Mechanotherapy: how physical therapists' prescription of exercise promotes tissue repair

K M Khan, A Scott, K M Khan, A Scott

Abstract

Mechanotransduction is the physiological process where cells sense and respond to mechanical loads. This paper reclaims the term "mechanotherapy" and presents the current scientific knowledge underpinning how load may be used therapeutically to stimulate tissue repair and remodelling in tendon, muscle, cartilage and bone. The purpose of this short article is to answer a frequently asked question "How precisely does exercise promote tissue healing?" This is a fundamental question for clinicians who prescribe exercise for tendinopathies, muscle tears, non-inflammatory arthropathies and even controlled loading after fractures. High-quality randomised controlled trials and systematic reviews show that various forms of exercise or movement prescription benefit patients with a wide range of musculoskeletal problems.1(-)4 But what happens at the tissue level to promote repair and remodelling of tendon, muscle, articular cartilage and bone? The one-word answer is "mechanotransduction", but rather than finishing there and limiting this paper to 95 words, we provide a short illustrated introduction to this remarkable, ubiquitous, non-neural, physiological process. We also re-introduce the term "mechanotherapy" to distinguish therapeutics (exercise prescription specifically to treat injuries) from the homeostatic role of mechanotransduction. Strictly speaking, mechanotransduction maintains normal musculoskeletal structures in the absence of injury. After first outlining the process of mechanotransduction, we provide well-known clinical therapeutic examples of mechanotherapy-turning movement into tissue healing.

Conflict of interest statement

Competing interests: None.

Figures

Figure 1
Figure 1
Tendon cell undergoing (A,B) shear and (C) compression during a tendon-loading cycle.
Figure 2
Figure 2
Tendon tissue provides an example of cell–cell communication. (A) The intact tendon consists of extracellular matrix (including collagen) and specialised tendon cells (arrowheads). (B) Tendon with collagen removed to reveal the interconnecting cell network. Cells are physically in contact throughout the tendon, facilitating cell–cell communication. Gap junctions are the specialised regions where cells connect and communicate small charged particles. They can be identified by their specific protein connexin 43. (C–E) Time course of cell–cell communication from (C) beginning, through (D) the midpoint to (E) the end. The signalling proteins for this step include calcium (red spheres) and inositol triphosphate (IP3).
Figure 3
Figure 3
Mechanical loading stimulates protein synthesis at the cellular level. (A) A larger scale image of the tendon cell network for orientation. We focus on one very small region. (B) Zooming in on this region reveals the cell membrane, the integrin proteins that bridge the intracellular and extra-cellular regions, and the cytoskeleton, which functions to maintain cell integrity and distribute mechanical load. The cell nucleus and the DNA are also illustrated. (C) With movement (shearing is illustrated), the integrin proteins activate at least two distinct pathways. (D) One involves the cytoskeleton that is in direct physical communication with the nucleus (ie, tugging the cytoskeleton sends a physical signal to the cell nucleus). Another pathway is triggered by integrins activating a series of biochemical signalling agents which are illustrated schematically. After a series of intermediate steps those biochemical signals also influence gene expression in the nucleus. (E). Once the cell nucleus receives the appropriate signals, normal cellular processes are engaged. mRNA is transcribed and shuttled to the endoplasmic reticulum in the cell cytoplasm, where it is translated into protein. The protein is secreted and incorporated into extracellular matrix. (F) In sum, the mechanical stimulus on the outside of the cell promotes intracellular processes leading to matrix remodelling.

References

    1. Loudon JK, Santos MJ, Franks L, et al. The effectiveness of active exercise as an intervention for functional ankle instability: A systematic review. Sports Med 2008;38:553–63
    1. Rabin A. Is there evidence to support the use of eccentric strengthening exercises to decrease pain and increase function in patients with patellar tendinopathy? Phys Ther 2006;86:450–6
    1. Smidt N, de Vet HC, Bouter LM, et al. Effectiveness of exercise therapy: A best-evidence summary of systematic reviews. Aust J Physiother 2005;51:71–85
    1. Taylor NF, Dodd KJ, Damiano DL. Progressive resistance exercise in physical therapy: A summary of systematic reviews. Phys Ther 2005;85:1208–23
    1. Duncan RL, Turner CH. Mechanotransduction and the functional response of bone to mechanical strain. Calcif Tissue Int 1995;57:344–58
    1. McElhaney JH, Stalnaker R, Bullard R. Electric fields and bone loss of disuse. J Biomech 1968;1:47–52
    1. Wall ME, Banes AJ. Early responses to mechanical load in tendon: Role for calcium signaling, gap junctions and intercellular communication. J Musculoskeletal Neuronal Interact 2005;5:70–84
    1. Cheema U, Brown R, Mudera V, et al. Mechanical signals and IGF-I gene splicing in vitro in relation to development of skeletal muscle. J Cell Physiol 2005;202:67–75
    1. Durieux AC, Desplanches D, Freyssenet D, et al. Mechanotransduction in striated muscle via focal adhesion kinase. Biochem Soc Trans 2007;35:1312–13
    1. Arnoczky SP, Tian T, Lavagnino M, et al. Activation of stress-activated protein kinases (SAPK) in tendon cells following cyclic strain: The effects of strain frequency, strain magnitude, and cytosolic calcium. J Orthop Res 2002;20:947–52
    1. Banes AJ, Tzsuzaki M, Wall ME, et al. In: Woo SL, Renstrom P, Arnoczky SP, eds. The molecular biology of tendinopathy: signaling and response pathways in tenocytes, in tendinopathy in athletes Malden: Blackwell Publishing, 2007
    1. Waggett AD, Benjamin M, Ralphs JR. Connexin 32 and 43 gap junctions differentially modulate tenocyte response to cyclic mechanical load. Eur J Cell Biol 2006;85:1145–54
    1. Knobloch TJ, Madhavan S, Nam J, et al. Regulation of chondrocytic gene expression by biomechanical signals. Crit Rev Eukaryot Gene Expr 2008;18:139–50
    1. Alberts B, Johnson A, Lewis J, et al. Molecular biology of the cell. New York: Garland Science, 2002
    1. Chen CS, Mrksich M, Huang S, et al. Geometric control of cell life and death. Science 1997;276:1425–8
    1. Chicurel ME, Singer RH, Meyer CJ, et al. Integrin binding and mechanical tension induce movement of mRNA and ribosomes to focal adhesions. Nature 1998;392:730–3
    1. Huang S, Ingber DE. The structural and mechanical complexity of cell-growth control. Nat Cell Biol 1999;1:E131–8
    1. Meyer CJ, Alenghat FJ, Rim P, et al. Mechanical control of cyclic AMP signalling and gene transcription through integrins. Nat Cell Biol 2000;2:666–8
    1. Lavagnino M, Arnoczky SP. In vitro alterations in cytoskeletal tensional homeostasis control gene expression in tendon cells. J Orthop Res 2005;23:1211–18
    1. Lavagnino M, Arnoczky SP, Tian T, et al. Effect of amplitude and frequency of cyclic tensile strain on the inhibition of MMP-1 mRNA expression in tendon cells: An in vitro study. Connect Tissue Res 2003;44:181–7
    1. Almekinders LC, Banes AJ, Ballenger CA. Effects of repetitive motion on human fibroblasts. Med Sci Sports Exerc 1993;25:603–7
    1. Archambault J, Tsuzaki M, Herzog W, et al. Stretch and interleukin-1beta induce matrix metalloproteinases in rabbit tendon cells in vitro. J Orthop Res 2002;20:36–9
    1. Archambault JM, Elfervig-Wall MK, Tsuzaki M, et al. Rabbit tendon cells produce MMP-3 in response to fluid flow without significant calcium transients. J Biomech 2002;35:303–9
    1. Banes AJ, Tsuzaki M, Hu P, et al. PDGF-BB, IGF-I and mechanical load stimulate DNA synthesis in avian tendon fibroblasts in vitro. J Biomech 1995;28:1505–13
    1. Banes AJ, Weinhold P, Yang X, et al. Gap junctions regulate responses of tendon cells ex vivo to mechanical loading. Clin Orthop 1999;367 Suppl:S356–70
    1. Tsuzaki M, Bynum D, Almekinders L, et al. ATP modulates load-inducible IL-1beta, COX 2, and MMP-3 gene expression in human tendon cells. J Cell Biochem 2003;89:556–62
    1. Upchurch GR, Jr, Loscalzo J, Banes AJ. Changes in the amplitude of cyclic load biphasically modulate endothelial cell DNA synthesis and division. Vasc Med 1997;2:19–24
    1. Vadiakas GP, Banes AJ. Verapamil decreases cyclic load-induced calcium incorporation in ros 17/2.8 osteosarcoma cell cultures. Matrix 1992;12:439–47
    1. Archambault JM, Hart DA, Herzog W. Response of rabbit achilles tendon to chronic repetitive loading. Connect Tissue Res 2001;42:13–23
    1. Hart DA, Natsu-ume T, Sciore P, et al. Mechanobiology: Similarities and differences between in vivo and in vitro analysis at the functional and molecular levels. Recent Res Devel Biophys Biochem 2002;2:153–177
    1. Scott A, Cook JL, Hart DA, et al. Tenocyte responses to mechanical loading in vivo: A role for local insulin-like growth factor 1 signaling in early tendinosis in rats. Arthritis Rheum 2007;56:871–81
    1. Tasevski V, Sorbetti JM, Chiu SS, et al. Influence of mechanical and biological signals on gene expression in human mg-63 cells: Evidence for a complex interplay between hydrostatic compression and vitamin D3 or TGF-beta1 on MMP-1 and MMP-3 mrna levels. Biochem Cell Biol 2005;83):96–107
    1. Frost HM. Bone “mass” and the “mechanostat”: A proposal. Anat Rec 1987;219:1–9
    1. Frost HM. Bone’s mechanostat: A 2003 update. Anat Rec A Discov Mol Cell Evol Biol 2003;275:1081–101
    1. Sallis RE. Exercise is medicine and physicians need to prescribe it!. Br J Sports Med 2009 Jan;43:3–4
    1. Blair SN. Physical inactivity: the biggest public health problem of the 21st century. Br J Sports Med 2009;43:1–2
    1. Olesen JL, Heinemeier KM, Haddad F, et al. Expression of insulin-like growth factor I, insulin-like growth factor binding proteins, and collagen mrna in mechanically loaded plantaris tendon. J Appl Physiol 2006;101:183–8
    1. Heinemeier KM, Olesen JL, Schjerling P, et al. Short-term strength training and the expression of myostatin and igf-i isoforms in rat muscle and tendon: Differential effects of specific contraction types. J Appl Physiol 2007;102:573–81
    1. Olesen JL, Heinemeier KM, Gemmer C, et al. Exercise-dependent IGF-I, IGFBPS, and type I collagen changes in human peritendinous connective tissue determined by microdialysis. J Appl Physiol 2007;102:214–20
    1. Ohberg L, Lorentzon R, Alfredson H. Eccentric training in patients with chronic achilles tendinosis: Normalised tendon structure and decreased thickness at follow up. Br J Sports Med 2004;38:8–11; discussion 11.
    1. Boyer MI, Goldfarb CA, Gelberman RH. Recent progress in flexor tendon healing. The modulation of tendon healing with rehabilitation variables. J Hand Ther 2005;18:80–5; quiz 86.
    1. Goldspink G. Gene expression in muscle in response to exercise. J Muscle Res Cell Motil 2003;24:121–6
    1. Jarvinen TA, Jarvinen TL, Kaariainen M, et al. Muscle injuries: Optimising recovery. Best Pract Res Clin Rheumatol 2007;21:317–31
    1. Alfredson H, Lorentzon R. Superior results with continuous passive motion compared to active motion after periosteal transplantation. A retrospective study of human patella cartilage defect treatment. Knee Surg Sports Traumatol Arthrosc 1999;7:232–8
    1. Challis MJ, Jull GJ, Stanton WR, et al. Cyclic pneumatic soft-tissue compression enhances recovery following fracture of the distal radius: A randomised controlled trial. Aust J Physiother 2007;53:247–52
    1. Challis MJ, Gaston P, Wilson K, et al. Cyclic pneumatic soft-tissue compression accelerates the union of distal radial osteotomies in an ovine model. J Bone Joint Surg Br 2006;88:411–15
    1. Challis MJ, Welsh MK, Jull GA, et al. Effect of cyclic pneumatic soft tissue compression on simulated distal radius fractures. Clin Orthop Relat Res 2005;433:183–8

Source: PubMed

3
Tilaa