predictability are changed. To optimize transfer, it is considered essential to progress the task sequentially from easy to more complex situations. If the aim is to transfer a skilled movement to functional tasks, then it is necessary to progress to function. More specifically, it may be necessary to replicate sensory characteristics (eg, limitation of visual feedback), environmental contexts (eg, unstable surfaces, distractions), and personal contexts (eg, anxiety, fatigue) [75] to ensure that the elements of the skill can be transferred to specific contexts [66]. Another issue to consider is the importance of the neutral position of the spine, which is variously described as being the normal lumbar, cervical lordosis, and thoracic kyphosis. Around the neutral position, the spine exhibits the least stiffness [1]. This region is important to consider because stability is dependent on the contribution of the trunk muscles. In addition, it has been argued that loading through the spine is optimal in neutral positions [5]. Clinically, training trunk-muscle control in neutral positions has been argued to be important for effective management of CLBP [4,6]. In summary, in a motor control model to progress a patient with LBP through the normal phases of motor learning, the basic sequence that needs to be undertaken is: (1) skill learning, (2) precision training, (3) activation in a variety of contexts (including a variety of postures and positions), (4) integration of the skill into tasks that includes activation of the superficial trunk muscles, and (5) specific functional retraining to ensure that the appropriate coordination of deep and superficial trunk muscles is maintained in a functional context. In contrast with the motor relearning approach to retraining the control and coordination of the trunk muscles, at the other end of the spectrum of core stability exercise are tasks aimed at restoring the ability of the trunk muscles themselves to meet the demands for spinal stability. Notably, as the magnitude of force acting on the trunk is increased, the relative contribution of the superficial trunk muscles is increased [5]. Thus, the system must be trained to meet the demands of high-level activities. The strategies employed overlap with those described above. Though these tasks form the central component of some core stability exercise programs, they may be

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254

251

used in a motor control model as progressive training to rehabilitate the integration of the local and global systems for higher-level activities. Many exercise strategies have been developed for this component of core stability exercise. Strategies that are consistent with this approach include Pilates exercise, dissociation exercises [18], limb movements with a neutral lumbar spine and pelvis [4], Swiss ball programs [76], McGills stability exercises [5], rhythmic stabilization exercises from proprioceptive neuromuscular facilitation [69], balance board and balance shoe training [77], floor exercise programs [6], and so on. The key features of exercise vary between approaches but include features such as control of the lumbopelvic position (particularly in neutral) during movement of the limbs [4,6,18,76], and on unstable surfaces [76,77], closed chain tasks [4], and resisted movements [78].

Efficacy of core stability exercise in the management of chronic LBP There are an increasing number of clinical studies that have investigated the efficacy of core stability exercise in the management of CLBP. Although early studies were based on audits of clinical outcome [6], more recently, high quality randomized controlled clinical trials have been conducted. The first study investigated core stability exercise, from a motor control perspective, in people with CLBP associated with spondylolisthesis [79]. In this study, subjects were allocated randomly to either participate in a motor-relearning program or a nontreatment group. The training period lasted for 10 weeks. At the completion of training and at follow-up at 30 months, there was a significant reduction in pain and disability in the motor-relearning group. There was no significant change in the nontreatment group. The second study involved training in acute first-episode unilateral LBP [80]. This group was selected because they have a reduced cross-sectional area of multifidus ipsilateral to their symptoms [54 ]. The intervention involved a 4-week program of motor relearning, focused on multifidus in conjunction with TrA. After 4 weeks, all pain and disability measures had recovered in all but one participant. This is consistent with epidemiologic data. The size of multifidus had recovered only in the motor-control training group [80]. The follow-up data provide potent evidence for the efficacy of the approach. After 3 years, people in the control group were 12.4 times more likely to have further episodes of pain that those in the exercise group [81]. Although the data deal with an acute

population, they are important to consider here because recurrence of pain is a major factor in CLBP. Other studies of chronic headache [70] and patellofemoral pain [72], using a similar clinical paradigm, have shown positive clinical outcomes. Additional studies of CLBP [82] and pelvic pain [83] have promising early results. Another focus of clinical studies has been to identify the mechanism of efficacy of the clinical approach. In general, studies of muscle size [84], muscle strength, muscle endurance, and muscle fiber composition [85] have failed to find improvement or a relationship with the clinical outcome. Improvements in motor control parameters such as muscle activation patterns [49,71,72] and performance of specific skilled activities [70] have found a positive relationship, however. Thus, there is increasing evidence of efficacy of core stability exercise, particularly from a motor-control perspective, and there is evidence that the improvements are related to the factor being addressed in the intervention, not a tertiary cause.

Summary In conclusion, core stability exercise is an evolving process, and refinement of the clinical rehabilitation strategies is ongoing. Two major foci are addressed in contemporary core stability programs: motor control and muscle capacity. Both of these factors have considerable foundation in the literature and can be seen as a progression of exercise rather than conflicting approaches. Importantly, the clinical efficacy of these approaches is being realized in clinical trials. Further work is required, however, to refine and validate the approach, particularly with reference to contemporary understanding of the neurobiology of chronic pain.

References
[1] Panjabi MM. The stabilizing system of the spine. Part 1. Function, dysfunction, adaptation, and enhancement. J Spinal Disord 1992;5(4):383 9. [2] Butler DS. Sensitive nervous system. Adelaide, Australia: NOIgroup Publications; 2000. [3] Janda V. Muscles, central nervous motor regulation and back problems. In: Korr IM, editor. The neurobiologic mechanisms in manipulative therapy. New York: Plenium Press; 1978. p. 27 41. [4] Richardson CA, Jull GA, Hodges PW, Hides JA. Therapeutic exercise for spinal segmental stabilization in LBP: scientific basis and clinical approach. Edinburgh: Churchill Livingstone; 1999.

252

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254 [22] Cresswell AG, Grundstrom H, Thorstensson A. Observations on intra-abdominal pressure and patterns of abdominal intra-muscular activity in man. Acta Physiol Scand 1992;144:409 18. [23] Urquhart D, Story I, Hodges P. Rotation transversus abdominis. In: Proceedings of the 7th International Physiotherapy Conference, Sydney, Australia. Melbourne: Australian Physiotherapy Association; 2002. p. 45. [24] Cresswell AG, Oddsson L, Thorstensson A. The influence of sudden perturbations on trunk muscle activity and intra-abdominal pressure while standing. Exp Brain Res 1994;98(2):336 41. [25] Hodges P, Kaigle-Holm A, Holm S, Ekstrom L, Cresswell AG, Hansson T, et al. In vivo evidence that postural activity of the respiratory muscles increases intervertebral stiffness of the spine: porcine studies. Soc Neurosci Abstr 2002;28:366. [26] Hodges PW, Cresswell AG, Daggfeldt K, Thorstensson A. In vivo measurement of the effect of intra-abdominal pressure on the human spine. J Biomech 2001;34: 347 53. [27] Gracovetsky S, Farfan H, Helleur C. The abdominal mechanism. Spine 1985;10(4):317 24. [28] Richardson CA, Snijders CJ, Hides JA, Damen L, Pas MS, Storm J. The relation between the transversus abdominis muscles, sacroiliac joint mechanics, and LBP. Spine 2002;27(4):399 405. [29] Macintosh JE, Bogduk N. The detailed biomechanics of the lumbar multifidus. Clin Biomech 1986;1:205 31. [30] Lewin T, Moffett B, Viidik A. The morphology of the lumbar synovial joints. Acta Morphologica Neerlanco Scandinav 1962;4:299 319. [31] Panjabi MM, Abumi K, Duranceau J, Oxland T. Spinal stability and intersegmental muscle forces. A biomechanical model. Spine 1989;14:194 200. [32] Wilke HJ, Wolf S, Claes LE, Arand M, Wiesend A. Stability increase of the lumbar spine with different muscle groups: a biomechanical in vitro study. Spine 1995;20(2):192 8. [33] Gracovetsky S, Farfan HF, Lamy C. A mathematical model of the lumbar spine using an optimised system to control muscles and ligaments. Orthop Clin North Am 1977;8:135 53. [34] Kaigle AM, Holm SH, Hansson TH. Experimental instability in the lumbar spine. Spine 1995;20(4):421 30. [35] Zetterberg C, Andersson GB, Schultz AB. The activity of individual trunk muscles during heavy physical loading. Spine 1987;12:1035 40. [36] Gardner-Morse MG, Stokes IA. The effects of abdominal muscle co-activation on lumbar spine stability. Spine 1998;23(1):86 91. [37] Pauly J. An electromyographic analysis of certain movements and exercises: I. Some deep muscles of the back. Anat Rec 1966;155:223 34. [38] Belenkii V, Gurfinkel VS, Paltsev Y. Elements of control of voluntary movements. Biofizika 1967;12(1): 135 41. [39] Hodges PW, Richardson CA. Contraction of the ab-

[5] McGill S. Low back disorders: evidence based prevention and rehabilitation. Champaign, IL: Human Kinetics Publishers, Inc.; 2002. [6] Saal JA, Saal JS. Non-operative treatment of herniated lumbar intervertebral disc with radiculopathy: an outcome study. Spine 1989;14:431 7. [7] Cholewicki J, McGill SM. Mechanical stability of the in vivo lumbar spine: implications for injury and chronic low back pain. Clin Biomech 1996;11(1): 1 15. [8] Crisco JJ, Panjabi MM. The intersegmental and multisegmental muscles of the lumbar spine: a biomechanical model comparing lateral stabilising potential. Spine 1991;7:793 9. [9] Gardner-Morse M, Stokes IAF, Laible JP. Role of muscles in lumbar spine stability in maximum extension efforts. J Orthopaed Res 1995;13:802 8. [10] Cholewicki J, Panjabi MM, Khachatryan A. Stabilizing function of trunk flexor-extensor muscles around a neutral spine posture. Spine 1997;22(19):2207 12. [11] Arendt-Nielsen L, Graven-Nielsen T, Svarrer H, Svensson P. The influence of low back pain on muscle activity and coordination during gait. Pain 1996;64(2): 231 40. [12] Zedka M, Prochazka A, Knight B, Gillard D, Gauthier M. Voluntary and reflex control of human back muscles during induced pain. J Physiol 1999;520(Pt 2): 591 604. [13] Radebold A, Cholewicki J, Panjabi MM, Patel TC. Muscle response pattern to sudden trunk loading in healthy individuals and in patients with chronic low back pain. Spine 2000;25(8):947 54. [14] Indahl A, Kaigle AM, Reikeras O, Holm SH. Interaction between the porcine lumbar intervertebral disc, zygapophysial joints, and paraspinal muscles. Spine 1997;22(24):2834 40. [15] Moseley GL, Hodges PW, Gandevia SC. External perturbation of the trunk in standing humans differentially activates components of the medial back muscles. J Physiol 2002; in press. [16] Bogduk N, Amevo B, Pearcy M. A biological basis for instantaneous centers of rotation of the vertebral column. Proc Inst Mech Eng 1995;209(3):177 83. [17] Bergmark A. Stability of the lumbar spine. A study in mechanical engineering. Acta Orthoped Scand 1989; 60(suppl 230):1 54. [18] Sahrman S. Diagnosis and treatment of movement impairment syndromes. St Louis: Mosby, Inc.; 2002. [19] Bogduk N, Pearcy M, Hadfeild G. Anatomy and biomechanics of psoas major. Clin Biomech 1992;7: 109 19. [20] McGill S, Juker D, Kropf P. Quantitative intramuscular myoelectric activity of quadratus lumborum during a wide variety of tasks. Clin Biomech 1996;11(3): 170 2. [21] Snijders CJ, Vleeming A, Stoeckart R, Mens JMA, Kleinrensink GJ. Biomechanical modelling of sacroiliac joint stability in different postures. Spine: State Art Rev 1995;9:419 32.

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254 dominal muscles associated with movement of the lower limb. Phys Ther 1997;77:132 44. Hodges PW, Richardson CA. Feedforward contraction of transversus abdominis in not influenced by the direction of arm movement. Exp Brain Res 1997;114: 362 70. Moseley GL, Hodges PW, Gandevia SC. Deep and superficial fibers of lumbar multifidus are differentially active during voluntary arm movements. Spine 2002; 27:E29 36. Hodges PW, Cresswell AG, Thorstensson A. Preparatory trunk motion accompanies rapid upper limb movement. Exp Brain Res 1999;124:69 79. Hodges PW, Richardson CA. Transversus abdominis and the superficial abdominal muscles are controlled independently in a postural task. Neurosci Lett 1999; 265(2):91 4. Wilder DG, Aleksiev AR, Magnusson ML, Pope MH, Spratt KF, Goel VK. Muscular response to sudden load. A tool to evaluate fatigue and rehabilitation. Spine 1996;21(22):2628 39. Henry SM, Fung J, Horak FB. EMG responses to maintain stance during multidirectional surface translations. J Neurophysiol 1998;80(4):1939 50. Hodges PW, Richardson CA. Delayed postural contraction of transversus abdominis associated with movement of the lower limb in people with LBP. J Spinal Disord 1998;11(11):46 56. Hodges PW, Moseley GL, Gabrielsson A, Gandevia SC. Experimental muscle pain changes feed forward postural responses of the trunk muscles. Exp Brain Res 2003; in press. Hodges P, Gandevia S. Changes in intra-abdominal pressure during postural and respiratory activation of the human diaphragm. J Appl Physiol 2000;89(3): 967 76. OSullivan PB, Twomey L, Allison GT. Altered abdominal muscle recruitment in patients with CLBP following a specific exercise intervention. J Orthop Sports Phys Ther 1998;27(2):114 24. Sihvonen T, Lindgren KA, Airaksinen O, Manninen H. Movement disturbances of the lumbar spine and abnormal back muscle electromyographic findings in recurrent LBP. Spine 1997;22:289 95. Leinonen V, Kankaanpaa M, Luukkonen M, Hanninen O, Airaksinen O, Taimela S. Disc herniation-related BP impairs feed-forward control of paraspinal muscles. Spine 2001;26(16):E367 72. Rantanen J, Hurme M, Falck B, Alaranta H, Nykvist F, Lehto M, et al. The lumbar multifidus muscle five years after surgery for a lumbar intervertebral disc herniation. Spine 1993;18(5):568 74. Roy SH, DeLuca CJ, Casavant DA. Lumbar muscle fatigue and chronic low back pain. Spine 1989; 14(9): 992 1001. Hides JA, Stokes MJ, Saide M, Jull GA, Cooper DH. Evidence of lumbar multifidus muscle wasting ipsilateral to symptoms in patients with acute/subacute low back pain. Spine 1994;19:165 77.

253

[40]

[41]

[42]

[43]

[44]

[45]

[46]

[47]

[48]

[49]

[50]

[51]

[52]

[53]

[54]

[55] Alaranta H, Tallroth K, Soukka A, Heliaara M. Fat content of lumbar extensor muscles in low back disability: a radiographic and clinical comparison. J Spinal Disord 1993;6(2):137 40. [56] Suzuki N, Ohe K. Abdominal and back muscle strength in patients with low back pain. Orthop Surg 1978;29:325 8. [57] Thorstensson A, Arvidson A. Trunk muscle strength and low back pain. Scand J Rehabil Med 1982;14: 69 75. [58] Arena JG, Sherman RA, Bruno GM, Young TR. Electromyographic recordings of 5 types of low back pain subjects and non-pain controls in different positions. Pain 1989;37(1):57 65. [59] Cram JR, Steger JC. EMG scanning in the diagnosis of chronic pain. Biofeedback Self Regul 1983;8:229 41. [60] Collins GA, Cohen MJ, Naliboff BD, Schandler SL. Comparative analysis of paraspinal and frontalis EMG, heart rate and skin conductance in chronic low back pain patients and normals to various postures and stresses. Scand J Rehabil Med 1982;14:39 46. [61] Shirado O, Ito T, Kaneda K, Strax TE. Flexion-relaxation phenomenon in the back muscles. A comparative study between healthy subjects and patients with chronic low back pain. Am J Phys Med Rehabil 1995; 74(2):139 44. [62] Kaigle AM, Wessberg P, Hansson TH. Muscular and kinematic behavior of the lumbar spine during flexionextension. J Spinal Disord 1998;11(2):163 74. [63] Gill KP, Callaghan MJ. The measurement of lumbar proprioception in individuals with and without low back pain. Spine 1998;23(3):371 7. [64] Brumagne S, Cordo P, Lysens R, Verschueren S, Swinnen S. The role of paraspinal muscle spindles in lumbosacral position sense in individuals with and without LBP. Spine 2000;25:989 94. [65] Hodges PW, Moseley GL. Pain and motor control of the lumbopelvic region: effect and possible mechanisms. Electromyogr Kinesiol 2003; in press. [66] Magill RA. Motor learning: concepts and applications. New York: McGraw-Hill; 2001. [67] Carr JH, Shepherd RB. A motor relearning programme for stroke. 2nd edition. London: Heinemann; 1987. [68] Fitts PM, Posner MI. Human performance. Belmont, CA: Brooks/Cole; 1967. [69] Voss DE, Ionta MK, Myers BJ. Proprioceptive neuromuscular facilitation: patterns and techniques. New York: Lippincott Williams & Wilkins Publishers; 1985. [70] Jull G, Trott P, Potter H, Zito G, Niere K, Shirley D, et al. A randomized controlled trial of exercise and manipulative therapy for cervicogenic headache. Spine 2002;27:1835 43. [71] Jull GA, Scott Q, Richardson C, Henry S, Hides J, Hodges P. New concepts for the control of pain in the lumbopelvic region. In: Vleeming A, Mooney V, Tilscher H, Dorman T, Snijders C, editors. Proceedings of the 3rd Interdisciplinary World Congress on Low Back and Pelvic Pain, Vienna, Austria. Rotterdam: European Conference Organisers; 1998. p. 128 31.

254

P.W. Hodges / Orthop Clin N Am 34 (2003) 245254 recovery is not automatic after resolution of acute, firstepisode low back pain. Spine 1996;21(23):2763 9. Hides JA, Jull GA, Richardson CA. Long term effects of specific stabilizing exercises for first episode low back pain. Spine 2001;26:243 8. Goldby L, Moore A, Doust J, Trew M. An RCT investigating the efficacy of manual therapy, exercises to rehabilitate spinal stabilization and an education booklet in the conservative treatment of chronic low back pain. In: Proceedings of International Federation of Manipulative Therapists. Perth, Australia: 2000. Stuge B. The efficacy of a specific stabilizing exercise program in the treatment of patients with peripartum pelvic pain after pregnancy. A randomized controlled trial. In: Proceedings of the 4th Interdisciplinary World Congress on Low Back and Pelvic Pain, Montreal, Canada. Rotterdam: European Conference Organisers; 2001. Daneels L, Cools A, Vanderstraeten G, Gambier D, Witrouw E, Bourgois J, et al. The effect of 3 different training modalities on the cross sectional area of paravertebral muscles. Scand J Med Sci Sports 2001;11(6): 335 41. Mannion AF, Taimela S, Muntener M, Dvorak J. Active therapy for chronic low back pain. Part 1. Effects on back muscle activation, fatigability, and strength. Spine 2001;26(8):897 908.

[72] Cowan SM, Bennel KL, Crossley KM, Hodges PW, McConnell J. Physical therapy alters recruitment of the vasti in patellofemoral pain syndrome. Med Sci Sports Exerc 2002;34:1879 85. [73] Schmidt RA, Lee TD. Motor control and learning: a behavioural emphasis. 3rd edition. Champaign, IL: Human Kinetics Publishers; 1999. [74] Hides JA, Richardson CA, Jull GA, Davies SE. Ultrasound imaging in rehabilitation. Aus J Physiother 1996;41:187 93. [75] Moseley GL, Hodges PW. Chronic pain and motor control. In: Boyling J, Jull GA, editors. Grieves modern manual therapy. Edinburgh: Churchill Livingstone; 2003. In press. [76] Creager CC. Therapeutic exercises using the Swiss ball. Berthoud (CO): Executive Physical Therapy; 1994. [77] Bullock-Saxton JE, Janda V, Bullock MI. Reflex activation of gluteal muscles in walking with balance shoes. Spine 1993;18(6):704 8. [78] Mannion AF, Muntener M, Taimela S, Dvorak J. A randomized clinical trial of three active therapies for chronic low back pain. Spine 1999;24(23):2435 48. [79] OSullivan PB, Twomey LT, Allison GT. Evaluation of specific stabilizing exercise in the treatment of chronic low back pain with radiologic diagnosis of spondylolysis or spondylolisthesis. Spine 1997;22(24):2959 67. [80] Hides JA, Richardson CA, Jull GA. Multifidus muscle

[81]

[82]

[83]

[84]

[85]

You might also like