Preface:
Significant research over the past two decades has established that extracellular matrix (ECM) elasticity, or stiffness, impacts fundamental cell processes including spreading, growth, proliferation, migration, differentiation, and organoid formation. Linearly elastic polyacrylamide hydrogels and polydimethylsiloxane (PDMS) elastomers coated with ECM proteins have become widely-used tools for assessing the role of stiffness, and results from these experiments are often assumed to reproduce the effect of the mechanical environment experienced by cells in vivo. However, tissues and ECMs are not linearly elastic materials – they in fact exhibit far more complex mechanical behaviors, including viscoelasticity, or a time-dependent response to loading or deformation, as well as mechanical plasticity and nonlinear elasticity. Recent work has revealed that matrix viscoelasticity regulates these same fundamental cell processes, and importantly can promote behaviors not observed with elastic hydrogels in both 2D and 3D culture microenvironments. These important findings have provided new insights into cell-matrix interactions and have given context as to how these interactions differentially modulate mechano-sensitive molecular pathways in cells. Moreover, these results indicate new design guidelines for the next generation of biomaterials that better match tissue and ECM mechanics for in vitro tissue models and applications in regenerative medicine.
While indications of the impacts of the mechanical properties of culture substrates on cell behaviours have long been present, it is only in recent times that this concept has become widely accepted by the scientific community. Earlier studies demonstrating the impact of substrate mechanics on cell structure and proliferation were overshadowed by an emphasis on cell biology on genetics and biochemistry1,2. The situation began to change in the late 1990’s when, using polyacrylamide hydrogels of varying elastic moduli coated with ECM proteins as cell culture substrates, Pelham and Wang showed that substrate stiffness affected cell-ECM adhesion, spreading, and migration3. Since this study, numerous groups have used polyacrylamide gels, and a variety of other material systems with tunable elastic moduli, to show that substrate stiffness impacts various other processes, including proliferation and apoptosis, stem cell differentiation, breast cancer progression and response to drugs4–6. Mechanistically, the current view is that cells exert traction forces using actomyosin-based contractility when coupled to substrates through integrin-based adhesions, or other cell-surface links, and they sense variations in substrate stiffness through differing magnitudes or extents of integrin and syndecan clustering and associated signaling, conformational changes in mechanosensitive proteins such as talin, vinculin, or lamin, activation of mechanosensitive ion channels (such as piezo1), and downstream activation of transcription factor activity7–10. While changes in ECM mechanics are sensed by cells over short timescales, these can impact long term cellular processes such as differentiation, fibrosis, and malignancy through continued sensing, mechanical memory, and changes in the epigenome11–13. Reported tissue elastic moduli vary from ~100s of Pascals in brain and fat tissue all the way up to 10s of GPa in bone14,15. Further, alterations in tissue mechanics are observed in development and in various diseases and have been linked to cell phenotype in these contexts16,17. Thus, the current consensus is that ECM stiffness plays a key role in regulating development, homeostasis, regenerative processes, and disease progression.
Living tissues and organisms appear as macroscopically solid objects, however they behave very differently to what one would expect of a perfectly elastic, or Hookean, solid when put under pressure or stretched. For example, whilst our skin and fat tissues eventually recover their shape after they are pinched or compressed, or after a wearable device is removed, they take time to do so. Tendons, when stretched slowly, are able to extend and then recoil back to their original size and shape, however, when rapidly extended, can further strain stiffen and eventually rupture18. Tissues are thus not purely elastic materials, like a rubber ball or a spring, because they exhibit a time-dependent mechanical response and dissipate a fraction of the energy it took to deform them, a property called viscoelasticity or poroelasticity, depending on the molecular mechanism. Macroscopically, loss of the ability to recover shape after applied mechanical stress or stretch is often a sign of injury, disease, or aging, as the affected tissues no longer recover shape after a bone break, a skin tear, or the drooping of the face after decades of gravitational stress19. However, even when tissues globally recover shape, local regions might not do so after forces are removed, experiencing irreversible or plastic deformations. Plastic deformation of the extracellular matrix is implicated in contributing to the conversion of an originally isotropic network of collagen fibers to a more aligned pattern that is often seen around tumors20–22, and irreversible changes in cell-cell boundaries caused by cell-derived forces at junction sites have recently been shown to be essential features of pattern formation during development in Drosophila23 and c-elegans24. Many soft tissues also exhibit nonlinear elasticity by strain-stiffening, or become increasingly difficult to extend as they are deformed, which may be advantageous in preventing large deformations that damage tissue25. For example, in blood vessel walls, distensibility at low strains accommodates pulsatile blood flow while increased stiffness at high strains provides elastic stability to prevent vessel rupture26. Biological tissues and ECMs thus exhibit complex, time and rate-dependent mechanical behaviors including a combination of viscoelasticity, poroelasticity, plasticity, and nonlinear elasticity (Box 1).
Box 1| Materials and mechanical concepts: linking material structure to functional responses under load.
Materials can be categorized by how they deform (or change shape) in response to mechanical loading, typically in a stress strain test. Mechanical stress is defined as the force per unit area, with units of Pascals (N/m2) and can be in shear or normal. Strain is a normalized measure of deformation. Constitutive equations describe the relationship between stress and strain for a given material. Biological tissues and ECMs can exhibit a combination of nonlinear elasticity, viscoelasticity, poroelasticity, and plasticity. Materials that are both viscoelastic and plastic are considered to be viscoplastic.

As cells interact with ECMs through dynamic processes that span a range of forces, from piconewtons up to hundreds of nanonewtons for individual cells, and span a range of timescales, from milliseconds to hours, it would be expected that time-dependent and strain-dependent mechanical responses in ECMs should impact cell-matrix interactions and mechanotransduction (Fig. 1). Indeed, an emerging body of evidence has demonstrated that these more complex mechanical characteristics of tissues and ECMs impact cells, sometimes in ways not anticipated from our previous understanding of mechanotransduction based on purely elastic substrates. Here, we review the complex mechanical behaviors of tissues and ECMs, discuss recent work elucidating the impact of ECM viscoelasticity on cells, and describe the potential for use of viscoelastic biomaterials in regenerative medicine.
Figure 1|. Mechanical interactions between cells and extracellular matrices.

Cells interact with ECMs mechanically, including by pulling, often through actomyosin-based contractility coupled to the ECM through integrin-based adhesions, and pushing, often through actin polymerization and microtubules. The mechanical properties of ECMs mediate these interactions resulting in cell mechanotransduction and impacting cell behaviors.
Tissue and ECM mechanics are complex
Viscoelasticity has been found to be a near universal characteristic of living tissues and ECMs. In response to a mechanical perturbation, viscoelastic materials exhibit an instantaneous elastic response, characteristic of purely elastic solids, followed by a time-dependent mechanical response and energy dissipation or loss, both characteristics of viscous liquids. Viscoelastic materials will ‘creep’, or deform in a time-dependent manner, in response to the application of an external step stress or load, and undergo ‘stress relaxation’, or reduce stress levels in a time-dependent manner, in response to a step deformation. Further, under an imposed sinusoidal deformation, stress and strain are completely in-phase for a purely elastic material, due to all of the inputted deformation energy being able to be ‘stored’ and ‘recovered’ during each cycle without any loss, whereas for a purely viscous fluid they are completely out-of-phase, a result of all of the inputted deformation energy being dissipated or ‘lost’ by internal friction in the system as it flows. Viscoelastic materials exhibit a response between these two extremes, with the in-phase component of the response described as the storage, or elastic, modulus and the out-of-phase response described as the loss, or viscous, modulus. The magnitude of the ratio of the loss modulus to the storage modulus in viscoelastic materials typically depends on the frequency. Viscoelastic solids are differentiated from viscoelastic fluids by maintaining stress or elastic resistance at long times under a constant deformation, or by reaching an equilibrium deformation under loading at long times. Everyday examples of viscoelastic solids include jello (gelatin), a “stress ball”, and bread dough, while silly putty serves as an example of a viscoelastic fluid. One of the softest and most dissipative viscoelastic tissues in mammals is the brain, which has been extensively studied at time scales and deformation magnitudes that span the range relevant to blasts and concussions on the fast (ms) and high stress (MPa) limit to the deformation caused by tumor growth on the slow (weeks) and low stress (10 Pa) limit. Depending on the time scale and deformation, brain tissue can dissipate at least as much energy as it stores in elastically recoverable deformation27, and at very long time scales it appears to flow like a glass or liquid28. Further, dissipation (and viscoelasticity) can resolve not only grey from white matter, but also different regions of the brain29. Other soft tissues are also viscoelastic, with rheological analysis showing that soft tissues generally exhibit loss, or viscous, moduli that are usually around 10 – 20% of their storage, or elastic, moduli at 1 Hz (Fig. 2a). Stress relaxation tests reveal that soft tissues, including liver, breast, muscle, skin, and adipose substantially relax their resistance to a deformation over timescales from tens to hundreds of seconds30–36 (Fig. 2b). Even stiffer skeletal tissues including bone, tendon, ligaments, and cartilage are viscoelastic, with loss moduli at about ~10% of the storage moduli. Embryos at various stages of development37, and regenerative structures such as fracture hematomas30 or blood clots38 also exhibit viscoelasticity.
Figure 2|. Biological tissues and extracellular matrices are viscoelastic and exhibit stress relaxation in response to a deformation.

a, Plot of loss modulus at ~1 Hz, a measure of viscosity (or dissipation), versus storage modulus at ~1 Hz, a measure of elasticity, for skeletal tissues, soft tissues, and reconstituted ECMs (rECMs). Grey dotted line indicates a loss modulus that is 10% of storage modulus. Data was taken from a set of randomly selected publications28,33,35,43,76–106. Shear storage and loss moduli were converted to storage and loss moduli by assuming a Poisson ratio of 0.5, and thus multiplying by a factor of 3. b, Stress relaxation tests on the indicated tissues. Data from refs.14,30,31,107,108. Data for a and b result from various modalities of measurement (shear, compression, tension), various measurement tools (mechanical testers, nanoindentation, AFM, shear rheometry), and tissue of different animal origins (human, rat, mouse, bovine, sheep, porcine, canine).
Importantly, changes in viscoelasticity have been associated with disease progression. Determination of elastic moduli, the basis of palpation that can identify stiff tumors, is not efficient for identifying most types of brain tumors, but rather changes in their dissipative properties, as revealed by magnetic resonance elastography, can identify the margins of gliomas and other types of brain tumor in situ39. Further, changes in brain viscoelasticity have been linked to aging40 and multiple sclerosis41. Similarly, breast cancer progression is associated with changes in both stiffness and energy dissipation42. Changes in viscoelasticity are likely to be associated with other types of cancers or other diseases, particularly those involving fibrosis or inflammation, as well as injuries, but data on these are largely missing, representing a critical gap in knowledge.
Materials that exhibit viscoplasticity represent a subset of viscoelastic materials, in that they exhibit permanent deformations when the applied stress exceeds a material ‘yield stress’ and remain at least partially deformed when the stress is removed. The response of these materials is viscoelastic to loads or deformations below their yield stress. For instance, molding clay and toothpaste are both viscoplastic. Reconstituted extracellular matrix materials used for cell culture, including common formulations of type-1 collagen gels, reconstituted basement membrane matrix, and fibrin gels, are typically viscoplastic22,43 unless they are sufficiently crosslinked covalently by enzymes such as Factor XIIIa or lysyloxidase44,45. Tissue viscoplasticity has been characterized even less than tissue viscoelasticity, representing another critical gap in knowledge.
Numerous mechanisms underlie the dissipative properties of tissues and ECMs, with some of these mechanisms also leading to viscoplasticity. Tissues consist of cells, ECM, and extracellular fluid. The ECM, composed of fibrous protein polymer networks, typically type-1 collagen fiber networks, interspersed with highly hydrated, flexible polysaccharides and other large molecules, is thought to be a key regulator of tissue mechanics and viscoelasticity46,47. Dissipation in networks of collagen or fibrin fibers depends on the nature of the bonds that link one fiber to another43,48. Most network crosslinks are non-covalent and arise from numerous weak bonds with dissociation rates fast enough to allow stresses to relax, or allow material creep, on a relevant time scale. These weak bonds can also exhibit load-dependent dynamics43, and the breaking of weak bonds under mechanical deformation or loading dissipates energy. Reformation of weak bonds following matrix deformation can stabilize the deformed state of the material, leading to plastic deformations. Using a theoretical fiber network model of collagen, a phase diagram was derived that classified the dominant mechanisms of plasticity based on the rate and magnitude of deformation and the mechanical properties of individual fibers21. It was shown that the experimentally observed viscoplasticity of collagen networks is caused by the formation of new cross-links if moderate strains are applied at small rates or due to permanent fiber elongation if large strains are applied over short periods. Both slipping of bonds between collagen fibers, and sliding of collagen fibrils, have been observed in vivo for tissues under load, for example in skin49 and tendon50, respectively. Polymer entanglements may function similarly to weak crosslinks, as release of an entanglement dissipates energy and allows the matrix to flow. These weak crosslinks or entanglement interactions co-exist with more stable covalent crosslinks, which act to diminish liquid-like flow and mechanical plasticity of the matrix overall, but do not eliminate dissipation by unbinding of the weak bonds or by deformations that can change sample volume. Elastin fibers also act to promote elastic recovery at the tissue-scale51,52. Protein unfolding is another mechanism of energy dissipation and viscoelasticity53,54, and has been reported in fibrin55, spectrin56,57 and intermediate filament58 networks in vitro. The relative importance of these distinct mechanisms of dissipation will likely vary substantially in their relevance to the viscoelastic spectrum displayed by different tissues.
Since tissues are largely water, the flow of water within the ECM can cause significant viscous dissipation and what are termed poroelastic effects, depending on the mesh size or porosity of the tissue and the rate of loading. Dissipation due to poroelasticity occurs under tension or compression, and results from volume changes due to water flow into or out of the network59. Variations in cell number or density and ECM composition, density, and conformation in a tissue, enables fluid to be differentially held by or released from the matrix when under an externally imposed load or strain, resulting in variations in response. In contrast, shear deformations change shape but not volume of the sample, and dissipation due to water movement within the matrix is much lower. As a result, the time- or frequency-dependent viscoelastic modulus measured in uniaxial strain for the ECM is much greater than it is for shear strain60. Poroelastic effects superpose with other mechanical behaviours of tissues and ECM, including nonlinear elasticity, viscoelasticity, and viscoplasticity.
Similar mechanisms apply to viscoelasticity of the cytoskeleton of cells61–63, with two important distinctions. The relatively impermeable cell membrane tends to prevent or retard poroelastic effects due to global cell deformation, but local contraction of the cytoskeleton can lead to intracellular poroelastic effects and transient pressure gradients that persist for biologically relevant times64. The second distinction is that covalent links between filaments of the cytoskeleton are very rare or non-existent. In addition, motor proteins apply random non-thermal forces to cytoskeletal filaments65, moving them faster than they would under thermal agitation alone, with the result that the active cytoskeleton is more fluidized than one without motors66. Cellular viscoelasticity can also manifest at the tissue-scale. For example, rigor mortis, the stiffening and solidification of muscle that occurs after death, happens in part because the links between actin and myosin fibers become both more numerous and permanent rather than rapidly forming and dissociating, while the living sarcomere hydrolyses ATP so that the actin-myosin links rapidly form and dissociate.
Finally, many tissues exhibit nonlinear elasticity and do not display the simple linear relationship between stress and strain that characterizes most conventional Hookean solid materials used in engineering, such as concrete, aluminum, or steel. Analogous to a nonlinear elastic material, a coiled bungee cord or rope, an exercise band, or an accordion is easy to straighten out initially, but becomes increasing difficult to stretch as it becomes fully extended. In addition to their role in mediating tissue viscoelasticity, networks of cross-linked collagen fibers are thought to govern nonlinear elasticity. For both shear and tensile deformations, collagen networks behave like linear elastic materials up to a threshold level of strain, beyond which they strain-stiffen concomitant with the alignment of fibers in the direction of maximum tensile strain25,67–71. The alignment of fibers can enable force transmission over hundreds of micrometers, facilitating long-range communication between cells70,72. A theoretical fiber network model of collagen showed that strong coupling between modes of deformation can give rise to significantly higher strain-stiffening of the networks in triaxial and biaxial tensile loading compared to uniaxial loading73. Nonlinear elasticity is also observed in cytoskeletal filament networks, including actin, vimentin, and neurofilaments, but the origins of nonlinear elasticity in these networks may have a stronger contribution of entropic elasticity, due to the semiflexible nature of the filaments25,74,75.
2D culture and the molecular clutch
The impact of substrate viscoelasticity on cells has been demonstrated powerfully through a set of 2D culture studies. In an early study, human mesenchymal stem cells (hMSCs) were cultured on collagen-coated polyacrylamide gels that had similar storage moduli, but varying loss moduli and creep responses110. Increased loss, or creep, in the substrates promoted cell spreading, focal adhesion formation, proliferation, and differentiation towards adipogenic, osteogenic, and smooth muscle cell lineages. Myosin and Rho-inhibition studies indicated the role of cytoskeletal tension in mediating the response to increased mechanical loss. In a follow-up study, increased activation of Rac1 and increases in motility and lamellipodial protrusions were found in hMSCs on substrates with higher loss and creep122. Another study compared fibroblasts and cancer cells cultured on covalently crosslinked, or elastic, versus ionically crosslinked, or viscoelastic and viscoplastic, alginate gels that presented RGD cell adhesion ligands. While cells were unable to spread on soft elastic gels, they were able to spread on soft viscoelastic gels through β1 integrin, myosin, and Rho, exhibiting robust focal adhesions and stress fibers and enhanced YAP activation, similar to their behavior on stiff and elastic substrates123. Increased spreading was associated with plastic deformation. To distinguish impacts of viscoelasticity versus viscoplasticity, viscoelastic but not viscoplastic substrates were formed using elastic polyacrylamide gels with linear acrylamide chains trapped inside111. An increased loss modulus, or faster stress relaxation, diminished fibroblast stiffness and cell spreading area, contrasting the results with viscoplastic alginate substrates. Similarly, hepatic stellate cells exhibited reduced spreading, stress fibers, and MRTF-A nuclear localization on viscoelastic compared to elastic substrates124. Interestingly, normal human hepatocytes also spread less and had lower motility on viscoelastic substrates, but hepatocellular carcinoma cells responded oppositely125.
To explain these seemingly disparate results, computational modeling has been applied. The primary sensing apparatus of substrate stiffness for cells in 2D culture is thought to be the myosin-actin-adhesion system, also known as the motor clutch module (Fig. 3), whose dynamics have successfully explained stiffness sensing of cells on elastic substrates126–128. To study the impact of ECM viscoelasticity on cell spreading, a generalized motor-clutch model that explicitly accounts for dissipative processes both in the ECM and in the cell has recently been developed129. In this model, myosin motors pull actomyosin networks at the leading edge of the cell towards the nucleus, generating actin retrograde flow. The retrograde flow is resisted by adhesion molecules that can randomly bind and unbind between actin bundles and ECM. At the cell leading edge, the polymerization of actin filaments, countered by retrograde flow, pushes the cell membrane forward, further resulting in the spreading of the cell. To account for processes that reinforce the adhesion (e.g., talin unfolding in the FA complex, which triggers recruitment of integrins130), the clutch binding rate is assumed to increase beyond a threshold level of force. Interestingly, the model shows that, for soft substrates, maximum cell spreading is achieved at an optimal level of viscosity in which the substrate relaxation time falls between the timescale for clutch binding and its characteristic binding lifetime. That is, viscosity serves to stiffen soft substrates on a timescale faster than the clutch off-rate, which enhances cell–ECM adhesion and cell spreading. On the other hand, for substrates that are stiff, the model predicts that viscosity will not influence cell spreading, since the bound clutches are saturated by the elevated stiffness. The model was tested and validated using experimental measurements on three different material systems and explained the different observed effects of viscosity on each substrate129. The clutch model has also been applied to describing myoblast interactions with purely viscous lipid bilayers131.
Figure 3|. The molecular clutch model of mechanotransduction explains the impact of matrix viscoelasticity on cell spreading in 2D.

a, Schematic of molecular clutch model of mechanotransduction as applied to viscoelastic substrates. Adapted from Ref.129. b, Molecular clutch model simulations predict optimal cell spreading when the timescale for stress relaxation is similar to the clutch binding timescale.
3D culture and mechanical confinement
The role of matrix viscoelasticity has also been investigated in 3D culture. Culture dimensionality is known to impact cell structure, adhesions, signaling, and nutrient transport132. 3D culture supports various behaviors, including epithelial morphogenesis, maintenance of pluripotency in human embryonic stem cells, and the differentiated state in chondrocytes133–135. Culture dimensionality has also been specifically implicated in mediating mechanotransduction. For example, while 2D culture studies have implicated the YAP transcriptional regulator as a universal mechanotransducer, mediating the response of cells to stiffness in all 2D culture contexts136, YAP-independent mechanotransduction is found in a 3D culture model of stiffness-induced breast cancer, which is consistent with analysis of human breast cancer patient samples137. Similarly, culture dimensionality impacts YAP/TAZ signaling in hMSCs138. YAP has been shown to play a role in mechanotransduction in some in vivo contexts, such as pancreatic cancer139, highlighting that the importance of using 3D culture models depends upon the specific biological process.
Various studies have explored the impact of matrix viscoelasticity on cells in 3D culture. Increased stress relaxation, enhanced creep, or a higher loss modulus in RGD-coupled PEG gels31, RGD-coupled alginate gels30,119, and interpenetrating networks of hyaluronic acid and collagen116 promotes spreading of adherent cells such as myoblasts, fibroblasts, and MSCs. Faster stress relaxation and increased loss also promote cell cycle progression and completion of mitosis in single cancer cells and fibroblasts, as well as osteogenic differentiation of MSCs30,107,140. Transcriptional responses are cell type specific, with human cortical progenitors and MSCs being sensitive to different ranges of stress relaxation and initial elastic moduli141. Maintenance of neural progenitor stemness is also facilitated by hydrogels with fast stress relaxation, while being inhibited in covalently crosslinked hydrogels142. In addition, chondrocytes and osteogenically differentiated MSCs can form wide volumes of interconnected cartilage-like or bone-like matrix, respectively, in viscoelastic hydrogels that exhibit fast stress relaxation30,143. Notably, viscoelastic hydrogels used in theses 3D culture studies are all viscoplastic.
Matrix viscoplasticity has been implicated in enabling mechanical remodeling of the matrix structure for cells cultured in 3D in collagen gels both locally20,22,144,145 and in microtissues146. The impact of viscoplasticity on cancer cell migration was explicitly tested in interpenetrating networks of reconstituted basement membrane matrix and alginate147. Cancer cells were found to be able to migrate through the nanoporous matrices in a protease-independent manner when the matrices exhibited sufficient mechanical plasticity. Cells mechanically opened up channels in the matrix using invadopodial protrusions, independent of proteases, and then migrated through the channels.
The impact of hydrogel viscoelasticity and viscoplasticity on cell spreading, proliferation, matrix deposition, and migration in 3D culture indicates a link to the concept of mechanical confinement. Many cellular processes involve changes in cell volume, shape, or movement (Fig. 4a). When these processes are physically restricted in 3D by the surrounding ECM or cells, the cells are considered to be mechanically confined148,149. The established view has been that pore size and matrix degradability are key regulators of mechanical confinement148. For example, in the context of cancer cell migration, it had been shown that rigid pore sizes below ~3 μm block migration, with cells unable to squeeze their stiff nucleus through smaller pores150–152. Note that PEG, alginate, and hyaluronic acid based hydrogels typically have nanometer scale pores. With rigid or elastic pores, matrix degradation was required for the cells to overcome confinement and migrate. However, given sufficient viscoelasticity or viscoplasticity, cells can overcome confinement to grow in size, deposit matrix, change their morphology as they spread or undergo mitosis, and migrate. This provides the new perspective that in addition to pore size and degradability, matrix mechanical viscoplasticity governs confinement (Fig. 4b). During cell-matrix remodeling, these properties are coupled: cell remodeling of viscoplastic matrices alters pore size147, degradation of the matrix changes its viscoelastic properties153, and changes in the matrix architecture likely impacts both viscoplasticity and degradability.
Figure 4|. Matrix viscoplasticity mediates mechanical confinement in 3D culture.

a, In confining 3D matrices, processes that involve volume change, morphological changes, or a combination of both are restricted. b, Confinement is governed by a combination of matrix pore size, matrix degradability, and matrix viscoplasticity. A sufficiently large value for any one of these properties releases confinement.
In viscoelastic and viscoplastic 3D matrices, various mechanisms of mechanotransduction have been reported. As with 2D culture, actomyosin based contractility coupled to the matrix through integrin mediated adhesions, and integrin-ligand clustering, are implicated30,154. While in principle, some of these impacts could likely be explained by molecular-clutch based models, these models have not yet been extended to 3D contexts involving mechanical confinement. Another mechanism involves cell volume expansion. Chondrocytes, MSCs, and cancer cells expand their volume, or grow as part of the cell cycle, in matrices with fast stress relaxation, but the volume expansion is restricted in matrices that exhibit slow stress relaxation, or are more elastic107,143,155. In MSCs, volume expansion activates TRPV4 stretch-activated ion channels, and the signaling cascade induced by the resulting calcium influx drives nuclear localization of RUNX2, but not YAP, to promote osteogenic differentiation in MSCs155. Similarly, growth during the G1 phase of the cell-cycle activates a TRPV4-PI3K/Akt-p27kip1 signaling axis to promote cell cycle progression in cancer cells107. Restriction of cell volume expansion promotes Il-1β signaling in chondrocytes, resulting in an osteoarthritic phenotype143. Finally, as matrix remodeling and deposition are often enhanced in matrices with increased viscoplasticity, the mechanical microenvironment to which cells respond is time-dependent, and cell-matrix interaction becomes a dynamic and potentially iterative process.
Viscoelastic biomaterials in medicine
One potentially impactful application for these findings lies in the design of biomaterials for regenerative medicine. This field originated with the goal of regenerating tissues and organs, or engineering replacements, for those damaged or lost to disease or trauma156. Biomaterials are typically utilized for cell and drug delivery, to spatially organize transplanted and resident cells, for regulation of gene expression, and to guide tissue structure and function in various regenerative, tissue and immune-engineering applications157. The demonstrated impact of matrix viscoelasticity on cell proliferation, gene expression, fate, and migration highlights it as potentially a key design parameter for biomaterials-based applications. Indeed, FDA-approved, tissue engineering products (e.g., Apligraft™ engineered skin, Infuse™ bone regeneration devices) are often based on viscoelastic matrices. Advances in materials processing techniques such as 3D printing, which often utilizes viscoelastic materials158,159, have allowed tissue and organ structure and properties to be more faithfully recapitulated. The utility of engineered tissues as improved models for basic studies of development and pathology, test beds for toxicology analysis, and improved drug screening have also led to significant interest in the development of microphysiological systems (e.g., tissue-on-chip) and cultured organoids160,161. These can more faithfully recapitulate tissue and organ biology than standard, 2D cell culture models, while also enabling the study of human biology as versus the animal biology of classic preclinical studies.
There is both direct evidence, and significant correlative data, that viscoelasticity is an important design parameter for biomaterials used in regenerative medicine. The first demonstration that matrix stiffness regulates regeneration utilized the transplantation of stem cells within viscoelastic hydrogels162. Strikingly, the impact of stiffness on stem cell fate in those gels related to the ability of cellular traction forces to remodel the polymers comprising the hydrogels154, suggesting that in fact it was the viscoelasticity of the gels that was key to their impact on cell fate in vivo. A subsequent study directly examined the impact of viscoelasticity by transplanting cells in hydrogels of matched initial elastic moduli, but varying rates of stress relaxation. Hydrogels with more rapid stress relaxation led to greater bone regeneration163; the optimal relaxation rate corresponded to that of human fracture hematomas isolated from patients163, which provide the environment in which bone regeneration naturally occurs. Similar viscoelastic hydrogels delivering inductive proteins were also found to promote extensive bone regeneration, likely due to the ability of host cells to readily invade the gels164,165. The beneficial impact of hydrogels in various applications including cartilage regeneration, vocal cord regeneration, and amelioration of pathologic remodeling of the myocardium following myocardial infarction may also relate to their viscoelastic properties166–169.
A key question is whether viscoelasticity has been a hidden variable that explains much past work in the biomaterials field more broadly. Some of the most widely used and successful biomaterials in regenerative medicine, including collagen gels, hyaluronic acid, and supramolecular assemblies170 are physically-crosslinked hydrogels (e.g., collagen and hyaluronic acid). The most widely used biomaterial for intestinal organoid formation in vitro, reconstituted basement membrane matrix, is also a physically-crosslinked viscoelastic hydrogel, as are others used to promote formation of skeletal muscle, liver, and neural organoids171–174. While there have been a number of studies aiming to delineate the impact of matrix degradation on tissue regeneration, a provocative possibility is that the impacts might, at least in part, relate to the viscoelastic behavior of these biomaterials. Several early studies concluded that more rapidly degrading hydrogels led to greater tissue regeneration than more slowly degrading gels175,176. However, those studies utilized alterations in polymer molecular weight to regulate gel dissolution, and these changes will also alter viscoelasticity. A number of studies examining 3D mechanotransduction have utilized covalently-crosslinked hydrogels and concluded that degradation of the gels was key to how cells interpreted gel cues177,178. However, the cellular activity leading to degradation of these materials will likely transition the local matrix to a more viscoelastic state. In addition, cells may be interacting with the matrix molecules they themselves deposit117, which might provide a viscoelastic substrate. Similarly, recent efforts to develop a synthetic analog to the naturally-derived, physical hydrogels for organoid formation demonstrate that gel degradability is critical to designing synthetic replacements179,180. While little is known regarding the role of viscoelasticity in the fate and functional state of cells of the innate and adaptive immune system, a recent study has implicated purely elastic covalently-crosslinked synthetic matrices, as contrasted to those fabricated with naturally derived physically-crosslinked viscoelastic extracellular matrix, as leading to inflammatory as versus regeneration-promoting immune cell responses181. Clearly, significant research will be required to delineate the specific roles of viscoelasticity, other physical properties and chemical composition in the cellular and tissue response to various biomaterials mediating tissue repair and formation.
Future outlook
Viscoelasticity is a near universal feature of living tissues and ECMs, and a rapidly expanding body of evidence is establishing that cells sense and respond to the viscoelastic properties of ECMs, challenging the current stiffness-centric view of cell-matrix mechanotransduction. There is a fundamental need for additional measurements of the viscoelasticity and viscoplasticity of tissues during development, and adult and pathologic tissues, as such measurements are currently quite limited. The change in viscoelasticity and viscoplasticity associated with diseases will be of particular interest, especially at the microscale relevant to cells. As both 2D and 3D culture studies have shown that changes in matrix viscoelasticity drive broad changes in proliferation, gene expression, migration, and differentiation, it is likely that changes in tissue viscoelasticity will play a role in disease progression and this relation could serve as a potent target for therapeutic approaches. More work is needed in the future to explore the relationships between viscoelasticity and viscoplasticity and higher order behaviours in development, tissue genesis and repair and disease aetiology.
While the impact of substrate viscoelasticity on cell spreading in 2D culture is increasingly understood, the impact of viscoelasticity must also be considered in the context of other physical cues of the matrix. Architectural features, including geometry, porosity, and topology (e.g., nanoscale roughness) have all been demonstrated to impact various aspects of cell behavior182–186. However, these have typically been studied in the context of high moduli, purely elastic matrices. It is unclear how cells will interpret these cues in the context of viscoelastic matrices. Cells generate forces and deformations on substrates in a highly dynamic manner, leading to a complex time-dependent mechanical response of the substrates, which may significantly alter the original architectural and the feature sizes to which cells respond. While externally applied stresses (e.g., compressive and shear forces) conveyed to cells from their matrices also regulate cellular gene expression and tissue structure and function187,188, their impacts have often been studied in the context of purely elastic substrates. Dissipation of externally applied forces by viscoelastic matrices is likely to diminish the magnitude and distance of action of these cues and may alter the mechanotransduction pathways they trigger.
Mechanistic understanding of mechanotransduction in viscoelastic and viscoplastic matrices in 3D is still limited. New tools and approaches that enable one to decipher cell-matrix interactions with greater spatiotemporal resolution are needed. This is particularly important in viscoplastic matrices as cell interactions with the matrix would be expected to dynamically alter local matrix architecture, ligand density, and viscoelasticity. Super-resolution imaging in 3D, molecular force sensors, and materials with dynamically tunable mechanical properties are emerging technologies that may address this need and provide a detailed readout of the dynamic molecular scale interactions and forces that occur between cells and viscoplastic matrices189–192, helping to develop a more holistic view of cell-matrix signaling. In addition, most synthetic hydrogel systems used in this field are nanoporous and do not capture the fibrillarity and ligand presentation of native ECMs.. Incorporation of collagen fibers into synthetic hydrogels116,121, or use of synthetic approaches to generating collagen-like fibers193, may help address this important limitation. Further, integrating advances in chemical synthesis routes that permit explicit control over composition, architecture and precise positioning of functional groups194,195 (e.g. RAFT, DNA origami) and real time, non-invasive tuning of properties191, with adaptive manufacturing processes that can program material composition and architecture across varying length scales159 likely will provide new material systems to explore the impacts of viscoelasticity and viscoplasticity both in vitro and in vivo. New synthetic semiflexible filament networks made from self-assembling helix-forming monomers or by electrospinning represent a novel class of materials that can more closely mimic the elastic properties of native ECM196,197 as well as incorporate energy dissipation198 and plastic deformation199. In addition, there are major gaps in our understanding of how matrix viscoelasticity impacts signaling pathways and regulation of transcription in 3D. Mechanical cues generally regulate genome architecture200, and a recent study found that matrix stiffness impacted genome accessibility in a 3D culture model of breast cancer, which mediated induction of malignancy by enhanced stiffness13. The connection between matrix viscoelasticity and cell signaling, transcription factor activation, and the epigenome is an area ripe for study.
While biomaterials design has historically operated in the dark, relative to the importance of viscoelasticity, viscoelasticity is likely to be a key technical specification in many applications moving forward (Fig. 5). Success will likely involve mimicking the mechanical characteristics of developing tissues, as this is often used as the model for regenerative strategies. The role of viscoelasticity in regulating the biology of the various cell types regulating regeneration, possibly including pluripotent stem cells, tissue resident stem and differentiated cells, and immune cells will also need to be delineated to rationally design materials to enhance tissue regeneration. Biomaterial design may also require decoupling of the local viscoelastic properties that cells sense, from the larger, tissue-scale properties required to achieve mechanical stability of the regenerating or engineered tissue. Thus, the advent of biomaterials with controlled viscoelasticity may be transformative in improving the success of biomaterials applications in regenerative medicine.
Figure 5|. Designing viscoelastic biomaterials for regenerative medicine.

a-b, Advanced imaging is utilized to detect the mechanical properties of the tissue, damaged and normal, in order to design materials with appropriate viscoelastic properties to guide the desired pattern of gene expression from interacting cells and morphogenesis. c-d, Introduction of the material, either alone or carrying various regeneration-promoting cargoes (e.g., cells) will then lead to (right panel) regeneration of the damaged tissue and reconstitution of function.
BOX 2: Biomaterials with tunable viscoelasticity.
To reproduce both the elastic and dissipative properties of tissues in simplified bioengineered materials used for cell culture, several novel approaches based on the principles of polymer physics have recently been reported. Polymers that are inert to cell binding and not susceptible to degradation by mammalian proteases are typically used, with cell-adhesion peptide motifs or protein coupling to the polymer serving as tunable design parameters. A purely elastic hydrogel involves formation of an ideal covalent polymer network, as uncrosslinked polymers and loose ends lead to energy dissipation109. In contrast, non-ideally crosslinked polymer networks, such as polyacrylamide crosslinked to just beyond the gel point, form materials with incomplete crosslinking that allow for loss and creep110. Varying the concentrations of acrylamide (monomer) and bisacrylamide (crosslinker), or inclusion of non-crosslinked linear acrylamide polymers into crosslinked polyacrylamide gel111, allows formation of a set of gels with the same storage modulus, but varying loss moduli.
Other approaches are based on hydrogel materials that are formed, at least in part, with weak (dynamic or physical) crosslinks between the polymers. Viscoelastic polyethylene glycol (PEG) hydrogels have been formed using dynamic covalent hydrazone bonds, boronate bonds, or thioester exchange31,112–114. In alginate gels, weak ionic crosslinking leads to viscoelastic gels115. Viscoelastic hyaluronic acid-based hydrogels can be formed by using hydrazone bonds or guest-host crosslinking116,117. Weak crosslinks can also be programmed into peptide-based hydrogels118. In these networks with weak bonds, viscoelasticity can be modulated independent of the initial elastic modulus by some combination of varying the following parameters: molecular weight of the constituent polymer; coupling of inert molecules to the constituent polymer as spacers; affinity of the weak bonds; ratio between weak and covalent bonds; and the total number of bonds30,116–121. Networks formed exclusively from weak crosslinks are expected to be viscoplastic, whereas single or double networks formed with a combination of covalent and weak crosslinks may or may not exhibit viscoplasticity at the bulk scale, depending on the molecular architecture.
Box 2 Figure|.

Strategies for forming hydrogels that are elastic, viscoelastic but not viscoplastic, or viscoelastic and viscoplastic.
Acknowledgements
O.C. acknowledges support from a National Institutes of Health National Cancer Institute grant (R37 CA214136), a National Science Foundation CAREER award (CMMI 1846367), and an American Cancer Society Research Scholar Grant (RSG-16-028-01). J.C-W acknowledges the support from the Australian Research Council Discovery Grants Scheme (DP190101969). PAJ acknowledges NIH awards EB017753, GM136259, and CA193417 and the Penn Materials Research Science and Engineering Center (DMR-1720530). D.J.M. acknowledges support from the NIH (R01 DE013033, U01CA214369) and the Harvard University Materials Research Science and Engineering Center (grant DMR-1420570). VBS acknowledges NIH awards R01EB017753, U01CA202177, U54CA193417 and R01CA232256 and the NSF Center for Engineering Mechanobiology (CMMI-154857).
Footnotes
Competing interests
The authors declare no competing interests.
REFERENCES
- 1.Katzberg AA Distance as a factor in the development of attraction fields between growing tissues in culture. Science 114, 431–432, doi: 10.1126/science.114.2965.431 (1951). [DOI] [PubMed] [Google Scholar]
- 2.Keese CR & Giaever I Substrate mechanics and cell spreading. Exp Cell Res 195, 528–532 (1991). [DOI] [PubMed] [Google Scholar]
- 3.Pelham RJ Jr. & Wang Y Cell locomotion and focal adhesions are regulated by substrate flexibility. Proc Natl Acad Sci U S A 94, 13661–13665, doi: 10.1073/pnas.94.25.13661 (1997). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Discher DE, Janmey P & Wang YL Tissue cells feel and respond to the stiffness of their substrate. Science 310, 1139–1143, doi: 10.1126/science.1116995 (2005). [DOI] [PubMed] [Google Scholar]
- 5.DuFort CC, Paszek MJ & Weaver VM Balancing forces: architectural control of mechanotransduction. Nat Rev Mol Cell Biol 12, 308–319, doi: 10.1038/nrm3112 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Vogel V & Sheetz M Local force and geometry sensing regulate cell functions. Nat Rev Mol Cell Biol 7, 265–275, doi: 10.1038/nrm1890 (2006). [DOI] [PubMed] [Google Scholar]
- 7.Humphrey JD, Dufresne ER & Schwartz MA Mechanotransduction and extracellular matrix homeostasis. Nat Rev Mol Cell Biol 15, 802–812, doi: 10.1038/nrm3896 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Kechagia JZ, Ivaska J & Roca-Cusachs P Integrins as biomechanical sensors of the microenvironment. Nat Rev Mol Cell Biol 20, 457–473, doi: 10.1038/s41580-019-0134-2 (2019). [DOI] [PubMed] [Google Scholar]
- 9.Janmey PA, Fletcher D & Reinhart-King CA Stiffness Sensing in Cells and Tissues. Physiol Rev, doi: 10.1152/physrev.00013.2019 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Bellin RM et al. Defining the role of syndecan-4 in mechanotransduction using surface-modification approaches. Proc Natl Acad Sci U S A 106, 22102–22107, doi: 10.1073/pnas.0902639106 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Yang C, Tibbitt MW, Basta L & Anseth KS Mechanical memory and dosing influence stem cell fate. Nat Mater 13, 645–652, doi: 10.1038/nmat3889 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Balestrini JL, Chaudhry S, Sarrazy V, Koehler A & Hinz B The mechanical memory of lung myofibroblasts. Integr Biol (Camb) 4, 410–421, doi: 10.1039/c2ib00149g (2012). [DOI] [PubMed] [Google Scholar]
- 13.Stowers RS et al. Matrix stiffness induces a tumorigenic phenotype in mammary epithelium through changes in chromatin accessibility. Nat Biomed Eng, doi: 10.1038/s41551-019-0420-5 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Levental I, Georges PC & Janmey PA Soft biological materials and their impact on cell function. Soft Matter 3, 299–306, doi: 10.1039/b610522j (2007). [DOI] [PubMed] [Google Scholar]
- 15.Swift J et al. Nuclear lamin-A scales with tissue stiffness and enhances matrix-directed differentiation. Science 341, 1240104, doi: 10.1126/science.1240104 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Wozniak MA & Chen CS Mechanotransduction in development: a growing role for contractility. Nat Rev Mol Cell Biol 10, 34–43, doi: 10.1038/nrm2592 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Jaalouk DE & Lammerding J Mechanotransduction gone awry. Nat Rev Mol Cell Biol 10, 63–73, doi: 10.1038/nrm2597 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Wang JH Mechanobiology of tendon. J Biomech 39, 1563–1582, doi: 10.1016/j.jbiomech.2005.05.011 (2006). [DOI] [PubMed] [Google Scholar]
- 19.Mazza E, Papes O, Rubin MB, Bodner SR & Binur NS Nonlinear elastic-viscoplastic constitutive equations for aging facial tissues. Biomech Model Mechanobiol 4, 178–189, doi: 10.1007/s10237-005-0074-y (2005). [DOI] [PubMed] [Google Scholar]
- 20.Malandrino A, Trepat X, Kamm RD & Mak M Dynamic filopodial forces induce accumulation, damage, and plastic remodeling of 3D extracellular matrices. PLoS Comput Biol 15, e1006684, doi: 10.1371/journal.pcbi.1006684 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Ban E et al. Mechanisms of Plastic Deformation in Collagen Networks Induced by Cellular Forces. Biophysical Journal 114, 450–461, doi: 10.1016/j.bpj.2017.11.3739 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study utilized computational modeling to show that observed plasticity of collagen networks is caused by the formation of new cross-links if moderate strains are applied at small rates or due to permanent fiber elongation if large strains are applied over short periods, matching experimental findings.
- 22.Nam S, Lee J, Brownfield DG & Chaudhuri O Viscoplasticity Enables Mechanical Remodeling of Matrix by Cells. Biophys J 111, 2296–2308, doi: 10.1016/j.bpj.2016.10.002 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Clement R, Dehapiot B, Collinet C, Lecuit T & Lenne PF Viscoelastic Dissipation Stabilizes Cell Shape Changes during Tissue Morphogenesis. Curr Biol 27, 3132–3142 e3134, doi: 10.1016/j.cub.2017.09.005 (2017). [DOI] [PubMed] [Google Scholar]
- 24.Lardennois A et al. An actin-based viscoplastic lock ensures progressive body-axis elongation. Nature 573, 266–270, doi: 10.1038/s41586-019-1509-4 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Storm C, Pastore JJ, MacKintosh FC, Lubensky TC & Janmey PA Nonlinear elasticity in biological gels. Nature 435, 191–194, doi: 10.1038/nature03521 (2005). [DOI] [PubMed] [Google Scholar]
- 26.Shadwick RE Mechanical design in arteries. J Exp Biol 202, 3305–3313 (1999). [DOI] [PubMed] [Google Scholar]
- 27.Li W, Shepherd DET & Espino DM Frequency dependent viscoelastic properties of porcine brain tissue. J Mech Behav Biomed Mater 102, 103460, doi: 10.1016/j.jmbbm.2019.103460 (2019). [DOI] [PubMed] [Google Scholar]
- 28.Bilston LE, Liu Z & Phan-Thien N Linear viscoelastic properties of bovine brain tissue in shear. Biorheology 34, 377–385, doi: 10.1016/s0006-355x(98)00022-5 (1997). [DOI] [PubMed] [Google Scholar]
- 29.Budday S, Sommer G, Holzapfel GA, Steinmann P & Kuhl E Viscoelastic parameter identification of human brain tissue. J Mech Behav Biomed Mater 74, 463–476, doi: 10.1016/j.jmbbm.2017.07.014 (2017). [DOI] [PubMed] [Google Scholar]
- 30.Chaudhuri O et al. Hydrogels with tunable stress relaxation regulate stem cell fate and activity. Nat Mater 15, 326–334, doi: 10.1038/nmat4489 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study demonstrated an approach to modulating the stress relaxation or loss modulus of alginate hydrogels independent of the initial elastic modulus, and found that increased stress relaxation promoted cell spreading, proliferation, and osteogenic differentiation of mesenchymal stem cells in 3D culture.
- 31.McKinnon DD, Domaille DW, Cha JN & Anseth KS Biophysically defined and cytocompatible covalently adaptable networks as viscoelastic 3D cell culture systems. Adv Mater 26, 865–872, doi: 10.1002/adma.201303680 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study demonstrated the use of hydrozone bonds to form viscoelastic PEG gels, and found that the viscoelastic gels enabled myoblast spreading in 3D culture.
- 32.Reihsner R & Menzel EJ Two-dimensional stress-relaxation behavior of human skin as influenced by non-enzymatic glycation and the inhibitory agent aminoguanidine. J Biomech 31, 985–993, doi: 10.1016/s0021-9290(98)00088-8 (1998). [DOI] [PubMed] [Google Scholar]
- 33.Geerligs M, Peters GW, Ackermans PA, Oomens CW & Baaijens FP Linear viscoelastic behavior of subcutaneous adipose tissue. Biorheology 45, 677–688 (2008). [PubMed] [Google Scholar]
- 34.Qiu S et al. Characterizing viscoelastic properties of breast cancer tissue in a mouse model using indentation. J Biomech 69, 81–89, doi: 10.1016/j.jbiomech.2018.01.007 (2018). [DOI] [PubMed] [Google Scholar]
- 35.Liu Z & Bilston L On the viscoelastic character of liver tissue: experiments and modelling of the linear behaviour. Biorheology 37, 191–201 (2000). [PubMed] [Google Scholar]
- 36.Perepelyuk M et al. Normal and Fibrotic Rat Livers Demonstrate Shear Strain Softening and Compression Stiffening: A Model for Soft Tissue Mechanics. PLoS One 11, e0146588, doi: 10.1371/journal.pone.0146588 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Forgacs G, Foty RA, Shafrir Y & Steinberg MS Viscoelastic properties of living embryonic tissues: a quantitative study. Biophys J 74, 2227–2234, doi: 10.1016/S0006-3495(98)77932-9 (1998). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Gersh KC, Nagaswami C & Weisel JW Fibrin network structure and clot mechanical properties are altered by incorporation of erythrocytes. Thromb Haemost 102, 1169–1175, doi: 10.1160/TH09-03-0199 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Streitberger KJ et al. High-resolution mechanical imaging of glioblastoma by multifrequency magnetic resonance elastography. PLoS One 9, e110588, doi: 10.1371/journal.pone.0110588 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Sack I et al. The impact of aging and gender on brain viscoelasticity. Neuroimage 46, 652–657, doi: 10.1016/j.neuroimage.2009.02.040 (2009). [DOI] [PubMed] [Google Scholar]
- 41.Streitberger KJ et al. Brain viscoelasticity alteration in chronic-progressive multiple sclerosis. PLoS One 7, e29888, doi: 10.1371/journal.pone.0029888 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Sinkus R et al. MR elastography of breast lesions: understanding the solid/liquid duality can improve the specificity of contrast-enhanced MR mammography. Magn Reson Med 58, 1135–1144, doi: 10.1002/mrm.21404 (2007). [DOI] [PubMed] [Google Scholar]; These studies39,42 utilized magnetic resonance elastography to analyze changes in tissue viscoelasticity during cancer, finding that there were striking differences in viscoelasticity between malignant compared to benign breast tumors, and glioblastoma compared to healthy brain parenchyma.
- 43.Nam S, Hu KH, Butte MJ & Chaudhuri O Strain-enhanced stress relaxation impacts nonlinear elasticity in collagen gels. Proc Natl Acad Sci U S A 113, 5492–5497, doi: 10.1073/pnas.1523906113 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Gerth C, Roberts WW & Ferry JD Rheology of fibrin clots. II. Linear viscoelastic behavior in shear creep. Biophys Chem 2, 208–217, doi: 10.1016/0301-4622(74)80046-3 (1974). [DOI] [PubMed] [Google Scholar]
- 45.Liu W et al. Fibrin fibers have extraordinary extensibility and elasticity. Science 313, 634, doi: 10.1126/science.1127317 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Connizzo BK & Grodzinsky AJ Multiscale Poroviscoelastic Compressive Properties of Mouse Supraspinatus Tendons Are Altered in Young and Aged Mice. J Biomech Eng 140, doi: 10.1115/1.4038745 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]; This and earlier related studies emphasize the importance of poroelastic relaxation in the design of tissues and their changes with injury, disease, and aging.
- 47.Sauer F et al. Collagen networks determine viscoelastic properties of connective tissues yet do not hinder diffusion of the aqueous solvent. Soft Matter 15, 3055–3064, doi: 10.1039/c8sm02264j (2019). [DOI] [PubMed] [Google Scholar]
- 48.Munster S et al. Strain history dependence of the nonlinear stress response of fibrin and collagen networks. Proc Natl Acad Sci U S A 110, 12197–12202, doi: 10.1073/pnas.1222787110 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Yang W et al. On the tear resistance of skin. Nat Commun 6, 6649, doi: 10.1038/ncomms7649 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Silver FH, Freeman JW & Seehra GP Collagen self-assembly and the development of tendon mechanical properties. J Biomech 36, 1529–1553, doi: 10.1016/s0021-9290(03)00135-0 (2003). [DOI] [PubMed] [Google Scholar]
- 51.Oxlund H, Manschot J & Viidik A The role of elastin in the mechanical properties of skin. J Biomech 21, 213–218, doi: 10.1016/0021-9290(88)90172-8 (1988). [DOI] [PubMed] [Google Scholar]
- 52.Vesely I The role of elastin in aortic valve mechanics. J Biomech 31, 115–123, doi: 10.1016/s0021-9290(97)00122-x (1998). [DOI] [PubMed] [Google Scholar]
- 53.DeBenedictis EP & Keten S Mechanical unfolding of alpha- and beta-helical protein motifs. Soft Matter 15, 1243–1252, doi: 10.1039/c8sm02046a (2019). [DOI] [PubMed] [Google Scholar]
- 54.Zhao XH Multi-scale multi-mechanism design of tough hydrogels: building dissipation into stretchy networks. Soft Matter 10, 672–687, doi: 10.1039/c3sm52272e (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Brown AE, Litvinov RI, Discher DE, Purohit PK & Weisel JW Multiscale mechanics of fibrin polymer: gel stretching with protein unfolding and loss of water. Science 325, 741–744, doi: 10.1126/science.1172484 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Paramore S, Ayton GS & Voth GA Extending a spectrin repeat unit. II: rupture behavior. Biophys J 90, 101–111, doi: 10.1529/biophysj.105.066977 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Takahashi H, Rico F, Chipot C & Scheuring S alpha-Helix Unwinding as Force Buffer in Spectrins. ACS Nano 12, 2719–2727, doi: 10.1021/acsnano.7b08973 (2018). [DOI] [PubMed] [Google Scholar]
- 58.Block J et al. Viscoelastic properties of vimentin originate from nonequilibrium conformational changes. Sci Adv 4, eaat1161, doi: 10.1126/sciadv.aat1161 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Oftadeh R, Connizzo BK, Nia HT, Ortiz C & Grodzinsky AJ Biological connective tissues exhibit viscoelastic and poroelastic behavior at different frequency regimes: Application to tendon and skin biophysics. Acta Biomater 70, 249–259, doi: 10.1016/j.actbio.2018.01.041 (2018). [DOI] [PubMed] [Google Scholar]
- 60.van Oosten AS et al. Uncoupling shear and uniaxial elastic moduli of semiflexible biopolymer networks: compression-softening and stretch-stiffening. Sci Rep 6, 19270, doi: 10.1038/srep19270 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Mollaeian K, Liu Y, Bi S & Ren J Atomic force microscopy study revealed velocity-dependence and nonlinearity of nanoscale poroelasticity of eukaryotic cells. J Mech Behav Biomed Mater 78, 65–73, doi: 10.1016/j.jmbbm.2017.11.001 (2018). [DOI] [PubMed] [Google Scholar]
- 62.Hu J et al. Size- and speed-dependent mechanical behavior in living mammalian cytoplasm. Proc Natl Acad Sci U S A 114, 9529–9534, doi: 10.1073/pnas.1702488114 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Mitchison TJ, Charras GT & Mahadevan L Implications of a poroelastic cytoplasm for the dynamics of animal cell shape. Semin Cell Dev Biol 19, 215–223, doi: 10.1016/j.semcdb.2008.01.008 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Moeendarbary E et al. The cytoplasm of living cells behaves as a poroelastic material. Nat Mater 12, 253–261, doi: 10.1038/nmat3517 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Guo M et al. Probing the stochastic, motor-driven properties of the cytoplasm using force spectrum microscopy. Cell 158, 822–832, doi: 10.1016/j.cell.2014.06.051 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Humphrey D, Duggan C, Saha D, Smith D & Kas J Active fluidization of polymer networks through molecular motors. Nature 416, 413–416, doi: 10.1038/416413a (2002). [DOI] [PubMed] [Google Scholar]
- 67.Vader D, Kabla A, Weitz D & Mahadevan L Strain-induced alignment in collagen gels. PLoS One 4, e5902, doi: 10.1371/journal.pone.0005902 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Hall MS et al. Fibrous nonlinear elasticity enables positive mechanical feedback between cells and ECMs. Proc Natl Acad Sci U S A 113, 14043–14048, doi: 10.1073/pnas.1613058113 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Steinwachs J et al. Three-dimensional force microscopy of cells in biopolymer networks. Nat Methods 13, 171–176, doi: 10.1038/nmeth.3685 (2016). [DOI] [PubMed] [Google Scholar]
- 70.Wang H, Abhilash AS, Chen CS, Wells RG & Shenoy VB Long-range force transmission in fibrous matrices enabled by tension-driven alignment of fibers. Biophys J 107, 2592–2603, doi: 10.1016/j.bpj.2014.09.044 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Licup AJ et al. Stress controls the mechanics of collagen networks. Proc Natl Acad Sci U S A 112, 9573–9578, doi: 10.1073/pnas.1504258112 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Han YL et al. Cell contraction induces long-ranged stress stiffening in the extracellular matrix. Proc Natl Acad Sci U S A 115, 4075–4080, doi: 10.1073/pnas.1722619115 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Ban E et al. Strong triaxial coupling and anomalous Poisson effect in collagen networks. Proc Natl Acad Sci U S A 116, 6790–6799, doi: 10.1073/pnas.1815659116 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Gardel ML et al. Elastic behavior of cross-linked and bundled actin networks. Science 304, 1301–1305, doi: 10.1126/science.1095087 (2004). [DOI] [PubMed] [Google Scholar]
- 75.Chaudhuri O, Parekh SH & Fletcher DA Reversible stress softening of actin networks. Nature 445, 295–298, doi: 10.1038/nature05459 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Shah JV & Janmey PA Strain hardening of fibrin gels and plasma clots. Rheol ActA 36, 262–268 (1997). [Google Scholar]
- 77.Chan RW Measurements of vocal fold tissue viscoelasticity: approaching the male phonatory frequency range. J Acoust Soc Am 115, 3161–3170, doi: 10.1121/1.1736272 (2004). [DOI] [PubMed] [Google Scholar]
- 78.Nasseri S, Bilston LE & Phan-Thien N Viscoelastic properties of pig kidney in shear, experimental results and modeling. Rheol ActA 41 (2002). [Google Scholar]
- 79.Hatami-Marbini H Viscoelastic shear properties of the corneal stroma. J Biomech 47, 723–728, doi: 10.1016/j.jbiomech.2013.11.019 (2014). [DOI] [PubMed] [Google Scholar]
- 80.Pereira H et al. Biomechanical and cellular segmental characterization of human meniscus: building the basis for Tissue Engineering therapies. Osteoarthritis Cartilage 22, 1271–1281, doi: 10.1016/j.joca.2014.07.001 (2014). [DOI] [PubMed] [Google Scholar]
- 81.Coluccino L et al. Anisotropy in the viscoelastic response of knee meniscus cartilage. J Appl Biomater Funct Mater 15, e77–e83, doi: 10.5301/jabfm.5000319 (2017). [DOI] [PubMed] [Google Scholar]
- 82.Chaudhuri O et al. Extracellular matrix stiffness and composition jointly regulate the induction of malignant phenotypes in mammary epithelium. Nat Mater 13, 970–978, doi: 10.1038/nmat4009 (2014). [DOI] [PubMed] [Google Scholar]
- 83.Jansen LE, Birch NP, Schiffman JD, Crosby AJ & Peyton SR Mechanics of intact bone marrow. J Mech Behav Biomed Mater 50, 299–307, doi: 10.1016/j.jmbbm.2015.06.023 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Suki B & Lutchen KR Lung Tissue Viscoelasticity. Wiley Encyclopedia of Biomedical Engineering (2006). [Google Scholar]
- 85.Holt B, Tripathi A & Morgan J Viscoelastic response of human skin to low magnitude physiologically relevant shear. J Biomech 41, 2689–2695, doi: 10.1016/j.jbiomech.2008.06.008 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Barnes SC et al. Viscoelastic properties of human bladder tumours. Journal of the Mechanical Behavior of Biomedical Materials 61, 250–257 (2016). [DOI] [PubMed] [Google Scholar]
- 87.Kiss MZ, Varghese T & Hall TJ Viscoelastic characterization of in vitro canine tissue. Phys Med Biol 49, 4207–4218, doi: 10.1088/0031-9155/49/18/002 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Klatt D et al. Viscoelastic properties of liver measured by oscillatory rheometry and multifrequency magnetic resonance elastography. Biorheology 47, 133–141, doi: 10.3233/BIR-2010-0565 (2010). [DOI] [PubMed] [Google Scholar]
- 89.Nicolle S & Palierne JF Dehydration effect on the mechanical behaviour of biological soft tissues: observations on kidney tissues. J Mech Behav Biomed Mater 3, 630–635, doi: 10.1016/j.jmbbm.2010.07.010 (2010). [DOI] [PubMed] [Google Scholar]
- 90.Nicolle S, Lounis M, Willinger R & Palierne JF Shear linear behavior of brain tissue over a large frequency range. Biorheology 42, 209–223 (2005). [PubMed] [Google Scholar]
- 91.Hrapko M, van Dommelen JA, Peters GW & Wismans JS The mechanical behaviour of brain tissue: large strain response and constitutive modelling. Biorheology 43, 623–636 (2006). [PubMed] [Google Scholar]
- 92.Netti P, D’amore A, Ronca D, Ambrosio L & Nicolais L Structure-mechanical properties relationship of natural tendons and ligaments. Journal of Materials Science: Materials in Medicine 7, 525–530 (1996). [Google Scholar]
- 93.Tanaka E et al. Dynamic shear properties of the porcine molar periodontal ligament. J Biomech 40, 1477–1483, doi: 10.1016/j.jbiomech.2006.06.022 (2007). [DOI] [PubMed] [Google Scholar]
- 94.Tanaka E et al. Comparison of dynamic shear properties of the porcine molar and incisor periodontal ligament. Ann Biomed Eng 34, 1917–1923, doi: 10.1007/s10439-006-9209-2 (2006). [DOI] [PubMed] [Google Scholar]
- 95.Troyer KL & Puttlitz CM Human cervical spine ligaments exhibit fully nonlinear viscoelastic behavior. Acta Biomater 7, 700–709, doi: 10.1016/j.actbio.2010.09.003 (2011). [DOI] [PubMed] [Google Scholar]
- 96.Fessel G & Snedeker JG Evidence against proteoglycan mediated collagen fibril load transmission and dynamic viscoelasticity in tendon. Matrix Biology 28, 503–510 (2009). [DOI] [PubMed] [Google Scholar]
- 97.Nagasawa K, Noguchi M, Ikoma K & Kubo T Static and dynamic biomechanical properties of the regenerating rabbit Achilles tendon. Clin Biomech (Bristol, Avon) 23, 832–838, doi: 10.1016/j.clinbiomech.2008.02.002 (2008). [DOI] [PubMed] [Google Scholar]
- 98.Koolstra JH, Tanaka E & Van Eijden TM Viscoelastic material model for the temporomandibular joint disc derived from dynamic shear tests or strain-relaxation tests. J Biomech 40, 2330–2334, doi: 10.1016/j.jbiomech.2006.10.019 (2007). [DOI] [PubMed] [Google Scholar]
- 99.Tanaka E et al. Shear properties of the temporomandibular joint disc in relation to compressive and shear strain. J Dent Res 83, 476–479, doi: 10.1177/154405910408300608 (2004). [DOI] [PubMed] [Google Scholar]
- 100.Tanaka E et al. Dynamic shear behavior of mandibular condylar cartilage is dependent on testing direction. J Biomech 41, 1119–1123, doi: 10.1016/j.jbiomech.2007.12.012 (2008). [DOI] [PubMed] [Google Scholar]
- 101.Toyras J, Nieminen MT, Kroger H & Jurvelin JS Bone mineral density, ultrasound velocity, and broadband attenuation predict mechanical properties of trabecular bone differently. Bone 31, 503–507, doi: 10.1016/s8756-3282(02)00843-8 (2002). [DOI] [PubMed] [Google Scholar]
- 102.Isaksson H et al. Precision of nanoindentation protocols for measurement of viscoelasticity in cortical and trabecular bone. J Biomech 43, 2410–2417, doi: 10.1016/j.jbiomech.2010.04.017 (2010). [DOI] [PubMed] [Google Scholar]
- 103.Cowin SC, Van Buskirk WC & Ashman RB in Handbook of Bioengineering (1987).
- 104.Les CM et al. Long-term ovariectomy decreases ovine compact bone viscoelasticity. J Orthop Res 23, 869–876, doi: 10.1016/j.orthres.2004.12.001 (2005). [DOI] [PubMed] [Google Scholar]
- 105.Polly BJ, Yuya PA, Akhter MP, Recker RR & Turner JA Intrinsic material properties of trabecular bone by nanoindentation testing of biopsies taken from healthy women before and after menopause. Calcif Tissue Int 90, 286–293, doi: 10.1007/s00223-012-9575-8 (2012). [DOI] [PubMed] [Google Scholar]
- 106.Abdel-Wahab AA, Alam K & Silberschmidt VV Analysis of anisotropic viscoelastoplastic properties of cortical bone tissues. J Mech Behav Biomed Mater 4, 807–820, doi: 10.1016/j.jmbbm.2010.10.001 (2011). [DOI] [PubMed] [Google Scholar]
- 107.Nam S et al. Cell cycle progression in confining microenvironments is regulated by a growth-responsive TRPV4-PI3K/Akt-p27(Kip1) signaling axis. Sci Adv 5, eaaw6171, doi: 10.1126/sciadv.aaw6171 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Purslow PP, Wess TJ & Hukins DW Collagen orientation and molecular spacing during creep and stress-relaxation in soft connective tissues. J Exp Biol 201, 135–142 (1998). [DOI] [PubMed] [Google Scholar]
- 109.Parada GA & Zhao XH Ideal reversible polymer networks. Soft Matter 14, 5186–5196, doi: 10.1039/c8sm00646f (2018). [DOI] [PubMed] [Google Scholar]
- 110.Cameron AR, Frith JE & Cooper-White JJ The influence of substrate creep on mesenchymal stem cell behaviour and phenotype. Biomaterials 32, 5979–5993, doi: 10.1016/j.biomaterials.2011.04.003 (2011). [DOI] [PubMed] [Google Scholar]; This study demonstrated an approach to modulating the loss modulus of PAM hydrogels independent of the elastic modulus, thereby creating a range of stiffness-matched substrates of varying viscoelasticity, showing that substrates that permitted increased creep under cell-generated stresses promoted increased cell spreading, proliferation, and tri-lineage differentiation of mesenchymal stem cells in 2D culture.
- 111.Charrier EE, Pogoda K, Wells RG & Janmey PA Control of cell morphology and differentiation by substrates with independently tunable elasticity and viscous dissipation. Nat Commun 9, 449, doi: 10.1038/s41467-018-02906-9 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study reported a method to produce viscoelastic solid substrates with separately tunable elastic and viscous moduli and showed that several cell types respond to viscoelastic substrates as though they were softer than purely elastic substrates of the same elastic modulus.
- 112.Tang SC et al. Adaptable Fast Relaxing Boronate-Based Hydrogels for Probing Cell-Matrix Interactions. Adv Sci 5, doi:ARTN 1800638 10.1002/advs.201800638 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Brown TE et al. Photopolymerized dynamic hydrogels with tunable viscoelastic properties through thioester exchange. Biomaterials 178, 496–503, doi: 10.1016/j.biomaterials.2018.03.060 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Marozas IA, Anseth KS & Cooper-White JJ Adaptable boronate ester hydrogels with tunable viscoelastic spectra to probe timescale dependent mechanotransduction. Biomaterials 223, 119430, doi: 10.1016/j.biomaterials.2019.119430 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Zhao XH, Huebsch N, Mooney DJ & Suo ZG Stress-relaxation behavior in gels with ionic and covalent crosslinks. J Appl Phys 107, doi:Artn 063509 10.1063/1.3343265 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Lou J, Stowers R, Nam S, Xia Y & Chaudhuri O Stress relaxing hyaluronic acid-collagen hydrogels promote cell spreading, fiber remodeling, and focal adhesion formation in 3D cell culture. Biomaterials 154, 213–222, doi: 10.1016/j.biomaterials.2017.11.004 (2018). [DOI] [PubMed] [Google Scholar]
- 117.Loebel C, Mauck RL & Burdick JA Local nascent protein deposition and remodelling guide mesenchymal stromal cell mechanosensing and fate in three-dimensional hydrogels. Nat Mater 18, 883–891, doi: 10.1038/s41563-019-0307-6 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study found that mesenchymal stem cells deposit matrix within a day of culture in proteolytically degradable covalently crosslinked or dynamically crosslinked viscoelastic hyaluronic acid hydrogels, and that the deposited proteins mediated mechanotransduction.
- 118.Dooling LJ, Buck ME, Zhang WB & Tirrell DA Programming Molecular Association and Viscoelastic Behavior in Protein Networks. Adv Mater 28, 4651–4657, doi: 10.1002/adma.201506216 (2016). [DOI] [PubMed] [Google Scholar]
- 119.Nam S, Stowers R, Lou J, Xia Y & Chaudhuri O Varying PEG density to control stress relaxation in alginate-PEG hydrogels for 3D cell culture studies. Biomaterials 200, 15–24, doi: 10.1016/j.biomaterials.2019.02.004 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Richardson BM, Wilcox DG, Randolph MA & Anseth KS Hydrazone covalent adaptable networks modulate extracellular matrix deposition for cartilage tissue engineering. Acta Biomater 83, 71–82, doi: 10.1016/j.actbio.2018.11.014 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Vining KH, Stafford A & Mooney DJ Sequential modes of crosslinking tune viscoelasticity of cell-instructive hydrogels. Biomaterials 188, 187–197, doi: 10.1016/j.biomaterials.2018.10.013 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Cameron AR, Frith JE, Gomez GA, Yap AS & Cooper-White JJ The effect of time-dependent deformation of viscoelastic hydrogels on myogenic induction and Rac1 activity in mesenchymal stem cells. Biomaterials 35, 1857–1868, doi: 10.1016/j.biomaterials.2013.11.023 (2014). [DOI] [PubMed] [Google Scholar]; This study demonstrated that increasing levels of dissipation in viscoelastic substrates matching skeletal muscle stiffness biased RhoGTPase activity to drive Rac1-mediated myogenic induction of mesenchymal stem cells in 2D culture.
- 123.Chaudhuri O et al. Substrate stress relaxation regulates cell spreading. Nat Commun 6, 6364, doi: 10.1038/ncomms7365 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Hui E, Gimeno KI, Guan G & Caliari SR Spatiotemporal Control of Viscoelasticity in Phototunable Hyaluronic Acid Hydrogels. Biomacromolecules 20, 4126–4134, doi: 10.1021/acs.biomac.9b00965 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Mandal K, Gong Z, Rylander A, Shenoy VB & Janmey PA Opposite responses of normal hepatocytes and hepatocellular carcinoma cells to substrate viscoelasticity. Biomater Sci 8, 1316–1328, doi: 10.1039/c9bm01339c (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Bangasser BL, Rosenfeld SS & Odde DJ Determinants of maximal force transmission in a motor-clutch model of cell traction in a compliant microenvironment. Biophys. J 105, 581–592, doi: 10.1016/j.bpj.2013.06.027 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Chan CE & Odde DJ Traction dynamics of filopodia on compliant substrates. Science 322, 1687–1691, doi: 10.1126/science.1163595 (2008). [DOI] [PubMed] [Google Scholar]
- 128.Bangasser BL et al. Shifting the optimal stiffness for cell migration. Nat. Commun 8, 15313, doi: 10.1038/ncomms15313 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Gong Z et al. Matching material and cellular timescales maximizes cell spreading on viscoelastic substrates. Proc Natl Acad Sci U S A 115, E2686–E2695, doi: 10.1073/pnas.1716620115 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Elosegui-Artola A et al. Mechanical regulation of a molecular clutch defines force transmission and transduction in response to matrix rigidity. Nat Cell Biol 18, 540–548, doi: 10.1038/ncb3336 (2016). [DOI] [PubMed] [Google Scholar]
- 131.Bennett M et al. Molecular clutch drives cell response to surface viscosity. Proc Natl Acad Sci U S A 115, 1192–1197, doi: 10.1073/pnas.1710653115 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study used analytical and Monte Carlo methods to Simulate the dynamics of motor clutches (i.e., focal adhesions) formed between the cell and a viscoelastic substrate and found that that intermediate viscosity maximizes cell spreading on soft substrates, while cell spreading is independent of viscosity on stiff substrates, in agreement with experiments on three different material systems.
- 132.Baker BM & Chen CS Deconstructing the third dimension: how 3D culture microenvironments alter cellular cues. J Cell Sci 125, 3015–3024, doi: 10.1242/jcs.079509 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Petersen OW, Ronnov-Jessen L, Howlett AR & Bissell MJ Interaction with basement membrane serves to rapidly distinguish growth and differentiation pattern of normal and malignant human breast epithelial cells. Proc Natl Acad Sci U S A 89, 9064–9068, doi: 10.1073/pnas.89.19.9064 (1992). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.von der Mark K, Gauss V, von der Mark H & Muller P Relationship between cell shape and type of collagen synthesised as chondrocytes lose their cartilage phenotype in culture. Nature 267, 531–532, doi: 10.1038/267531a0 (1977). [DOI] [PubMed] [Google Scholar]
- 135.Gerecht S et al. Hyaluronic acid hydrogel for controlled self-renewal and differentiation of human embryonic stem cells. Proc Natl Acad Sci U S A 104, 11298–11303, doi: 10.1073/pnas.0703723104 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Dupont S et al. Role of YAP/TAZ in mechanotransduction. Nature 474, 179–183, doi: 10.1038/nature10137 (2011). [DOI] [PubMed] [Google Scholar]
- 137.Lee JY et al. YAP-independent mechanotransduction drives breast cancer progression. Nat Commun 10, 1848, doi: 10.1038/s41467-019-09755-0 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Caliari SR, Vega SL, Kwon M, Soulas EM & Burdick JA Dimensionality and spreading influence MSC YAP/TAZ signaling in hydrogel environments. Biomaterials 103, 314–323, doi: 10.1016/j.biomaterials.2016.06.061 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Tito Panciera AC, Daniele Di Biagio, Giusy Battilana, Alessandro Gandin,, Stefano Giulitti, M. F, Silvio Bicciato, Valeria Panzetta, Sabato Fusco,, Luca Azzolin, A. T, Angelo Paolo Dei Tos, Matteo Fassan, Vincenzo Vindigni,, Franco Bassetto, A. R, Giovanna Brusatin, Michelangelo Cordenonsi and & Piccolo S Reprogramming normal cells into tumour precursors requires ECM stiffness and oncogenemediated changes of cell mechanical properties. Nature Materials, doi: 10.1038/s41563-020-0615-x (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Nam S & Chaudhuri O Mitotic cells generate protrusive extracellular forces to divide in three-dimensional microenvironments. Nature Physics 14, 621–+, doi: 10.1038/s41567-018-0092-1 (2018). [DOI] [Google Scholar]
- 141.Darnell M et al. Material microenvironmental properties couple to induce distinct transcriptional programs in mammalian stem cells. Proc Natl Acad Sci U S A 115, E8368–E8377, doi: 10.1073/pnas.1802568115 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]; This work revealed that the transcriptional responses of cells in 3D culture to stress relaxation, matrix stiffness and adhesion ligand density exhibit significant independent effects and coupling among these properties, demonstrating a significant cell type and context dependence of viscoelasticity sensing.
- 142.Madl CM et al. Maintenance of neural progenitor cell stemness in 3D hydrogels requires matrix remodelling. Nat Mater 16, 1233–1242, doi: 10.1038/nmat5020 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Lee HP, Gu L, Mooney DJ, Levenston ME & Chaudhuri O Mechanical confinement regulates cartilage matrix formation by chondrocytes. Nat Mater 16, 1243–1251, doi: 10.1038/nmat4993 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Mohammadi H, Arora PD, Simmons CA, Janmey PA & McCulloch CA Inelastic behaviour of collagen networks in cell-matrix interactions and mechanosensation. J R Soc Interface 12, 20141074, doi: 10.1098/rsif.2014.1074 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Kim J et al. Stress-induced plasticity of dynamic collagen networks. Nat Commun 8, 842, doi: 10.1038/s41467-017-01011-7 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Liu AS et al. Matrix viscoplasticity and its shielding by active mechanics in microtissue models: experiments and mathematical modeling. Sci Rep 6, 33919, doi: 10.1038/srep33919 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Wisdom KM et al. Matrix mechanical plasticity regulates cancer cell migration through confining microenvironments. Nature Communications 9:4144 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study demonstrated that mechanical plasticity in nanoporous matrices allows protease independent migration of cancer cells, with cells using invadopodial protrusions to mechanically open up micron-size channels to migrate through.
- 148.Paul CD, Mistriotis P & Konstantopoulos K Cancer cell motility: lessons from migration in confined spaces. Nat Rev Cancer 17, 131–140, doi: 10.1038/nrc.2016.123 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Caiazzo M et al. Defined three-dimensional microenvironments boost induction of pluripotency. Nat Mater 15, 344–352, doi: 10.1038/nmat4536 (2016). [DOI] [PubMed] [Google Scholar]
- 150.Sabeh F, Shimizu-Hirota R & Weiss SJ Protease-dependent versus -independent cancer cell invasion programs: three-dimensional amoeboid movement revisited. J Cell Biol 185, 11–19, doi: 10.1083/jcb.200807195 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Wolf K et al. Physical limits of cell migration: control by ECM space and nuclear deformation and tuning by proteolysis and traction force. J Cell Biol 201, 1069–1084, doi: 10.1083/jcb.201210152 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Harada T et al. Nuclear lamin stiffness is a barrier to 3D migration, but softness can limit survival. J Cell Biol 204, 669–682, doi: 10.1083/jcb.201308029 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Schultz KM, Kyburz KA & Anseth KS Measuring dynamic cell-material interactions and remodeling during 3D human mesenchymal stem cell migration in hydrogels. Proc Natl Acad Sci U S A 112, E3757–3764, doi: 10.1073/pnas.1511304112 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study examined how the viscoelastic properties of PEG hydrogels with degradable crosslinks were altered due to cellular degradation during migration of mesenchymal stem cells, reporting that the cells converted the elastic hydrogel into a viscoelastic fluid.
- 154.Huebsch N et al. Harnessing traction-mediated manipulation of the cell/matrix interface to control stem-cell fate. Nat Mater 9, 518–526, doi: 10.1038/nmat2732 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Lee HP, Stowers R & Chaudhuri O Volume expansion and TRPV4 activation regulate stem cell fate in three-dimensional microenvironments. Nat Commun 10, 529, doi: 10.1038/s41467-019-08465-x (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]; This study identified the role of cell volume expansion and activation of mechanosensitive ion channels in mediating how mesenchymal stem cells sense matrix viscoelasticity.
- 156.Langer R & Vacanti JP Tissue engineering. Science 260, 920–926, doi: 10.1126/science.8493529 (1993). [DOI] [PubMed] [Google Scholar]
- 157.Huebsch N & Mooney DJ Inspiration and application in the evolution of biomaterials. Nature 462, 426–432, doi: 10.1038/nature08601 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Grosskopf AK et al. Viscoplastic Matrix Materials for Embedded 3D Printing. ACS Appl Mater Interfaces 10, 23353–23361, doi: 10.1021/acsami.7b19818 (2018). [DOI] [PubMed] [Google Scholar]
- 159.Truby RL & Lewis JA Printing soft matter in three dimensions. Nature 540, 371–378, doi: 10.1038/nature21003 (2016). [DOI] [PubMed] [Google Scholar]
- 160.Clevers H Modeling Development and Disease with Organoids. Cell 165, 1586–1597, doi: 10.1016/j.cell.2016.05.082 (2016). [DOI] [PubMed] [Google Scholar]
- 161.Prantil-Baun R et al. Physiologically Based Pharmacokinetic and Pharmacodynamic Analysis Enabled by Microfluidically Linked Organs-on-Chips. Annu Rev Pharmacol Toxicol 58, 37–64, doi: 10.1146/annurev-pharmtox-010716-104748 (2018). [DOI] [PubMed] [Google Scholar]
- 162.Huebsch N et al. Matrix elasticity of void-forming hydrogels controls transplanted-stem-cell-mediated bone formation. Nat Mater 14, 1269–1277, doi: 10.1038/nmat4407 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Darnell M et al. Substrate Stress-Relaxation Regulates Scaffold Remodeling and Bone Formation In Vivo. Adv Healthc Mater 6, doi: 10.1002/adhm.201601185 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Kolambkar YM et al. An alginate-based hybrid system for growth factor delivery in the functional repair of large bone defects. Biomaterials 32, 65–74, doi: 10.1016/j.biomaterials.2010.08.074 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Kolambkar YM et al. Spatiotemporal delivery of bone morphogenetic protein enhances functional repair of segmental bone defects. Bone 49, 485–492, doi: 10.1016/j.bone.2011.05.010 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Lin X et al. A viscoelastic adhesive epicardial patch for treating myocardial infarction. Nat Biomed Eng 3, 632–643, doi: 10.1038/s41551-019-0380-9 (2019). [DOI] [PubMed] [Google Scholar]
- 167.Ruvinov E & Cohen S Alginate biomaterial for the treatment of myocardial infarction: Progress, translational strategies, and clinical outlook: From ocean algae to patient bedside. Adv Drug Deliv Rev 96, 54–76, doi: 10.1016/j.addr.2015.04.021 (2016). [DOI] [PubMed] [Google Scholar]
- 168.Chhetri DK & Mendelsohn AH Hyaluronic acid for the treatment of vocal fold scars. Curr Opin Otolaryngol Head Neck Surg 18, 498–502, doi: 10.1097/MOO.0b013e32833f85d1 (2010). [DOI] [PubMed] [Google Scholar]
- 169.Atala A, Kim W, Paige KT, Vacanti CA & Retik AB Endoscopic treatment of vesicoureteral reflux with a chondrocyte-alginate suspension. J Urol 152, 641–643; discussion 644, doi: 10.1016/s0022-5347(17)32671-x (1994). [DOI] [PubMed] [Google Scholar]
- 170.Boekhoven J & Stupp SI 25th anniversary article: supramolecular materials for regenerative medicine. Adv Mater 26, 1642–1659, doi: 10.1002/adma.201304606 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.Sato T & Clevers H Growing self-organizing mini-guts from a single intestinal stem cell: mechanism and applications. Science 340, 1190–1194, doi: 10.1126/science.1234852 (2013). [DOI] [PubMed] [Google Scholar]
- 172.Shansky J, Del Tatto M, Chromiak J & Vandenburgh H A simplified method for tissue engineering skeletal muscle organoids in vitro. In Vitro Cell Dev Biol Anim 33, 659–661, doi: 10.1007/s11626-997-0118-y (1997). [DOI] [PubMed] [Google Scholar]
- 173.Balikov DA, Neal EH & Lippmann ES Organotypic Neurovascular Models: Past Results and Future Directions. Trends Mol Med, doi: 10.1016/j.molmed.2019.09.010 (2019). [DOI] [PubMed] [Google Scholar]
- 174.Prior N, Inacio P & Huch M Liver organoids: from basic research to therapeutic applications. Gut 68, 2228–2237, doi: 10.1136/gutjnl-2019-319256 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Alsberg E et al. Regulating bone formation via controlled scaffold degradation. J Dent Res 82, 903–908, doi: 10.1177/154405910308201111 (2003). [DOI] [PubMed] [Google Scholar]
- 176.Simmons CA, Alsberg E, Hsiong S, Kim WJ & Mooney DJ Dual growth factor delivery and controlled scaffold degradation enhance in vivo bone formation by transplanted bone marrow stromal cells. Bone 35, 562–569, doi: 10.1016/j.bone.2004.02.027 (2004). [DOI] [PubMed] [Google Scholar]
- 177.Khetan S et al. Degradation-mediated cellular traction directs stem cell fate in covalently crosslinked three-dimensional hydrogels. Nat Mater 12, 458–465, doi: 10.1038/nmat3586 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Bryant SJ & Anseth KS Hydrogel properties influence ECM production by chondrocytes photoencapsulated in poly(ethylene glycol) hydrogels. J Biomed Mater Res 59, 63–72, doi: 10.1002/jbm.1217 (2002). [DOI] [PubMed] [Google Scholar]
- 179.Gjorevski N et al. Designer matrices for intestinal stem cell and organoid culture. Nature 539, 560–564, doi: 10.1038/nature20168 (2016). [DOI] [PubMed] [Google Scholar]
- 180.Cruz-Acuna R et al. Synthetic hydrogels for human intestinal organoid generation and colonic wound repair. Nat Cell Biol 19, 1326–1335, doi: 10.1038/ncb3632 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]; These studies179,180 demonstrated the use of synthetic covalently crosslinked hydrogels for organoid formation, identifying gel degradability as a critical design parameter.
- 181.Sadtler K et al. Divergent immune responses to synthetic and biological scaffolds. Biomaterials 192, 405–415, doi: 10.1016/j.biomaterials.2018.11.002 (2019). [DOI] [PubMed] [Google Scholar]
- 182.Ehrig S et al. Surface tension determines tissue shape and growth kinetics. Sci Adv 5, eaav9394, doi: 10.1126/sciadv.aav9394 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Petersen A et al. A biomaterial with a channel-like pore architecture induces endochondral healing of bone defects. Nat Commun 9, 4430, doi: 10.1038/s41467-018-06504-7 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Jain N & Vogel V Spatial confinement downsizes the inflammatory response of macrophages. Nat Mater 17, 1134–1144, doi: 10.1038/s41563-018-0190-6 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Reimer A et al. Scalable topographies to support proliferation and Oct4 expression by human induced pluripotent stem cells. Sci Rep 6, 18948, doi: 10.1038/srep18948 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Han P, F. J, Gomez GA, Yap AS, O’Neill GM, Cooper-White JJ. Five Piconewtons: The Difference between Osteogenic and Adipogenic Fate Choice in Human Mesenchymal Stem Cells. ACS Nano 13, 11129–11143, doi: 10.1021/acsnano.9b03914 (2019). [DOI] [PubMed] [Google Scholar]
- 187.Vining KH & Mooney DJ Mechanical forces direct stem cell behaviour in development and regeneration. Nat Rev Mol Cell Biol 18, 728–742, doi: 10.1038/nrm.2017.108 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Panciera T, Azzolin L, Cordenonsi M & Piccolo S Mechanobiology of YAP and TAZ in physiology and disease. Nat Rev Mol Cell Biol 18, 758–770, doi: 10.1038/nrm.2017.87 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Chen BC et al. Lattice light-sheet microscopy: imaging molecules to embryos at high spatiotemporal resolution. Science 346, 1257998, doi: 10.1126/science.1257998 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Grashoff C et al. Measuring mechanical tension across vinculin reveals regulation of focal adhesion dynamics. Nature 466, 263–266, doi: 10.1038/nature09198 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Rosales AM & Anseth KS The design of reversible hydrogels to capture extracellular matrix dynamics. Nat Rev Mater 1, doi: 10.1038/natrevmats.2015.12 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Liu AP, Chaudhuri O & Parekh SH New advances in probing cell-extracellular matrix interactions. Integr Biol (Camb) 9, 383–405, doi: 10.1039/c6ib00251j (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Baker BM et al. Cell-mediated fibre recruitment drives extracellular matrix mechanosensing in engineered fibrillar microenvironments. Nat Mater 14, 1262–1268, doi: 10.1038/nmat4444 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Braunecker WA & Matyjaszewski K Controlled/living radical polymerization: Features, developments, and perspectives. Progress in Polymer Science 32, 93–146 (2007). [Google Scholar]
- 195.Ong LL et al. Programmable self-assembly of three-dimensional nanostructures from 10,000 unique components. Nature 552, 72–77, doi: 10.1038/nature24648 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Liu KZ, Mihaila SM, Rowan A, Oosterwijk E & Kouwer PHJ Synthetic Extracellular Matrices with Nonlinear Elasticity Regulate Cellular Organization. Biomacromolecules 20, 826–834, doi: 10.1021/acs.biomac.8b01445 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Wang YM et al. Biomimetic Strain-Stiffening Self-Assembled Hydrogels. Angewandte Chemie-International Edition, doi: 10.1002/anie.201911364 (2020). [DOI] [PubMed] [Google Scholar]
- 198.Wang YF et al. Architected lattices with adaptive energy absorption. Extreme Mechanics Letters 33, doi: 10.1016/j.eml.2019.100557 (2019). [DOI] [Google Scholar]
- 199.Davidson MD et al. Mechanochemical Adhesion and Plasticity in Multifiber Hydrogel Networks. Advanced Materials 32, doi: 10.1002/adma.201905719 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Shivashankar GV Mechanical regulation of genome architecture and cell-fate decisions. Curr Opin Cell Biol 56, 115–121, doi: 10.1016/j.ceb.2018.12.001 (2019). [DOI] [PubMed] [Google Scholar]
