Next Article in Journal
WBM-DLNets: Wrapper-Based Metaheuristic Deep Learning Networks Feature Optimization for Enhancing Brain Tumor Detection
Next Article in Special Issue
Lower Limb Biomechanics during the Golf Downswing in Individuals with and without a History of Knee Joint Injury
Previous Article in Journal
Comparison of Low and High Temperature Sintering for Processing of Bovine Bone as Block Grafts for Oral Use: A Biological and Mechanical In Vitro Study
Previous Article in Special Issue
The Role of Fascial Tissue Layer in Electric Signal Transmission from the Forearm Musculature to the Cutaneous Layer as a Possibility for Increased Signal Strength in Myoelectric Forearm Exoprosthesis Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Response to Mechanical Properties and Physiological Challenges of Fascia: Diagnosis and Rehabilitative Therapeutic Intervention for Myofascial System Disorders

1
Department of Orthopaedic Surgery, Okayama Rosai Hospital, 1-10-25 Midorimachi, Minamiku, Okayama 702-8055, Japan
2
Department of Central Rehabilitation, Okayama Rosai Hospital, 1-10-25 Midorimachi, Minamiku, Okayama 702-8055, Japan
*
Author to whom correspondence should be addressed.
Bioengineering 2023, 10(4), 474; https://doi.org/10.3390/bioengineering10040474
Submission received: 19 March 2023 / Revised: 6 April 2023 / Accepted: 12 April 2023 / Published: 14 April 2023

Abstract

:
Damage to the fascia can cause significant performance deficits in high-performance sports and recreational exercise and may contribute to the development of musculoskeletal disorders and persistent potential pain. The fascia is widely distributed from head to toe, encompassing muscles, bones, blood vessels, nerves, and internal organs and comprising various layers of different depths, indicating the complexity of its pathogenesis. It is a connective tissue composed of irregularly arranged collagen fibers, distinctly different from the regularly arranged collagen fibers found in tendons, ligaments, or periosteum, and mechanical changes in the fascia (stiffness or tension) can produce changes in its connective tissue that can cause pain. While these mechanical changes induce inflammation associated with mechanical loading, they are also affected by biochemical influences such as aging, sex hormones, and obesity. Therefore, this paper will review the current state of knowledge on the molecular level response to the mechanical properties of the fascia and its response to other physiological challenges, including mechanical changes, innervation, injury, and aging; imaging techniques available to study the fascial system; and therapeutic interventions targeting fascial tissue in sports medicine. This article aims to summarize contemporary views.

1. Introduction

Injuries to various fascial tissues can significantly reduce performance in sports [1] and may contribute to the development and persistence of symptoms of musculoskeletal disorders, including low back pain [2]. Although damage to the fascial tissue has been shown to be influenced primarily by exercise, aging, sex hormones, obesity, and inflammation [3,4,5,6,7,8,9,10,11,12,13,14], there is limited information about the elastic fiber composition of fascial tissue, extracellular matrix properties, vascularity, extent of innervation, and their role in disease and treatment [15]. Although these properties are fundamental for optimal sports performance, the fascia is not considered in the physical examination and rehabilitation of sports injuries, omitting anatomical tissues that could play an important role in sports injury rehabilitation. For example, acute pain associated with the groin is common in athletes injured by overuse. Acute strains have been shown to occur at the musculotendinous junction, particularly in the adductor longus, rectus femoris, and iliopsoas muscles [16]. X-rays and MRIs are performed to rule out the diagnosis of serious diseases, such as fractures [17,18,19,20], and if serious diseases are ruled out, pain provocation tests are performed using palpation, stretching, and resistance testing [21,22]. However, these evaluation methods focus on the joint range of motion and muscle strength and do not consider the assessment of disorders of the fascia [23,24].
An accurate understanding of the mechanical behavior of the fascia plays an important role in investigating pathological phenomena and comprehensive analysis of its functionality. Furthermore, the fascia and muscle in sports injuries can be identified using conventional imaging modalities such as magnetic resonance imaging (MRI) and ultrasound imaging, which are valuable guides for appropriately rehabilitating sports injuries. Furthermore, physical examination of these tissues at different sites, depending on the pathology, may influence the rehabilitation outcome. This article will review the current state of knowledge on the molecular to mechanical properties of fascia and its response to other physiological challenges, including mechanical changes, innervation, injury, and aging; ultrasound imaging and MRI available to study the fascial system; and therapeutic interventions targeting fascial tissue in sports medicine. The purpose of this article is to summarize contemporary views on the topics highlighted in the following sections.

2. Role of Fascial Tissue and Pathological Reactions

Fascia is widely distributed from head to toe; it encases and permeates muscles, bones, blood vessels, nerves, and internal organs, constitutes various layers of different depths [25], and is a connective tissue composed of irregularly arranged collagen fibers, clearly different from the regularly arranged collagen fibers found in tendons, ligaments, or periosteal sheets [26]. Additionally, it supports important functions of the human body, such as posture, movement, and homeostasis [25,26,27,28], and also contains various sensory receptors for proprioception, nociception, and even hormones [28].
Structurally, fascial tissues are composed of various cell types (fibroblasts, myofibroblasts, myofascial cells, and telocytes), as well as fibrous (type I and type III collagen fibers, elastin, fibrillin), aqueous (a complex mixture of water and glycosaminoglycans) components, and neural elements (free nerve endings and mechanoreceptors) [29,30]. The fact that the fascia can transmit tension far is the basis of the “biotensegrity” framework [31,32]. Biotensegrity is the application of the principle of tensegrity to the understanding of human movement, where tensegrity is an architectural principle according to which a structure (or tensegrity system) is stabilized by continuous tension with discontinuous compression and functions as a single structure [32]. As the tension in the fascia increases, the connective tissue can disperse the force around it and propagate it along the fascial system [31,32,33,34]. Forces passively imposed on the muscle by stretching are distributed throughout the tissue via the intramuscular connective tissue [33,34]. Fascia transmits tension, influences other muscles, plays a role in the proper coordination of body movements, and can reflect the direction of force vectors. Fascia can actively contract, and changes in tension are caused by contractile cells [35]. Myofibroblasts are present in developing and normal adult tissues and are responsible for altering tissue tension [35]. Normal fibroblasts are highly sensitive to physical stimuli.
The transition from fibroblast to myofibroblast is influenced by mechanical stress. Upon mechanical tension, fibroblasts differentiate into proto-myofibroblasts, which contain actin stress fibers in their cytoplasm that terminate in a fiber bundle adhesion complex [25,36,37]. The adhesion complex bridges the internal cytoskeleton and integrins of myofibroblasts with extracellular matrix (ECM) fibronectin fibers. Thus, this allowed contractile forces to be generated in the nearby ECM when traction is applied; moreover, forces within the ECM are maintained over time and are further enhanced by remodeling and collagen deposition [37]. In addition, chronic strain, such as sitting or overuse of muscles [38,39,40], infection and inflammation [40], and immobilization of the limb by trauma, fracture, or casting [28,29,30], can produce further contraction of myofibroblast smooth muscle actin fibers and contribute to joint contractures. These environments make it difficult to maintain a relaxed state, resulting in decreased mechanical tension, and consequently, myofibroblasts either dedifferentiate or undergo apoptosis [37]. The tipping point between exercise and rest is unknown; however, multiple repetitions of the contraction cycle may result in graded and irreversible tissue contraction [37].

3. Factors Influencing the Pathological Development of Fascia Tissue (Nerve, Disorder, Aging, Sex Hormone)

Since blood vessels and nerves are scattered throughout the fascia, invasion of these structures via fascial changes is common [37,38]. Tissue changes translate into changes in the mobility of the nerve, resulting in a decrease in the independence of the nerve from its surroundings [39,40]. Changes in the fibrous tissue around the nerve can cause entrapment lesions [39,40], and patients may notice numbness, dysesthesia, and pain [41]. Existing histological studies indicate that the fascia is the largest sensory organ because of its large surface area [42,43]. However, it has been shown that the type and density of innervation, including receptors, varies depending on the fascia present in the different parts of the body, indicating that this tissue is more complex than imagined [44,45,46,47,48,49,50,51]. The role of the fascial tissue as a sensory receptor and the environmental changes associated with structural changes in fascial tissues may increase the stimulation of free nerve endings in the fascia, potentially increasing inflammation and pain. Reportedly, free nerve endings have nociceptor characteristics, and the density and length of nociceptors increase in inflamed thoracolumbar fascia [49,50]. Moreover, repetitive identical postures, sports, and repetitive motion may produce movement patterns that increase tissue thickness and limit slippage between fascial layers. In addition, structural changes in the fascia that occur after fascial adhesions due to trauma, overuse, or surgery can alter the activation of nerve receptors embedded in the fascia [52,53,54]. However, few effective treatments are available due to the complexity of the mechanism [55,56,57].
Mechanical stress induces the release and activation of molecules stored in the ECM, triggering the cleavage products of collagen XVIII and other basement membrane components; because the ECM is the primary site of the inflammatory response that occurs in tissues. Moreover, the ECM interacts with immune cells, especially under dynamic conditions such as growth and regeneration, and the fascial tissue requires significant changes in the local ECM microenvironment to allow cellular adaptation and remodeling of the ECM [58,59,60].
Following acute injury due to overload or anoxia in fascial tissues, the immune response aims to phagocytose injured cells. The acute inflammatory response is usually brief and reversible. It involves the release of inflammatory cytokines from injured cells and macrophages, other substances (such as bradykinin, substance P, and proteases), and molecules that sensitize nociceptive nerves [61] and promotes immune cell infiltration. However, prolonged or repeated loading results in persistent inflammation [62,63,64], and the long-term presence of macrophages and cytotoxic cytokines in and around the tissue ultimately leads to progressive tissue damage. Cytotoxic cytokines (e.g., interleukin-1β, tumor necrosis factor (TNF), transforming growth factor-β (TGFβ-1)) promote fibrosis through fibroblast overgrowth and collagen matrix deposition [65]. Notably, cytokine overproduction can also lead to nociceptive afferent nerve maintenance sensitization and increase the production and release of substance P (nociceptive neuropeptide). Prolonged or repeated loading results in persistent inflammation and the prolonged presence of macrophages and cytotoxic cytokines in and around the tissue [63,64]. Eventually, tissue damage progresses, and overproduction of cytokines is triggered. This overproduction of cytokines maintains sensitization of nociceptive afferents and increases the production and release of substance P [65]. Substance P stimulates TGFβ-1 production by fibroblasts. Furthermore, the substances P and TGFβ-1 have been shown to induce fibrogenic processes independently [65]. Therefore, it is suggested that neurogenic processes (substance P) and loading/repair processes (TGFβ-1) may contribute to increased collagen in fascial tissue. Fibrosis around fascial tissue affects secondary dynamic biomechanical properties, anchoring structures to one another or inducing chronic compression [60]. In addition, inflammatory cytokines leak into the bloodstream, causing extensive secondary tissue damage and impaired function of central nociceptors [62,66,67]. Therefore, it is suggested that the pathogenesis of myofascial tissue injury is related to the maintenance of musculoskeletal function during daily life and exercise in the elderly and the prevention of overuse injuries in athletes. The following methods have been reported to decrease inflammatory cytokines. Early treatment with anti-inflammatory agents can prevent or reduce pain induced by TNF signaling and decrease downstream collagen production [68]. Stretching of the fascial tissue can promote the resolution of inflammation both in vivo and in vitro, and manual therapy can prevent overuse-induced fibrosis in some fascial tissues [69,70]. Nevertheless, there is limited information on exercise-induced changes in myofascial tissue and its molecular response to aging. Hence, further research is needed.
Physiological aging is a highly individualized process characterized by the progressive degeneration of tissues and organ systems. For example, a sedentary lifestyle and repetitive overuse of muscles with a limited range of motion can lead to myofascial pain syndrome (MPS), producing pain in multiple areas of the musculoskeletal system resulting from myofascial changes/fibrosis. In fact, physically active workers are less likely to develop symptoms of MPS compared to sedentary workers [67]. Functionally, these pathological changes alter the mechanical properties of myofascial tissue and skeletal muscles, causing pain and age-related decreases in muscle strength and range of motion that cannot be explained by muscle mass loss alone [5]. Physical inactivity is more likely to cause pain due to collagen changes in fascia, decreased fascial sliding due to hyaluronic acid (HA) aggregation, increased contractility of myofibroblasts, and increased production of inflammatory cytokines [46,63,64,65]. Furthermore, aging is associated with fluctuations in fascia thickness. Indeed, age-related modifications are specific to different body regions. Fascia thickness in the lower extremities decreases with age (−12.3–25.8%), while that in the lower back increases (+40.0–76.7%) [71]. These connective tissue changes have been suggested to decrease joint flexibility [72]. Moreover, structural and molecular changes in fascia with aging affect force transmission in the locomotor system [73]. Fascial tissue becomes denser, and fibrosis develops with age, reducing muscle force transmission and joint range of motion [74,75]; consequently, body pain due to MPS is often attributed to age-related loss of mobility and can be viewed as a natural consequence of a sedentary lifestyle.
Today, the percentage of women participating in physical activity and sports has never been higher [76,77]. There is also a growing awareness of the potential effects of cyclical menstrual hormones (estrogen and progesterone) on exercise performance [78] and metabolic demand [79]. Estrogen is a major regulator of the female body composition; however, it is also involved in muscle damage and recovery [80,81]. Reportedly, markers of muscle damage (creatine kinase) and inflammation (interleukin 6) are significantly greater in the follicular phase than in the mid-luteal phase [82]. Thus, the menstrual cycle phase may be involved in inflammation and recovery. Therefore, contraceptive methods that can alter hormonal fluctuations, such as oral contraceptives, are sometimes used by female athletes [83,84,85]. Oral contraceptives generally contain estrogen and progestin [86] and should be cautiously used. Reportedly, females taking oral contraceptives have a lower anterior cruciate ligament (ACL) elasticity than those not taking oral contraceptives [87], and muscle-tendon stiffness in the lower extremities of young females is lower during the ovulatory phase [88,89]. However, oral contraceptives may adversely affect inflammatory markers, as evidenced by the fact that Olympic-level elite female athletes using oral contraceptives had higher levels of C-reactive protein, a marker of inflammation, than amenorrheic athletes, suggesting increased muscle damage and poor recovery potential [90,91]. In fascial tissues, acute and chronic loading stimulates collagen remodeling [6]. Moreover, the increase in collagen synthesis with exercise is lower in females than in males, and sex differences in injury frequency and estrogen receptor expression in human fascial tissue suggest that estrogen may play an important regulatory role in ECM remodeling [6,7,8]. Estrogen replacement in older postmenopausal females inhibits collagen synthesis during exercise; in contrast, it has a stimulatory effect on collagen synthesis at rest [9]. Furthermore, estrogen and estrogen receptor beta (Erβ) inhibit fibrosis by decreasing TGFβ expression, connective tissue growth factor production and function, matrix metalloproteinases 2 and 9 expressions and activity, fibroblast conversion to myofibroblasts, and collagen I and III production [89,90]. Thus, long-term estrogen deficiency is known to be associated with increased fibrosis. The presence of sex hormone receptors in fascial tissues may help explain sex differences in the prevalence of myofascial pain [88].
Changes in the physical and chemical properties of hyaluronic acid (HA) are associated with changes in the viscoelasticity, mechanical plasticity, and nonlinear elasticity of the extracellular matrix, as previously described [92,93], which may contribute to fascial disease. HA occurs between the deep fascia and muscle, facilitating sliding between these two structures and within the loose connective tissue of the fascia, ensuring smooth sliding of adjacent fibrous fascial layers [94]. Although exercise promotes HA production and recycling, the biomechanical properties of free connective tissue may change in response to the amount of lactic acid accumulated after intense exercise. Furthermore, a decrease in pH due to lactate accumulation increases the viscosity of HA, resulting in instant stiffness [95]. In contrast, immobility increases the concentration of HA without effective HA recycling, which may increase viscosity and decrease lubrication and sliding of connective tissue and muscle layers [96]. In addition, the thickening of the fascia caused by aging [66] increases the distance between surfaces and leads to an increase in HA viscosity [97]. The increase in HA viscosity within connective tissue may inhibit the sliding of fascial collagen fibers between layers [98]. The patient may perceive this increase in overall fascial thickness as increased stiffness and pain. An important component of pain treatment is reversing these changes in HA; when HA becomes sticky rather than lubricated, densification of the fascia occurs, and the distribution of force lines within the fascia is distorted [67]. Additionally, small repetitive movements, immobility, or overuse syndromes of movement that result in negative modifications of loose connective tissue can distort loose connective tissue between fascial layers and densification. From the properties of HA within the ECM, this change, reportedly, is reversible by modification of temperature, pH, and mechanical loading (such as massage) [74,98].

4. Imaging Diagnosis

Ultrasound and MRI are the commonly used imaging modalities for fascial injuries. Ultrasonography of fascia allows the deep fascia to be observed and provides a more convenient assessment of subcutaneous and perimuscular connective tissue thickness compared with other imaging modalities. Moreover, ultrasonography can also provide dynamic images. This allows us to identify slippage and movement between the muscle and the adjacent fascial layer and on the fascial side for the movement we wish to evaluate. Another feature is that the thickness of the fascia varies with age [99,100,101]. The thickness of the fascia and its relationship with the underlying muscles distinguish aponeurotic fasciae from epimysial fasciae, which are covered by a fibrous sheath and maintain their position as a muscle group or broad muscle attachment; in contrast, epimysial fasciae are covered by a fibrous sheath and maintain their position as a muscle group or broad muscle attachment. The epimysial fasciae characterize each muscle and determine its shape and volume. By observing each, the relationship between muscle, fascia, bone, and surrounding soft tissues can be identified [102,103]. Ultrasonography is also useful for muscle atrophy and fascial tears. Muscle atrophy can be observed by measuring the volume under ultrasound. In contrast, fascial tears can be identified by observing the damage to the continuity of the fascia. Observing the fascia while moving it and assessing the proximal and distal areas of the tear can also determine if the fascia is completely or incompletely torn [104,105]. MRI can assess fascial thickening and signal changes, adjacent soft tissue, and bone marrow edema; T1, T2; moreover, fat suppression images on MRI can identify various injuries [106]. Furthermore, MRI can identify fascial thickening, perifascial fluid retention, contiguous tears, adjacent soft tissue edema, and bone marrow edema at fascial bone marrow edema in the adherent area. Fascial thickening and complete tears are easily identified on MRI. For example, fibrous tissues may demonstrate a low signal in the T1- and T2-weighted images and a high signal in fat suppression in some cases. Acute myofascial tears show low signal areas of tear, high signal on fat suppression, and intermediate signal changes on T1. High signal areas may indicate fluid retention in the soft tissues surrounding the injury. The T1- and T2-weighted images showing heterogeneous signal intensity may indicate a neoplastic lesion. In the case of a foreign body reaction, the incidence of a low signal on T1-weighted images and a high signal on T2-weighted images is high. In addition, an infection may be present with a high signal on fat suppression and a low signal on the T1-weighted images, indicating contrasting enhancements. Thus, the degree of inflammation, localization, and spread of infection can be evaluated [106,107]. In addition, ultrasonography and MRI are easy to identify the fascia, and the contrast between the fascia and its surroundings is clear; therefore, there is a high degree of inter-specialty reliability in evaluating the fascia.

5. Myofascial Release (MFR) for Muscle and Fascia Dysfunction

Myofascia degenerates due to various causes. The main causes are circulatory failures due to trauma or reduced physical activity, disuse syndrome, overuse syndrome due to repetitive motion, and chronic poor posture. This causes the densification of the fascia because of the twisted collagen fibers, which harden the substrate because of dehydration. In addition, sustained muscle contraction, such as in overuse syndrome, causes hyaluronic acid aggregation, which is a factor that reduces the sliding properties of the fascia [58,69,70]. Furthermore, it has also been hypothesized that because of the continuity of fascia, dysfunction of fascia in one part of the body can cause stress in other body parts [108]. Thus, gliding between the fascia and its deeper tissues, such as muscle tissue, is inhibited, reducing the ability to maintain an antigravity posture and efficient athletic performance. Therefore, it causes poor performance in sports and interference with daily activities. Training is necessary to improve performance, especially in athletes; however, inadequate rest periods can cause high-frequency, high-intensity training that leads to continued pointless training [109]. Maladaptive training before tissue recovery and rebuilding can lead to the accumulation of microdamage in affected tissues, resulting in overuse injuries, thereby compromising the athletes’ competitive performance due to pain and dysfunction [110]. As evidenced by the fact that 39% of athletes experience unexplained musculoskeletal pain weekly [111], athletes may continue to train despite the risk of disability. Since it is difficult to determine training at the appropriate load and rest periods, we believe that MFR, as discussed below, can reduce the risk of overuse injury. MFR aims to improve the mobility and length of myofascial tissue, thereby reducing pain and allowing it to function normally [112]. As a treatment, the injection of local anesthetics into the interfascial space was expected to be a new anesthetic technique [113]. Nevertheless, it has since been clinically applied and used as a hydrorelease. Reportedly, the hydrorelease has shown therapeutic efficacy, including improvement in pain and a range of motion after arthroscopic surgery [113]. Furthermore, there was an improvement in low back pain after 5 min of intervention on the multifidus muscle in patients with acute low back pain [114]. In addition, the therapist and patient can see where the fascia is being stripped using ultrasound imaging, possibly because the patient may experience changes in muscle tension and pain, thus, increasing the level of expectations and satisfaction with the treatment. MFR in physical therapy (manual therapy) uses techniques systematized by Barnes and colleagues [115,116]. This is a technique that not only stretches the fascia but also unties and untwists it. Recently, self-myofascial release (SMFR) using foam rolling (FR) has become popular. The effects of FR include reducing delayed-onset muscle soreness after exercise [117], increasing the pain threshold, and making the pain less perceptible [118]. It has also been reported to improve the range of motion without decreasing exercise performance [119].
MFR has been applied to treat a wide range of disease areas, including tension-type headaches [120], postmenopausal venous insufficiency [121], and nonspecific chronic low back pain [122]. This is believed to be because MFR does not require joint movement and applies mild myofascial stretching and pressure, making it safe and applicable to many age groups and diseases.
Furthermore, Ichikawa et al. compared fascial gliding and flexibility of the vastus lateralis muscle with MFR and hot pack treatment and reported improved gliding and flexibility in MFR [123]. It was suggested that continuous lengthening and pressure were necessary to improve these. Furthermore, several reports have shown the effectiveness of MFR in combination with conventional therapies rather than as a standalone treatment [124,125]. For example, Ozsoy et al. reported that in elderly patients with nonspecific low back pain, the group that received MFR and trunk stabilization exercises showed a significant increase in MFR and trunk compared to the group that received only MFR. In addition, the results reported that trunk endurance and mobility improved in the group that received MFR and trunk stabilization exercises [126], suggesting that MFR in combination with conventional exercise therapy is effective.
Recently, scattered reports have shown the effectiveness of MFR in combination with conventional treatment for postoperative orthopedic patients [127,128]. Therefore, the Oswestry disability index (ODI) was examined at the beginning of the study, and a 1-month follow-up after the intervention was discontinued. The exercise was performed three times a week under the supervision of a physical therapist. The results showed that ODI improved significantly more with MFR than with SE alone, with a minimally clinically important difference of more than 10% [129,130]. The mechanism of these effects is suggested to be that MFR causes normalization of the apoptosis rate of damaged fascia, changes in cell morphology, and reorientation of fibroblasts [131], reducing fascial shortening and thickening, improving normal muscle length and flexibility of fascial tissue [102]. Furthermore, by activating the descending pain suppression system [132], it is speculated that pain modulation occurred and contributed to the improvement of ODI.
However, studies using MFR in the postoperative orthopedic setting are still limited, making generalized conclusions difficult to establish. Postoperative patients have been reported to experience tissue fibrosis during wound healing [133]. Kawanishi et al. reported improved pain [134] and walking ability [135] by improving subcutaneous tissue gliding in patients with femoral metaphyseal fractures. The effectiveness of a broad range of orthopedic postoperative therapies that focus on the fascia and the connective tissue as a whole needs to be clarified.
SMFR is used to prevent injuries and maintain performance in sports [136,137]. For example, in soccer, in the early period after half-time, both physical and cognitive performance is reduced, and the risk of injury increases [138]. Kaya et al. tested the effects of SMFR by replicating the running distance and half-time experienced during a game with soccer players of various levels [139]. When SMFR was performed during the recreated half-time period, the subsequent decrease in sprint performance was suppressed [139]. This effect, reported by Okamoto et al., was attributed to SMFR improving vascular endothelial function, increasing blood flow to the muscles targeted by SMFR, increasing the supply of oxygen and other vital nutrients, and promoting more efficient removal of metabolites [140]. Athletes, young and old, male and female, novice and elite, are prone to delayed-onset muscle soreness (DOMS) after intense regular exercise; DOMS results in fibrous tissue adhesions that limit joint range of motion [115,141,142]. Adhesions of fibrous tissue and fascia occur due to disease or injury and reduce joint ROM, muscle length, muscle endurance, and motor coordination [115,143,144]. Therefore, suppressing DOMS may also inhibit fascial adhesions. Static stretching (SS) is generally used before exercise to improve the range of motion and prevent injury. It is also believed to decrease force and power, making it difficult to use before exercise [145]. However, in a study in which SMFR was performed prior to exercise, it improved exercise performance without decreasing force and power, in addition to improving the range of motion [142,146,147]. SMFR has also been reported to reduce muscle fatigue before exercise [147,148]. Thus, the use of SMFR before exercise is thought to be effective in restoring range of motion, fatigue, and performance after exercise [146,147]. A systematic review investigating the effects of post-exercise massage reported that massage, such as MFR, stimulates the parasympathetic nervous system, indirectly enhances the immune system by improving local circulation, and decreases inflammatory cytokines [149]. SMFR, especially after exercise, has been shown to be beneficial for recovery after exercise-induced muscle damage (EIMD), DOMS, and other impairments of physical performance [150,151,152,153,154]. The improved performance after SMFR has also been reported to last up to 72 h [152,154]. SMFR has been reported to be a safe intervention used for performance (especially flexibility) and recovery from previous training and competition and can reduce DOMS [154,155]. These findings suggest that SMFR, as a routine practice before and after exercise, may help prevent fascial adhesions in athletes and reduce the incidence of injury.
When the fascia is restricted in any part, it causes stress and impairment in other areas [156], depending on the continuity of the myofascial structure, and reduces muscle flexibility [138]. This is a particularly important issue for athletes subjected to long-term repetitive strain, and using SMFR with FR is an important part of an athlete’s training. Therefore, SMFR with FR is recommended for inclusion in athletes’ training regarding performance loss, injury prevention, and recovery [138].
While some reports have shown the effectiveness of MFR, as mentioned earlier, there are several manual therapies for myofascial as follows: osteopathic soft tissue techniques [157], strain counter strain [158], myofascial trigger point therapy [159], muscle energy technique [160]. However, there is no evidence to suggest which manual therapy is optimal [161]. Furthermore, while MFR for chronic low back pain patients is effective for low back pain and ADL disability, the clinical significance of MFR is unclear [162]. Therefore, further studies are needed to clarify clinical relevance and build evidence in the future.

6. Conclusions

In myofascial tissue, normal fibroblasts are highly sensitive to physical stimuli. The transition from fibroblast to myofibroblast is influenced by mechanical stress and may result in gradual and irreversible tissue contraction as the contraction cycle is repeated many times. The acute inflammatory response of fascial tissue is usually brief and reversible but repeated identical postures, sports, and repetitive motion promote fibrosis via fibroblast overgrowth and collagen matrix deposition. Fibrosis around fascial tissue is translated into nerve mobility changes, resulting in dysfunction of central nociceptors. In addition, the thickening of the fascia increases the distance between surfaces, increasing HA viscosity. The increase in HA viscosity within the connective tissue inhibits the sliding of fascial collagen fibers between layers. Furthermore, MRI imaging allows the deep fascia to be observed and provides a broad and detailed assessment of fascial thickening and signal changes, as well as adjacent soft tissue and bone marrow edema; in contrast, ultrasound imaging allows dynamic observation of the fascia and is a useful tool for assessing the proximity and distal extent of the injury to determine treatment efficacy. The goal of myofascial release for myofascial dysfunction is to reduce pain and allow myofascial tissue mobility to function normally. Several manual therapies have been developed in addition to myofascial release; however, there is no evidence to suggest which manual therapy is best, and the effectiveness of a broad range of therapies targeting the entire connective tissue system needs to be clarified. Additionally, the effectiveness of a broad range of therapies that look at connective tissue as a whole also needs to be clarified.

Author Contributions

Validation, Y.K. and S.M.; resources, T.O. and A.K.; writing—original draft preparation, Y.K., S.M. and T.S.; writing—review and editing, M.T. and M.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Ethical review and approval were waived for this study due to review study.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are openly available in Bioengineering at Bioengineering|Special Issue: Physical Examination and Rehabilitation of Fasciae and Muscles in Sports Injuries (mdpi.com).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ljungqvist, A.; Schwellnus, M.P.; Bachl, N.; Collins, M.; Cook, J.; Khan, K.M.; Maffulli, N.; Pitsiladis, Y.; Riley, G.; Golspink, G.; et al. International Olympic Committee consensus statement: Molecular basis of connective tissue and muscle injuries in sport. Clin. Sports Med. 2008, 27, 231–239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Wilke, J.; Schleip, R.; Klingler, W.; Stecco, C. The lumbodorsal fascia as a potential source of low back pain: A narrative review. BioMed Res. Int. 2017, 2017, 5349620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Shireman, P.K.; Contreras-Shannon, V.; Ochoa, O.; Karia, B.P.; Michalek, J.E.; McManus, L.M. MCP-1 deficiency causes altered inflammation with impaired skeletal muscle regeneration. J. Leukoc. Biol. 2007, 81, 775–785. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, H.; Melton, D.W.; Porter, L.; Sarwar, Z.U.; McManus, L.M.; Shireman, P.K. Altered macrophage phenotype transition impairs skeletal muscle regeneration. Am. J. Pathol. 2014, 184, 1167–1184. [Google Scholar] [CrossRef] [Green Version]
  5. Zhang, C.; Gao, Y. Effects of aging on the lateral transmission of force in rat skeletal muscle. J. Biomech. 2014, 47, 944–948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Miller, B.F.; Hansen, M.; Olesen, J.L.; Schwarz, P.; Babraj, J.A.; Smith, K.; Rennie, M.J.; Kjaer, M. Tendon collagen synthesis at rest and after exercise in women. J. Appl. Physiol. 2007, 102, 541–546. [Google Scholar] [CrossRef]
  7. Magnusson, S.P.; Hansen, M.; Langberg, H.; Miller, B.; Haraldsson, B.; Westh, E.K.; Koskinen, S.; Aagaard, P.; Kjaer, M. The adaptability of tendon to loading differs in men and women. Int. J. Exp. Pathol. 2007, 88, 237–240. [Google Scholar] [CrossRef]
  8. Fede, C.; Albertin, G.; Petrelli, L.; Sfriso, M.M.; Biz, C.; De Caro, R.; Stecco, C. Hormone receptor expression in human fascial tissue. Eur. J. Histochem. 2016, 60, 2710. [Google Scholar] [CrossRef] [Green Version]
  9. Hansen, M.; Kongsgaard, M.; Holm, L.; Skovgaard, D.; Magnusson, S.P.; Qvortrup, K.; Larsen, J.O.; Aagaard, P.; Dahl, M.; Serup, A.; et al. Effect of estrogen on tendon collagen synthesis, tendon structural characteristics, and biomechanical properties in postmenopausal women. J. Appl. Physiol. 2009, 106, 1385–1393. [Google Scholar] [CrossRef] [Green Version]
  10. Ugwoke, C.K.; Cvetko, E.; Umek, N. Pathophysiological and therapeutic roles of fascial hyaluronan in obesity-related myofascial disease. Int. J. Mol. Sci. 2022, 23, 11843. [Google Scholar] [CrossRef]
  11. Mackey, A.L.; Kjaer, M.; Dandanell, S.; Mikkelsen, K.H.; Holm, L.; Døssing, S.; Kadi, F.; Koskinen, S.O.; Jensen, C.H.; Schrøder, H.D.; et al. The influence of anti-inflammatory medication on exercise-induced myogenic precursor cell responses in humans. J. Appl. Physiol. 2007, 103, 425–431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Christensen, B.; Dandanell, S.; Kjaer, M.; Langberg, H. Effect of anti-inflammatory medication on the running-induced rise in patella tendon collagen synthesis in humans. J. Appl. Physiol. 2011, 110, 137–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Calve, S.; Simon, H.G. Biochemical and mechanical environment cooperatively regulate skeletal muscle regeneration. FASEB J. 2012, 26, 2538–2545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Pagel, C.N.; Wijesinghe, D.K.W.; Taghavi Esfandouni, N.; Mackie, E.J. Osteopontin, inflammation and myogenesis: Influencing regeneration, fibrosis and size of skeletal muscle. J. Cell Commun. Signal. 2014, 8, 95–103. [Google Scholar] [CrossRef] [Green Version]
  15. Stecco, C.; Corradin, M.; Macchi, V.; Morra, A.; Porzionato, A.; Biz, C.; de Caro, R. Plantar fascia anatomy and its relationship with Achilles tendon and paratenon. J. Anat. 2013, 223, 665–676. [Google Scholar] [CrossRef]
  16. Serner, A.; Weir, A.; Tol, J.L.; Thorborg, K.; Roemer, F.; Guermazi, A.; Yamashiro, E.; Hölmich, P. Characteristics of acute groin injuries in the adductor muscles: A detailed MRI study in athletes. Scand. J. Med. Sci. Sports 2018, 28, 667–676. [Google Scholar] [CrossRef]
  17. Davies, A.G.; Clarke, A.W.; Gilmore, J.; Wotherspoon, M.; Connell, D.A. Imaging of groin pain in the athlete. Skeletal. Radiol. 2010, 39, 629–644. [Google Scholar] [CrossRef]
  18. Knapik, J.J.; Reynolds, K.L.; Hoedebecke, K.L. Stress fractures: Etiology, epidemiology, diagnosis, treatment, and prevention. J. Spec. Oper. Med. 2017, 17, 120–130. [Google Scholar] [CrossRef]
  19. Armfield, D.R.; Towers, J.D.; Robertson, D.D. Radiographic and MR imaging of the athletic hip. Clin. Sports Med. 2006, 25, 211–239. [Google Scholar] [CrossRef]
  20. Georgiadis, A.G.; Zaltz, I. Slipped capital femoral epiphysis: How to evaluate with a review and update of treatment. Pediatr. Clin. N. Am. 2014, 61, 1119–1135. [Google Scholar] [CrossRef]
  21. Serner, A.; Tol, J.L.; Jomaah, N.; Weir, A.; Whiteley, R.; Thorborg, K.; Robinson, M.; Hölmich, P. Diagnosis of acute groin injuries: A prospective study of 110 athletes. Am. J. Sports Med. 2015, 43, 1857–1864. [Google Scholar] [CrossRef] [PubMed]
  22. Serner, A.; Weir, A.; Tol, J.L.; Thorborg, K.; Roemer, F.; Guermazi, A.; Hölmich, P. Can standardised clinical examination of athletes with acute groin injuries predict the presence and location of MRI findings? Br. J. Sports Med. 2016, 50, 1541–1547. [Google Scholar] [CrossRef] [PubMed]
  23. Thorborg, K.; Branci, S.; Nielsen, M.P.; Tang, L.; Nielsen, M.B.; Hölmich, P. Eccentric and isometric hip adduction strength in male soccer players with and without adductor-related groin pain: An assessor-blinded comparison. Orthop. J. Sports Med. 2014, 2, 2325967114521778. [Google Scholar] [CrossRef]
  24. Reiman, M.P.; Thorborg, K. Clinical examination and physical assessment of hip joint-related pain in athletes. Int. J. Sports Phys. Ther. 2014, 9, 737–755. [Google Scholar]
  25. Willard, F.H.; Vleeming, A.; Schuenke, M.D.; Danneels, L.; Schleip, R. The thoracolumbar fascia: Anatomy, function and clinical considerations. J. Anat. 2012, 221, 507–536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Langevin, H.M.; Keely, P.; Mao, J.; Hodge, L.M.; Schleip, R.; Deng, G.; Hinz, B.; Swartz, M.A.; De Valois, B.A.; Zick, S.; et al. Connecting (T)issues: How Research in Fascia Biology Can Impact Integrative Oncology. Cancer Res. 2016, 76, 6159–6162. [Google Scholar] [CrossRef] [Green Version]
  27. Barker, P.J.; Briggs, C.A. Attachments of the posterior layer of lumbar fascia. Spine 1999, 24, 1757–1764. [Google Scholar] [CrossRef]
  28. Nordez, A.; Gross, R.; Andrade, R.; Le Sant, G.; Freitas, S.; Ellis, R.; McNair, P.J.; Hug, F. Non-Muscular Structures Can Limit the Maximal Joint Range of Motion during Stretching. Sports Med. 2017, 47, 1925–1929. [Google Scholar] [CrossRef]
  29. Fede, C.; Pirri, C.; Fan, C.; Petrelli, L.; Guidolin, D.; de Caro, R.; Stecco, C. A closer look at the cellular and molecular components of the deep/muscular fasciae. Int. J. Mol. Sci. 2021, 22, 1411. [Google Scholar] [CrossRef]
  30. Pirri, C.; Fede, C.; Pirri, N.; Petrelli, L.; Fan, C.; de Caro, R.; Stecco, C. Diabetic foot: The role of fasciae, a narrative review. Biology 2021, 10, 759. [Google Scholar] [CrossRef]
  31. Sharkey, J. Fascia and living tensegrity considerations in: Lower extremity and pelvic entrapment neuropathies. Int. J. Anat. Res. 2021, 9, 7881–7885. [Google Scholar] [CrossRef]
  32. Sharkey, J. Fascia and Tensegrity the Quintessence of a Unified Systems Conception. Int. J. Anat. Appl. Physiol. 2021, 7, 174–178. [Google Scholar]
  33. Bordoni, B.; Myers, T. A Review of the Theoretical Fascial Models: Biotensegrity, Fascintegrity, and Myofascial Chains. Cureus 2020, 12, e7092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Dischiavi, S.L.; Wright, A.A.; Hegedus, E.J.; Bleakley, C.M. Biotensegrity and myofascial chains: A global approach to an integrated kinetic chain. Med. Hypotheses 2018, 110, 90–96. [Google Scholar] [CrossRef]
  35. Tomasek, J.J.; Gabbiani, G.; Hinz, B.; Chaponnier, C.; Brown, R.A. Myofibroblasts and mechano-regulation of connective tissue remodelling. Nat. Rev. Mol. Cell Biol. 2002, 3, 349–363. [Google Scholar] [CrossRef] [PubMed]
  36. Stempien-Otero, A.; Kim, D.H.; Davis, J. Molecular networks underlying myofibroblast fate and fibrosis. J. Mol. Cell Cardiol. 2016, 97, 153–161. [Google Scholar] [CrossRef] [Green Version]
  37. Castella, L.F.; Buscemi, L.; Godbout, C.; Meister, J.J.; Hinz, B. A new lock-step mechanism of matrix remodelling based on subcellular contractile events. J. Cell Sci. 2010, 123, 1751–1760. [Google Scholar] [CrossRef] [Green Version]
  38. Akbar, M.; McLean, M.; Garcia-Melchor, E.; Crowe, L.A.; McMillan, P.; Fazzi, U.G.; Martin, D.; Arthur, A.; Reilly, J.H.; McInnes, I.B.; et al. Fibroblast activation and inflammation in frozen shoulder. PLoS ONE 2019, 14, e0215301. [Google Scholar] [CrossRef] [Green Version]
  39. Kagawa, E.; Nimura, A.; Nasu, H.; Kato, R.; Akita, K. Fibrous connection between cervical nerve and zygapophysial joint and implication of the cervical spondylotic radiculopathy: An anatomic cadaveric study. Spine 2021, 46, E704–E709. [Google Scholar] [CrossRef]
  40. Wertsch, J.J.; Melvin, J. Median nerve anatomy and entrapment syndromes: A review. Arch. Phys. Med. Rehabil. 1982, 63, 623–627. [Google Scholar]
  41. Simons, D.G.; Janet, G.T.; Lois, S.S. Travell & Simons’ Myofascial Pain and Dysfunction: The Trigger Point Manual, 2nd ed.; Lippincott Williams & Wilkins: Baltimore, MD, USA, 1999; pp. 940–955. [Google Scholar]
  42. Sakada, S. Mechanoreceptors in fascia, periosteum and periodontal ligament. Bull. Tokyo Med. Dent. Univ. 1974, 21, 11–13. [Google Scholar] [PubMed]
  43. Stilwell, D.L. Regional variations in the innervation of deep fasciae and aponeuroses. Anat. Rec. 1957, 127, 635–653. [Google Scholar] [CrossRef] [PubMed]
  44. Corey, S.M.; Vizzard, M.A.; Badger, G.J.; Langevin, H.M. Sensory innervation of the nonspecialized connective tissues in the low back of the rat. Cells Tissues Organs. 2011, 194, 521–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hoheisel, U.; Rosner, J.; Mense, S. Innervation changes induced by inflammation of the rat thoracolumbar fascia. Neuroscience 2015, 300, 351–359. [Google Scholar] [CrossRef] [PubMed]
  46. Barry, C.M.; Kestell, G.; Gillan, M.; Haberberger, R.V.; Gibbins, I.L. Sensory nerve fibers containing calcitonin gene-related peptide in gastrocnemius, latissimus dorsi and erector spinae muscles and thoracolumbar fascia in mice. Neuroscience 2015, 291, 106–117. [Google Scholar] [CrossRef]
  47. Stecco, C.; Gagey, O.; Belloni, A.; Pozzuoli, A.; Porzionato, A.; Macchi, V.; Aldegheri, R.; de Caro, R.; Delmas, V. Anatomy of the deep fascia of the upper limb. Second part: Study of innervation. Morphologie 2007, 91, 38–43. [Google Scholar] [CrossRef]
  48. Satoh, M.; Yoshino, H.; Fujimura, A.; Hitomi, J.; Isogai, S. Three-layered architecture of the popliteal fascia that acts as a kinetic retinaculum for the hamstring muscles. Anat. Sci. Int. 2016, 91, 341–349. [Google Scholar] [CrossRef]
  49. Marpalli, S.; Mohandas Rao, K.G.; Venkatesan, P.; George, B.M. The morphological and microscopical characteristics of posterior layer of human thoracolumbar fascia; A potential source of low back pain. Morphologie 2021, 105, 308–315. [Google Scholar] [CrossRef]
  50. Sanchis-Alfonso, V.; Roselló-Sastre, E. Immunohistochemical analysis for neural markers of the lateral retinaculum in patients with isolated symptomatic patellofemoral malalignment. A neuroanatomic basis for anterior knee pain in the active young patient. Am. J. Sports Med. 2000, 28, 725–731. [Google Scholar] [CrossRef]
  51. Alhilou, A.M.; Shimada, A.; Svensson, C.I.; Ernberg, M.; Cairns, B.E.; Christidis, N. Density of nerve fibres and expression of substance P, NR2B-receptors and nerve growth factor in healthy human masseter muscle: An immunohistochemical study. J. Oral. Rehabil. 2021, 48, 35–44. [Google Scholar] [CrossRef]
  52. Stecco, A.; Gesi, M.; Stecco, C.; Stern, R. Fascial components of the myofascial pain syndrome. Curr. Pain Headache Rep. 2013, 17, 352. [Google Scholar] [CrossRef] [PubMed]
  53. Stecco, C.; Stern, R.; Porzionato, A.; MacChi, V.; Masiero, S.; Stecco, A.; de Caro, R. Hyaluronan within fascia in the etiology of myofascial pain. Surg. Radiol. Anat. 2011, 33, 891–896. [Google Scholar] [CrossRef] [PubMed]
  54. Casato, G.; Stecco, C.; Busin, R. Role of fasciae in nonspecific low back pain. Eur. J. Transl. Myol. 2019, 29, 8330. [Google Scholar] [CrossRef]
  55. Lucas, N.; Macaskill, P.; Irwig, L.; Moran, R.; Bogduk, N. Reliability of physical examination for diagnosis of myofascial trigger points: A systematic review of the literature. Clin. J. Pain. 2009, 25, 80–89. [Google Scholar] [CrossRef] [PubMed]
  56. Quintner, J.L.; Bove, G.M.; Cohen, M.L. A critical evaluation of the trigger point phenomenon. Rheumatology 2015, 54, 392–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Meister, M.R.; Sutcliffe, S.; Ghetti, C.; Chu, C.M.; Spitznagle, T.; Warren, D.K.; Lowder, J.L. Development of a standardized, reproducible screening examination for assessment of pelvic floor myofascial pain. Am. J. Obstet. Gynecol. 2019, 220, 255.e1–255.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. O’Reilly, M.S.; Boehm, T.; Shing, Y.; Fukai, N.; Vasios, G.; Lane, W.S.; Flynn, E.; Birkhead, J.R.; Olsen, B.R.; Folkman, J. Endostatin: An endogenous inhibitor of angiogenesis and tumor growth. Cell 1997, 88, 277–285. [Google Scholar] [CrossRef] [Green Version]
  59. Bloch, W.; Huggel, K.; Sasaki, T.; Grose, R.; Bugnon, P.; Addicks, K.; Timpl, R.; Werner, S. The angiogenesis inhibitor endostatin impairs blood vessel maturation during wound healing. FASEB J. 2000, 14, 2373–2376. [Google Scholar] [CrossRef]
  60. Wenzel, D.; Schmidt, A.; Reimann, K.; Hescheler, J.; Pfitzer, G.; Bloch, W.; Fleischmann, B.K. Endostatin, the proteolytic fragment of collagen XVIII, induces vasorelaxation. Circ. Res. 2006, 98, 1203–1211. [Google Scholar] [CrossRef] [Green Version]
  61. Fedorczyk, J.M.; Barr, A.E.; Rani, S.; Gao, H.G.; Amin, M.; Amin, S.; Litvin, J.; Barbe, M.F. Exposure-dependent increases in IL-1beta, substance P, CTGF, and tendinosis in flexor digitorum tendons with upper extremity repetitive strain injury. J. Orthop. Res. 2010, 28, 298–307. [Google Scholar] [CrossRef] [Green Version]
  62. Barr, A.E.; Barbe, M.F. Inflammation reduces physiological tissue tolerance in the development of work-related musculoskeletal disorders. J. Electromyogr. Kinesiol. 2004, 14, 77–85. [Google Scholar] [CrossRef] [PubMed]
  63. Gao, H.G.; Fisher, P.W.; Lambi, A.G.; Wade, C.K.; Barr-Gillespie, A.E.; Popoff, S.N.; Barbe, M.F. Increased serum and musculotendinous fibrogenic proteins following persistent low-grade inflammation in a rat model of long-term upper extremity overuse. PLoS ONE 2013, 8, e71875. [Google Scholar] [CrossRef] [Green Version]
  64. Barbe, M.F.; Gallagher, S.; Popoff, S.N. Serum biomarkers as predictors of stage of work-related musculoskeletal disorders. J. Am. Acad. Orthop. Surg. 2013, 21, 644–646. [Google Scholar] [CrossRef]
  65. Frara, N.; Fisher, P.W.; Zhao, Y.; Tarr, J.T.; Amin, M.; Popoff, S.N.; Barbe, M.F. Substance P increases CCN2 dependent on TGF-beta yet Collagen Type I via TGF-beta1 dependent and independent pathways in tenocytes. Connect. Tissue Res. 2018, 59, 30–44. [Google Scholar] [CrossRef] [PubMed]
  66. Driscoll, M.; Blyum, L. The presence of physiological stress shielding in the degenerative cycle of musculoskeletal disorders. J. Bodyw. Mov. Ther. 2011, 15, 335–342. [Google Scholar] [CrossRef]
  67. Xin, D.L.; Hadrévi, J.; Elliott, M.E.; Amin, M.; Harris, M.Y.; Barr-Gillespie, A.E.; Barbe, M.F. Effectiveness of conservative interventions for sickness and pain behaviors induced by a high repetition high force upper extremity task. BMC Neurosci. 2017, 18, 36. [Google Scholar] [CrossRef] [Green Version]
  68. Kasper, D.L.; Fauci, A.S.; Hauser, S.L.; Longo, D.L.; Jameson, J.L.; Loscalzo, J. Harrison’s Principles of Internal Medicine, 20th ed.; Mcgraw-hill: New York, NY, USA, 2018; pp. 222–223, 2637–2639, 2644–2645. [Google Scholar]
  69. Abdelmagid, S.M.; Barr, A.E.; Rico, M.; Amin, M.; Litvin, J.; Popoff, S.N.; Safadi, F.F.; Barbe, M.F. Performance of repetitive tasks induces decreased grip strength and increased fibrogenic proteins in skeletal muscle: Role of force and inflammation. PLoS ONE 2012, 7, e38359. [Google Scholar] [CrossRef] [Green Version]
  70. Berrueta, L.; Muskaj, I.; Olenich, S.; Butler, T.; Badger, G.J.; Colas, R.A.; Spite, M.; Serhan, C.N.; Langevin, H.M. Stretching impacts inflammation resolution in connective tissue. J. Cell Physiol. 2016, 231, 1621–1627. [Google Scholar] [CrossRef] [Green Version]
  71. Bove, G.M.; Harris, M.Y.; Zhao, H.; Barbe, M.F. Manual therapy as an effective treatment for fibrosis in a rat model of upper extremity overuse injury. J. Neurol. Sci. 2016, 361, 168–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Wilke, J.; Kalo, K.; Niederer, D.; Vogt, L.; Banzer, W. Gathering hints for myofascial force transmissionunder in vivo conditions: Are remote exercise effects age dependent? J. Sport Rehabil. 2019, 28, 758–763. [Google Scholar] [CrossRef]
  73. Wilke, J.; Schleip, R.; Yucesoy, C.A.; Banzer, W. Not merely a protective packing organ? A review of fascia and its force transmission capacity. J. Appl. Physiol. 2018, 124, 234–244. [Google Scholar] [CrossRef] [PubMed]
  74. Pavan, P.G.; Stecco, A.; Stern, R.; Stecco, C. Painful connections: Densification versus fibrosis of fascia. Curr. Pain Headache Rep. 2014, 18, 441. [Google Scholar] [CrossRef] [PubMed]
  75. Sölch, D. Ageing and restricted mobility. Frailty from the perspective of myofascial structural models. Z. Gerontol. Geriatr. 2015, 48, 35–40. [Google Scholar] [CrossRef]
  76. Carole, A. Gender equality and sport. Women. Beyond 2000, 2007, 1–40. [Google Scholar]
  77. Fink, J.S. Female athletes, women’s sport, and the sport media commercial complex: Have we really “come a long way, baby”? Sport Manag. Rev. 2015, 18, 331–342. [Google Scholar] [CrossRef]
  78. McNulty, K.L.; Elliott-Sale, K.J.; Dolan, E.; Swinton, P.A.; Ansdell, P.; Goodall, S.; Thomas, K.; Hicks, K.M. The effects of menstrual cycle phase on exercise performance in eumenorrheic women: A systematic review and meta-analysis. Sports Med. 2020, 50, 1813–1827. [Google Scholar] [CrossRef]
  79. Benton, M.J.; Hutchins, A.M.; Dawes, J.J. Effect of menstrual cycle on resting metabolism: A systematic review and metaanalysis. PLoS ONE 2020, 15, e0236025. [Google Scholar] [CrossRef] [PubMed]
  80. Van Pelt, R.E.; Gavin, K.M.; Kohrt, W.M. Regulation of body composition and bioenergetics by estrogens. Endocrinol. Metab. Clin. North Am. 2015, 44, 663–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Larsen, B.; Cox, A.; Colbey, C.; Drew, M.; McGuire, H.; de St Groth, B.F.; Hughes, D.; Vlahovich, N.; Waddington, G.; Burke, L.; et al. Inflammation and oral contraceptive use in female athletes before the Rio Olympic Games. Front. Physiol. 2020, 11, 497. [Google Scholar] [CrossRef] [PubMed]
  82. Hackney, A.C.; Kallman, A.L.; Aǧgön, E. Female sex hormones and the recovery from exercise: Menstrual cycle phase affects responses. Biomed. Hum. Kinet. 2019, 11, 87–89. [Google Scholar] [CrossRef] [Green Version]
  83. Lee, N.; Wingo, P.; Gwin, M.; Rubin, G.; Kendrick, J.; Webster, L. The reduction in risk of ovarian cancer associated with oral-contraceptive use. N. Engl. J. Med. 1987, 316, 650–655. [Google Scholar] [CrossRef]
  84. Frye, C.A. An overview of oral contraceptives: Mechanism of action and clinical use. Neurology 2006, 66, S29–S36. [Google Scholar] [CrossRef]
  85. Daniels, K.; Abma, J. Current contracpetive status among women aged 15–49. NCHS Data Brief. 2018, 327. [Google Scholar]
  86. Hackney, A.C. Sex Hormones and Physical Activity in Women: An Evolutionary Framework. In Sex Hormones, Exercise and Women; Springer: Cham, Switzerland, 2017; pp. 139–149. [Google Scholar]
  87. Aucouturier, J.; Baker, J.S.; Duché, P. Fat and carbohydrate metabolism during submaximal exercise in children. Sports Med. 2008, 38, 213–238. [Google Scholar] [CrossRef]
  88. Isacco, L.; Duché, P.; Boisseau, N. Influence of hormonal status on substrate utilization at rest and during exercise in the female population. Sports Med. 2012, 42, 327–342. [Google Scholar] [CrossRef]
  89. Lee, H.; Petrofsky, J.S.; Daher, N.; Berk, L.; Laymon, M. Differences in anterior cruciate ligament elasticity and force for knee flexion in women: Oral contraceptive users versus non-oral contraceptive users. Eur. J. Appl. Physiol. 2014, 114, 285–294. [Google Scholar] [CrossRef]
  90. Eiling, E.; Bryant, A.L.; Petersen, W.; Murphy, A.; Hohmann, E. Effects of menstrual-cycle hormone fluctuations on musculotendinous stiffness and knee joint laxity. Knee Surg Sport. Traumatol. Arthrosc. 2007, 15, 126–132. [Google Scholar] [CrossRef]
  91. Cauci, S.; Francescato, M.P.; Curcio, F. Combined oral contraceptives increase high-sensitivity C-reactive protein but not haptoglobin in female athletes. Sports Med. 2017, 47, 175–185. [Google Scholar] [CrossRef] [PubMed]
  92. Chaudhuri, O.; Cooper-White, J.; Janmey, P.A.; Mooney, D.J.; Shenoy, V.B. Effects of extracellular matrix viscoelasticity on cellular behaviour. Nature 2020, 584, 535–546. [Google Scholar] [CrossRef] [PubMed]
  93. Grolman, J.M.; Weinand, P.; Mooney, D.J. Extracellular matrix plasticity as a driver of cell spreading. Proc. Natl. Acad. Sci. USA 2020, 117, 25999–26007. [Google Scholar] [CrossRef] [PubMed]
  94. Stecco, C.; Fede, C.; Macchi, V.; Porzionato, A.; Petrelli, L.; Biz, C.; Stern, R.; De Caro, R. The fasciacytes: A new cell devoted to fascial gliding regulation. Clin. Anat. 2018, 31, 667–676. [Google Scholar] [CrossRef]
  95. Juel, C.; Bangsbo, J.; Graham, T.; Saltin, B. Lactate and potassium fluxes from human skeletal muscle during and after intense, dynamic, knee extensor exercise. Acta Physiol. Scand. 1990, 140, 147–159. [Google Scholar] [CrossRef] [PubMed]
  96. Cowman, M.K.; Schmidt, T.A.; Raghavan, P.; Stecco, A. Viscoelastic properties of hyaluronan in physiological conditions. F1000Reseasrch 2015, 4, 622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Tadmor, R.; Chen, N.; Israelachvili, J.N. Thin film rheology and lubricity of hyaluronic acid solutions at a normal physiological concentration. J. Biomed. Mater. Res. 2002, 61, 514–523. [Google Scholar] [CrossRef] [PubMed]
  98. Hughes, E.J.; McDermott, K.; Funk, M.F. Evaluation of hyaluronan content in areas of densification compared to adjacent areas of fascia. J. Bodyw. Mov. Ther. 2019, 23, 324–328. [Google Scholar] [CrossRef]
  99. Pirri, C.; Fede, C.; Stecco, A.; Guidolin, D.; Fan, C.; De Caro, R.; Stecco, C. Ultrasound imaging of crural fascia and epimysial fascia thicknesses in basketball players with previous ankle sprains versus healthy subjects. Diagnostics 2021, 11, 177. [Google Scholar] [CrossRef]
  100. Pirri, C.; Guidolin, D.; Fede, C.; Macchi, V.; De Caro, R.; Stecco, C. Ultrasound imaging of brachial and antebrachial fasciae. Diagnostics 2021, 11, 2261. [Google Scholar] [CrossRef] [PubMed]
  101. Pirri, C.; Pirri, N.; Guidolin, D.; Macchi, V.; De Caro, R.; Stecco, C. Ultrasound imaging of the superficial fascia in the upper limb: Arm and forearm. Diagnostics 2022, 12, 1884. [Google Scholar] [CrossRef]
  102. Pirri, C.; Pirri, N.; Porzionato, A.; Boscolo-Berto, R.; De Caro, R.; Stecco, C. Inter- and intra-rater reliability of ultrasound measurements of superficial and deep fasciae thickness in upper limb. Diagnostics 2022, 12, 2195. [Google Scholar] [CrossRef]
  103. Yang, C.; Huang, X.; Li, Y.; Sucharit, W.; Sirasaporn, P.; Eungpinichpong, W. Acute effects of percussive massage therapy on thoracolumbar fascia thickness and ultrasound echo intensity in healthy male individuals: A randomized controlled trial. Int. J. Environ. Res. Public Health 2023, 20, 1073. [Google Scholar] [CrossRef]
  104. Wilke, J.; Macchi, V.; De Caro, R.; Stecco, C. Fascia thickness, aging and flexibility: Is there an association? J. Anat. 2019, 234, 43–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Fantoni, I.; Biz, C.; Fan, C.; Pirri, C.; Fede, C.; Petrelli, L.; Ruggieri, P.; De Caro, R.; Stecco, C. Fascia Lata alterations in hip osteoarthritis: An observational cross-sectional study. Life 2021, 11, 1136. [Google Scholar] [CrossRef] [PubMed]
  106. Flores, D.V.; Mejía Gómez, C.; Estrada-Castrillón, M.; Smitaman, E.; Pathria, M.N. MR imaging of muscle trauma: Anatomy, biomechanics, pathophysiology, and imaging appearance. RadioGraphics 2018, 38, 124–148. [Google Scholar] [CrossRef] [PubMed]
  107. Draghi, F.; Gitto, S.; Bortolotto, C.; Draghi, A.G.; Ori Belometti, G. Imaging of plantar fascia disorders: Findings on plain radiography, ultrasound and magnetic resonance imaging. Insights Imaging 2017, 8, 69–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Schleip, R. Fascial plasticity—A new neurobiological explanation: Part 1. J. Bodyw. Mov. Ther. 2003, 7, 11–19. [Google Scholar] [CrossRef]
  109. Soligard, T.; Schwellnus, M.; Alonso, J.M.; Bahr, R.; Clarsen, B.; Dijkstra, H.P.; Gabbett, T.; Gleeson, M.; Hägglund, M.; Hutchinson, M.R.; et al. How much is too much? (Part 1) International Olympic Committee consensus statement on load in sport and risk of injury. Br. J. Sports Med. 2016, 50, 1030–1041. [Google Scholar] [CrossRef] [Green Version]
  110. Bahr, R. No injuries, but plenty of pain? On the methodology for recording overuse symptoms in sports. Br. J. Sports Med. 2009, 43, 966–972. [Google Scholar] [CrossRef]
  111. Clarsen, B.; Myklebust, G.; Bahr, R. Development and validation of a new method for the registration of overuse injuries in sports injury epidemiology: The Oslo Sports Trauma Research Centre (OSTRC) overuse injury questionnaire. Br. J. Sports Med. 2013, 47, 495–502. [Google Scholar] [CrossRef] [Green Version]
  112. Barnes, J. Myofascial Release: The Search for Excellence, 10th ed.; Rehabilitation Services Inc.: Paoli, PA, USA, 1990. [Google Scholar]
  113. Domingo, T.; Blasi, J.; Casals, M.; Mayoral, V.; Ortiz-Sagristá, J.C.; Miguel-Pérez, M. Is interfascial block with ultrasound-guided puncture useful in treatment of myofascial pain of the trapezius muscle? Clin. J. Pain 2011, 27, 297–303. [Google Scholar] [CrossRef]
  114. Kanamoto, H.; Orita, S.; Inage, K.; Shiga, Y.; Abe, K.; Eguchi, Y.; Ohtori, S. Effect of ultrasound-guided hydrorelease of the multifidus muscle on acute low back pain. J. Ultrasound Med. 2021, 40, 981–987. [Google Scholar] [CrossRef] [PubMed]
  115. Barnes, M.F. The basic science of myofascial release: Morphologic change in connective tissue. J. Bodyw. Mov. Ther. 1997, 1, 231–238. [Google Scholar] [CrossRef]
  116. Ward, R.C. Myofascial Release Concepts; Wiliams & Wilkins: Baltimore, MD, USA, 1993. [Google Scholar]
  117. MacDonald, G.Z.; Penney, M.D.; Mullaley, M.E.; Cuconato, A.L.; Drake, C.D.; Behm, D.G.; Button, D.C. An acute bout of self-myofascial release increases range of motion without a subsequent decrease in muscle activation or force. J. Strength Cond. Res. 2013, 27, 812–821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Pearcey, G.E.; Bradbury-Squires, D.J.; Kawamoto, J.E.; Drinkwater, E.J.; Behm, D.G.; Button, D.C. Foam rolling for delayed-onset muscle soreness and recovery of dynamic performance measures. J. Athl. Train. 2015, 50, 5–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Cavanaugh, M.T.; Döweling, A.; Young, J.D.; Quigley, P.J.; Hodgson, D.D.; Whitten, J.H.; Reid, J.C.; Aboodarda, S.J.; Behm, D.G. An acute session of roller massage prolongs voluntary torque development and diminishes evoked pain. Eur. J. Appl. Physiol. 2017, 117, 109–117. [Google Scholar] [CrossRef]
  120. Ajimsha, M.S. Effectiveness of direct vs indirect technique myofascial release in the management of tension-type headache. J. Bodyw. Mov. Ther. 2011, 15, 431–435. [Google Scholar] [CrossRef] [PubMed]
  121. Ramos-González, E.; Moreno-Lorenzo, C.; Matarán-Peñarrocha, G.A.; Guisado-Barrilao, R.; Aguilar-Ferrándiz, M.E.; Castro-Sánchez, A.M. Comparative study on the effectiveness of myofascial release manual therapy and physical therapy for venous insufficiency in postmenopausal women. Complement. Ther. Med. 2012, 20, 291–298. [Google Scholar] [CrossRef]
  122. Arguisuelas, M.D.; Lisón, J.F.; Doménech-Fernández, J.; Martínez-Hurtado, I.; Salvador Coloma, P.; Sánchez-Zuriaga, D. Effects of myofascial release in erector spinae myoelectric activity and lumbar spine kinematics in non-specific chronic low back pain: Randomized controlled trial. Clin. Biomech. 2019, 63, 27–33. [Google Scholar] [CrossRef]
  123. Ichikawa, K.; Takei, H.; Usa, H.; Mitomo, S.; Ogawa, D. Comparative analysis of ultrasound changes in the vastus lateralis muscle following myofascial release and thermotherapy: A pilot study. J. Bodyw. Mov. Ther. 2015, 19, 327–336. [Google Scholar] [CrossRef]
  124. Desai, M.J.; Bean, M.C.; Heckman, T.W.; Jayaseelan, D.; Moats, N.; Nava, A. Treatment of myofascial pain. Pain Manag. 2013, 3, 67–79. [Google Scholar] [CrossRef]
  125. Hou, C.R.; Tsai, L.C.; Cheng, K.F.; Chung, K.C.; Hong, C.Z. Immediate effects of various physical therapeutic modalities on cervical myofascial pain and trigger-point sensitivity. Arch. Phys. Med. Rehabil. 2002, 83, 1406–1414. [Google Scholar] [CrossRef]
  126. Ozsoy, G.; Ilcin, N.; Ozsoy, I.; Gurpinar, B.; Buyukturan, O.; Buyukturan, B.; Kararti, C.; Sas, S. The effects of myofascial release technique combined with core stabilization exercise in elderly with non-specific low back pain: A randomized controlled, single-blind study. Clin. Interv. Aging. 2019, 14, 1729–1740. [Google Scholar] [CrossRef] [Green Version]
  127. E Silva, D.C.C.M.; de Andrade Alexandre, D.J.; Silva, J.G. Immediate effect of myofascial release on range of motion, pain and biceps and rectus femoris muscle activity after total knee replacement. J. Bodyw. Mov. Ther. 2018, 22, 930–936. [Google Scholar] [CrossRef] [PubMed]
  128. Zalta, J. Massage therapy protocol for post-anterior cruciate ligament reconstruction patellofemoral pain syndrome: A case report. Int. J. Ther. Massage Bodyw. 2008, 1, 11–21. [Google Scholar] [CrossRef]
  129. Hägg, O.; Fritzell, P.; Nordwall, A.; Swedish Lumbar Spine Study Group. The clinical importance of changes in outcome scores after treatment for chronic low back pain. Eur. Spine J. 2003, 12, 12–20. [Google Scholar] [CrossRef] [PubMed]
  130. Elsayyad, M.M.; Abdel-Aal, N.M.; Helal, M.E. Effect of adding neural mobilization versus myofascial release to stabilization exercises after lumbar spine fusion: A randomized controlled trial. Arch. Phys. Med. Rehabil. 2021, 102, 251–260. [Google Scholar] [CrossRef]
  131. Meltzer, K.R.; Cao, T.V.; Schad, J.F.; King, H.; Stoll, S.T.; Standley, P.R. In vitro modeling of repetitive motion injury and myofascial release. J. Bodyw. Mov. Ther. 2010, 14, 162–171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Melzack, R.; Wall, P.D. Pain mechanisms: A new theory. Pain Forum. 1996, 5, 3–11. [Google Scholar] [CrossRef]
  133. Coelho, N.M.; McCulloch, C.A. Contribution of collagen adhesion receptors to tissue fibrosis. Cell Tissue Res. 2016, 365, 521–538. [Google Scholar] [CrossRef]
  134. Kawanishi, K.; Kudo, S.; Yokoi, K. Relationship between gliding and lateral femoral pain in patients with trochanteric fracture. Arch. Phys. Med. Rehabil. 2020, 101, 457–463. [Google Scholar] [CrossRef]
  135. Kawanishi, K.; Fukuda, D.; Niwa, H.; Okuno, T.; Miyashita, T.; Kitagawa, T.; Kudo, S. Relationship between tissue gliding of the lateral thigh and gait parameters after trochanteric fractures. Sensors 2022, 22, 3842. [Google Scholar] [CrossRef]
  136. Park, J.J.; Lee, H.S.; Kim, J.H. Effect of acute self-myofascial release on pain and exercise performance for cycling club members with iliotibial band friction syndrome. Int. J. Environ. Res. Public Health 2022, 19, 15993. [Google Scholar] [CrossRef] [PubMed]
  137. Sulowska-Daszyk, I.; Skiba, A. The influence of self-myofascial release on muscle flexibility in long-distance runners. Int. J. Environ. Res. Public Health 2022, 19, 457. [Google Scholar] [CrossRef] [PubMed]
  138. Russell, M.; West, D.J.; Harper, L.D.; Cook, C.J.; Kilduff, L.P. Half-time strategies to enhance second-half performance in team-sports players: A review and recommendations. Sports Med. 2015, 45, 353–364. [Google Scholar] [CrossRef] [PubMed]
  139. Kaya, S.; Cug, M.; Behm, D.G. Foam rolling during a simulated half-time attenuates subsequent soccer-specific performance decrements. J. Bodyw. Mov. Ther. 2021, 26, 193–200. [Google Scholar] [CrossRef] [PubMed]
  140. Okamoto, T.; Masuhara, M.; Ikuta, K. Acute effects of self-myofascial release using a foam roller on arterial function. J. Strength Cond. Res. 2014, 28, 69–73. [Google Scholar] [CrossRef] [Green Version]
  141. Cheung, K.; Hume, P.; Maxwell, L. Delayed onset muscle soreness: Treatment strategies and performance factors. Sports Med. 2003, 33, 145–164. [Google Scholar] [CrossRef]
  142. Halperin, I.; Aboodarda, S.J.; Button, D.C.; Andersen, L.L.; Behm, D.G. Roller massager improves range of motion of plantar flexor muscles without subsequent decreases in force parameters. Int. J. Sports Phys. Ther. 2014, 9, 92–102. [Google Scholar]
  143. Curran, P.F.; Fiore, R.D.; Crisco, J.J. A comparison of the pressure exerted on soft tissue by 2 myofascial rollers. J. Sport Rehabil. 2008, 17, 432–442. [Google Scholar] [CrossRef] [Green Version]
  144. Swann, E.; Graner, S.J. Uses of manual-therapy techniques in pain management. Athl. Ther. Today 2002, 7, 14–17. [Google Scholar] [CrossRef]
  145. Behm, D.G.; Chaouachi, A. A review of the acute effects of static and dynamic stretching on performance. Eur. J. Appl. Physiol. 2011, 111, 2633–2651. [Google Scholar] [CrossRef]
  146. Macdonald, G.Z.; Button, D.C.; Drinkwater, E.J.; Behm, D.G. Foam rolling as a recovery tool after an intense bout of physical activity. Med. Sci. Sports Exerc. 2014, 46, 131–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Healey, K.C.; Hatfield, D.L.; Blanpied, P.; Dorfman, L.R.; Riebe, D. The effects of myofascial release with foam rolling on performance. J. Strength Cond. Res. 2014, 28, 61–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Schroeder, A.N.; Best, T.M. Is self myofascial release an effective preexercise and recovery strategy? A literature review. Curr. Sports Med. Rep. 2015, 14, 200–208, Erratum in Curr. Sports Med. Rep. 2015, 14, 352. [Google Scholar] [CrossRef] [PubMed]
  149. Tejero-Fernández, V.; Membrilla-Mesa, M.; Galiano-Castillo, N.; Arroyo-Morales, M. Immunological effects of massage after exercise: A systematic review. Phys. Ther. Sport 2015, 16, 187–192. [Google Scholar] [CrossRef]
  150. Hendricks, S.; Hill, H.; Hollander, S.D.; Lombard, W.; Parker, R. Effects of foam rolling on performance and recovery: A systematic review of the literature to guide practitioners on the use of foam rolling. J. Bodyw. Mov. Ther. 2020, 24, 151–174. [Google Scholar] [CrossRef]
  151. Skinner, B.; Moss, R.; Hammond, L. A systematic review and meta-analysis of the effects of foam rolling on range of motion, recovery and markers of athletic performance. J. Bodyw. Mov. Ther. 2020, 24, 105–122. [Google Scholar] [CrossRef]
  152. Wiewelhove, T.; Döweling, A.; Schneider, C.; Hottenrott, L.; Meyer, T.; Kellmann, M.; Pfeiffer, M.; Ferrauti, A. A Meta-Analysis of the Effects of Foam Rolling on Performance and Recovery. Front. Physiol. 2019, 10, 376. [Google Scholar] [CrossRef] [Green Version]
  153. Hughes, G.A.; Ramer, L.M. duration of myofascial rolling for optimal recovery, range of motion, and performance: A systematic review of the literature. Int. J. Sports Phys. Ther. 2019, 14, 845–859. [Google Scholar] [CrossRef]
  154. Ferreira, R.M.; Martins, P.N.; Goncalves, R.S. Effects of Self-myofascial Release Instruments on Performance and Recovery: An Umbrella Review. Int. J. Exerc. Sci. 2022, 15, 861–883. [Google Scholar]
  155. Jay, K.; Sundstrup, E.; Søndergaard, S.D.; Behm, D.; Brandt, M.; Særvoll, C.A.; Jakobsen, M.D.; Andersen, L.L. Specific and cross over effects of massage for muscle soreness: Randomized controlled trial. Int. J. Sports Phys. Ther. 2014, 9, 82–91. [Google Scholar]
  156. Ostiak, W.; Kaczmarek-Maciejewska, M.; Kasprzak, P. Foot and shin in terms of Anatomy Trains. J. Orthop. Trauma Surg. Relat. Res. 2011, 5, 38–46. [Google Scholar]
  157. Clinical Guideline Subcommittee on Low Back Pain; American Osteopathic Association. American Osteopathic Association guidelines for osteopathic manipulative treatment (OMT) for patients with low back pain. J. Am. Osteopath. Assoc. 2010, 110, 653–666. [Google Scholar]
  158. Dardzinski, J.A.; Ostrov, B.E.; Hamann, L.S. Myofascial pain unresponsive to standard treatment: Successful use of a strain and counterstrain technique with physical therapy. J. Clin. Rheumatol. 2000, 6, 169–174. [Google Scholar] [CrossRef] [PubMed]
  159. Lew, J.; Kim, J.; Nair, P. Comparison of dry needling and trigger point manual therapy in patients with neck and upper back myofascial pain syndrome: A systematic review and meta-analysis. J. Man. Manip. Ther. 2021, 29, 136–146. [Google Scholar] [CrossRef] [PubMed]
  160. Sbardella, S.; La Russa, C.; Bernetti, A.; Mangone, M.; Guarnera, A.; Pezzi, L.; Paoloni, M.; Agostini, F.; Santilli, V.; Saggini, R.; et al. Muscle energy technique in the rehabilitative treatment for acute and chronic non-specific neck pain: A systematic Review. Healthcare 2021, 9, 746. [Google Scholar] [CrossRef]
  161. Chen, Z.; Wu, J.; Wang, X.; Wu, J.; Ren, Z. The effects of myofascial release technique for patients with low back pain: A systematic review and meta-analysis. Complement. Ther. Med. 2021, 59, 102737. [Google Scholar] [CrossRef] [PubMed]
  162. Arguisuelas, M.D.; Lisón, J.F.; Sánchez-Zuriaga, D.; Martínez-Hurtado, I.; Doménech-Fernández, J. Effects of myofascial release in nonspecific chronic low back pain: A randomized clinical trial. Spine 2017, 42, 627–634. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kodama, Y.; Masuda, S.; Ohmori, T.; Kanamaru, A.; Tanaka, M.; Sakaguchi, T.; Nakagawa, M. Response to Mechanical Properties and Physiological Challenges of Fascia: Diagnosis and Rehabilitative Therapeutic Intervention for Myofascial System Disorders. Bioengineering 2023, 10, 474. https://doi.org/10.3390/bioengineering10040474

AMA Style

Kodama Y, Masuda S, Ohmori T, Kanamaru A, Tanaka M, Sakaguchi T, Nakagawa M. Response to Mechanical Properties and Physiological Challenges of Fascia: Diagnosis and Rehabilitative Therapeutic Intervention for Myofascial System Disorders. Bioengineering. 2023; 10(4):474. https://doi.org/10.3390/bioengineering10040474

Chicago/Turabian Style

Kodama, Yuya, Shin Masuda, Toshinori Ohmori, Akihiro Kanamaru, Masato Tanaka, Tomoyoshi Sakaguchi, and Masami Nakagawa. 2023. "Response to Mechanical Properties and Physiological Challenges of Fascia: Diagnosis and Rehabilitative Therapeutic Intervention for Myofascial System Disorders" Bioengineering 10, no. 4: 474. https://doi.org/10.3390/bioengineering10040474

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop