Review

Mechanobiomaterials: Harnessing mechanobiology principles for tissue repair and regeneration

  • Xiao Lin a, 1 ,
  • Hua Yang b, 1 ,
  • Yi Xia c ,
  • Kang Wu a ,
  • Fengcheng Chu a ,
  • Huan Zhou c ,
  • Huajian Gao , d, ** ,
  • Lei Yang , a, c, *
Expand
  • a Orthopedic Institute, Department of Orthopedics, The First Affiliated Hospital, Soochow University, Suzhou, China
  • b School of Mechanical Engineering, Hebei University of Technology, Tianjin, China
  • c Center for Health Sciences and Engineering (CHSE), Hebei Key Laboratory of Biomaterials and Smart Theranostics, School of Health Sciences and Biomedical Engineering, Hebei University of Technology, Tianjin, China
  • d Mechano-X Institute, Applied Mechanics Laboratory, Department of Engineering Mechanics, Tsinghua University, Beijing, China
* E-mail addresses: (L. Yang),
** E-mail addresses: (H. Gao).

1These authors contributed equally to this work.

Received date: 2023-12-22

  Revised date: 2024-04-23

  Accepted date: 2024-05-12

  Online published: 2024-05-16

Abstract

Mechanical stimuli are known to play critical roles in mediating tissue repair and regeneration. Recently, this knowledge has led to a paradigm shift toward proactive programming of biological functionalities of biomaterials by leveraging mechanics-geometry-biofunction relationships, which are beginning to shape the newly emerging field of mechanobiomaterials. To profile this emerging field, this article aims to elucidate the fundamental principles in modulating biological responses with material-tissue mechanical interactions, illustrate recent findings on the relationships between material properties and biological responses, discuss the importance of mathematical/physical models and numerical simulations in optimizing material properties and geometry, and outline design strategies for mechanobiomaterials and their potential for tissue repair and regeneration. Given that the field of mechanobiomaterials is still in its infancy, this article also discusses open questions and challenges that need to be addressed.

Cite this article

Xiao Lin , Hua Yang , Yi Xia , Kang Wu , Fengcheng Chu , Huan Zhou , Huajian Gao , Lei Yang . Mechanobiomaterials: Harnessing mechanobiology principles for tissue repair and regeneration[J]. Mechanobiology in Medicine, 2024 , 2(3) : 100079 -19 . DOI: 10.1016/j.mbm.2024.100079

1. Introduction

The repair and functional restoration of tissue lesions (e.g., cartilage defects, intervertebral disc degeneration, volumetric muscle loss, and myocardial infarction) remain a great challenge for clinicians, biologists, and biomedical engineers. Compared to chemical drugs and other therapeutic agents, various implantable and wearable biomedical devices have already been playing equal, if not more important, roles in mediating tissue repair and functional restoration. The physiochemical characteristics of biomedical devices, especially non-active devices, are largely rooted in their constituent materials and subsequently affect the biological responses of the human body. To date, the influence of the biochemical cues of materials on cellular responses has been widely investigated, providing a foundation for the rational design of a material's chemical characteristics for tissue repair and regeneration. In addition to biochemical cues, the mechanical interaction between biomedical devices and tissues is ubiquitous. However, traditional mechanical factors have only been considered for the development of load-bearing biomedical devices to guarantee their reliability in the body's loading environment.
From a mechanics perspective, human and animal tissues have specialized structures and mechanical properties that enable an appropriate response or tolerance to dynamic loadings during physiological activities. In addition, the extracellular matrix (ECM), with varying composition and architecture, transduces mechanical forces at the tissue level to the mechanical stimuli experienced by cells. To accommodate the dynamic mechanical environment, cells sense extracellular mechanical stimuli and regulate their metabolic activities accordingly.1 However, tissue damage or degeneration can dramatically change the mechanical stimuli perceived by cells. Trauma- or aging-induced pathological changes in tissue mechanical (including matrix degradation, loss of tissue hydration, and production of incorrect fibrous ECM) and structural properties (including changes in tissue shape, size, and microarchitecture) result in substantial changes in stress and strain distribution across tissues. At the cellular/subcellular scale, the degradation of the ECM affects the mechanical interaction between the ECM and cells and consequently alters the mechanical stimuli experienced by cells in a particular zone. Therefore, traumatic damage and/or pathological changes in tissues at different length scales can result in abnormal mechanical stress and strain, even under normal physiological loads, causing dysregulation of mechanobiological signaling of cells, compromising tissue reparative capacity, and/or even damaging adjacent healthy tissues. For example, an altered subchondral bone structure after osteoarthritis (OA) aberrantly changes the patterns of mechanical stress within the articular cartilage (AC), which subsequently disturbs chondrocyte metabolism and AC homeostasis, causing degeneration of the AC and continuous progression of OA pathology.2
Correction of the aberrant cellular mechanical environment and rebuilding tissue mechano-adaptive processes can potentially prevent or reverse tissue pathogenesis. Meanwhile, a number of important mechanobiological signaling pathways and molecular mechanisms related to tissue regeneration have been identified,3,4 allowing new strategies to boost tissue regeneration and repair.5,6 Recently, an improved understanding of cellular mechanobiological behavior and its critical role in mediating tissue growth and homeostasis3,7,8 has led to advances in the rational design and development of biomaterials to regulate the responses of the surrounding tissue/cells through mechanical interactions.9 Over the past decade, many studies have reported enhanced tissue repair and regeneration by modulating the mechanical properties and geometry of biomaterials.10,11,12
Mechanobiomaterials is an emerging field of research in which biomaterials are proactively designed or programed with target mechanical properties, responses or behaviors for precisely mediating biomechanical environment of living systems, in order to regenerate or restore partially or totally any tissue, organ or to function of the body. The major objective of mechanobiomaterials is to design biomaterials that can dynamically and consistently adapt and respond to the changing mechanical environment of tissues and cells to enhance their regenerative potential through appropriate mechanical stimuli.9,13 This is a fast-emerging field at the interface of biology, biomechanics, materials science, and engineering, aimed at developing mechanically responsive or active materials that can interact with cells or tissues and thus proactively alter or program their mechanobiological responses. A large amount of research in this field has focused on modulating or programming cell behavior (e.g., proliferation, migration, protein secretion, and differentiation) via biomaterial-induced mechanical cues to enhance tissue repair and restoration. Modulating the mechanical properties and geometries of materials has already demonstrated great potential for repairing heart,10 intervertebral disc,11 cartilage,14 muscle15 and skin16 injuries as well as opening up new applications such as skin appearance improvement,17 orodental reconstruction,18 and intestinal circumferential expansion.19
Numerous studies have investigated development strategies for mechanobiomaterials and their underlying biological and biomechanical mechanisms. The present article provides an overview of this emerging area of research, summarizing the fundamentals of material-tissue mechanical interactions at multiple scales, the relationships between material mechanics and biological responses, and the principles or strategies for designing mechanobiomaterials for tissue repair and regeneration. Hopefully, this study will foster future research in this area and bolster the evolution of next-generation biomaterials and biomedical devices.

2. Biomechanical and mechanobiological insights into biomaterial design

Correction of pathological or stimulation of reparative mechanobiological pathways through the regulation of cellular mechanical environment at lesion sites could enhance tissue regeneration during the clinical treatment of tissue lesions. Interactions between biomaterials and tissues/cells occur at multiple length scales, and wearable and implantable materials can deliver mechanical stimuli (e.g., stretch, compression, and shear) to cells either in contact with biomaterials or within tissues, eliciting biological responses at the molecular, cellular, and tissue levels through mechanobiological mechanisms.20,21,22 Compared with chemical or biological stimuli, mechanical stimuli provide a relatively predictable platform for the precise control and modulation of cellular responses. Understanding how biomaterials induce mechanical stimuli and how biological systems respond to such stimuli are the two pillars for the rational design of mechanobiomaterials.

2.1. Mechanical interactions between biomaterials and tissues/cells

The mechanical interplay between biomaterials and tissues/cells involves multi-scale interactions (Fig. 1). At the cellular or subcellular scale, living cells constantly push or pull the surrounding ECM. When in contact with cells, biomaterials can trigger mechanobiological responses through cell-probing mechanisms, and their mechanical properties can affect the mechanical stimuli perceived by cells and subsequently dictate cellular function and fate.21 It should be mentioned that the topographical features of biomaterials play a crucial role in governing cellular behaviors by affecting the mechanical interplay between cells and biomaterials.23,24 At the tissue or organ scale, endogenous mechanical forces are constantly present throughout the tissue or organ in various forms (e.g., heartbeat, lung expansion, and joint/muscle compression). The presence of biomaterials may change the dynamic mechanical environment of the tissue (e.g., dynamic distribution of stress and strain), and thus alter the mechanical stimuli perceived by cells. For example, the stress and strain distributions within biomaterials and local tissues can be significantly altered, depending on the mechanical properties and geometries of the biomaterials. In addition, responsive biomaterials that change their mechanical properties and geometry under external stimuli (e.g., electric or magnetic force, vibration, and light) can directly exert mechanical stimuli on tissues and cells.
Fig. 1. Mechanisms involved in mechanical interactions between biomaterials and tissues/cells.
The mechanical environment of tissue-resident endogenous cells can be regulated based on an understanding of material-tissue mechanical interactions at multiple scales. Given that exogenous cells are usually incorporated into biomaterials for applications, such as cell-based therapy, mechanical stimuli to these cells can also be controlled by material-tissue mechanical interactions. Mechanical stimuli can be perceived by cells and can affect cell fate through various mechanisms, as discussed in the next section.

2.2. Transduction of mechanical signals at cellular level

Cells perceive and respond to extracellular and intracellular mechanical stimuli, which is crucial for tissue development, homeostasis, and regeneration.25 Extracellular mechanical stimuli (such as stretching, compression, and tissue shearing) (Fig. 2)26 and intracellular stimuli (such as osmotic pressure and cytoskeletal forces generated by the cytoskeleton interacting with the plasma membrane and intracellular organelles)27,28 can arise in static, incremental, and/or cyclical forms. Cells within tissues are subject to a variety of forces, to which they can either adapt or react by altering their shapes. Experimental setups can be devised to replicate these conditions and investigate how cells react to applied forces under controlled circumstances. Among the most prevalent types of physical forces exerted on living cells are pressure, stretching, and fluid shear. The specific types of forces experienced by cells vary depending on the methods and substrates used to apply the forces.29
Fig. 2. Typical modes of mechanical stimuli experienced by cells. External forces result in compressive, stretch (tensile), and shear stresses on cells.26
Cell stretching devices typically utilize elastic membranes as substrate materials. These devices stretch cells adhering to the membrane either by directly stretching the elastic basement membrane itself or by inducing deformation through methods such as liquid or vacuum. Cell stretch devices concentrate on studying the impacts of stretch direction, frequency, intensity, and duration. Various stretching modes such as uniaxial, biaxial, and equiaxial can be applied to cells within these devices. Cell compression system is designed to exert mechanical pressure on living cells. For example, the uniform compression method involves placing a glass cylinder with uniform and steady compression on a slab of hydrogel or plate over confluent cell layers in large quantities.30 Shear stress could be exerted through a microfluidic device capable of incrementally adjusting magnitude across substrates with varying stiffness levels.31 The mechanical characterization of cells necessitates measurement techniques capable of precisely applying and monitoring deformations at the micro-to nanometer scale and forces in the nano-to piconewton range, given the cells' small size and low elastic modulus. Additionally, approaches that assess local cellular mechanical properties could exert extracellular dynamic forces in the nano-to piconewton range across the entire cell, known as force-application techniques, including optical tweezers (laser traps), atomic force microscopy (AFM), acoustic force spectroscopy (AFS), cell stretchers, flow rheometry and so on.32
Mechanobiological transduction is the ability of living cells to respond to numerous mechanical stimuli and convert them into intracellular signals to direct cell behaviors (cell movement, secretion, endocytosis, proliferation, differentiation, apoptosis, etc.).33,34 Tension transduction is a typical mechanism involved in cell perception and transduction of mechanical stimuli, and is mainly transduced by the cell membrane and actin cytoskeleton. Another typical mechanism is mediated by mechanosensitive ion channels, which can be mechanically activated to allow ions to selectively traverse the cell membrane.35
Tensile stress mainly includes membrane tension due to osmotic pressure differences across the lipid bilayer and cytoskeleton tension originating from actin attached to the plasma membrane.36 Cells control membrane tension by modulating membrane area and cytoskeletal attachment,37 as shown by the membrane area regulation/cytoskeleton model of cell spreading and polarization.38 Fig. 3(a) shows typical phases during the cell spreading and polarization processes. Rounded cells exhibit weak cell-substrate attachment and high membrane tension. Stable adhesion and global remodeling of the cytoskeleton decrease membrane tension. The protrusion and shrinking of lamellipodia on opposite sides of the cell increase the membrane tension in the polarized phase.
Fig. 3. Typical mechanisms involved in cell reception of mechanical signals. (a) Model of cell spreading and polarization. Phase 1 involves rounded up cells with weak cell-substrate attachment but higher membrane tension. In phase 2, the cells begin to spread by unfolding membrane reservoirs and increase the membrane tension. Stable adhesions start to mature during phase 3, global remodeling of the cytoskeleton occurs with formation of actin bundles and the membrane tension is now at its low resting level with a constant membrane area. In the final polarized phase, lamellipodia protrude on one side of the cell, whereas the other side of the cell shrinks and membrane tension increases.38,59,60 (b) Schematic diagram of ion channel activation by mechanical force. Schematic diagram of Piezo1 channel activation by mechanical force.61
In the basal equilibrium stress state, integrins are in equilibrium with the actin filaments (F-actin) and ECM. Mechanical stimulation of cells alters the stress equilibrium instantaneously, leading to remodeling of the cytoskeleton, which in turn influences the transduction of forces throughout the cell. For example, cells polymerize cytoskeletal proteins to create stress cables and induce actomyosin contraction or self-loading to match applied loads.39,40 Variations in cytoskeleton length, density, alignment, and cross-linking, together with their kinetics and dynamics, directly regulate the mechanical properties of cells, such as contractility, rigidity, motility, and various cellular responses.41,42,43 Meanwhile, there is a dose-dependent relationship between the degree of externally applied force and cytoskeletal stiffness, and the interconnected structural tensile elements globally re-orientate in response to local stress in a manner consistent with the tensegrity model.44,45
Mechanosensitive ion channels are force sensors that translate mechanical inputs into electrical and/or biochemical intracellular signals.46 When mechanical stimulation reaches the cell membrane, stress is distributed to all components, including the double layer, cytoskeleton, and ECM, which pulls the ion channels and transforms them from a closed state to an open state (Fig. 3(b)), allowing specific ions to flow based on channel selectivity.47,48,49,50 Mechanosensitive ion channels are known to sense various mechanical signals, including radial pressure, membrane stretching, compression, shear stress, matrix stiffness, ultrasound, matrix nano-topology, and osmotic pressure,51,52 and their opening probability increases with the strength of mechanical stimuli.53 The activation of ion channels by different mechanical stimuli mediates diverse cell and tissue functions46,54,55 and pathological processes.56,57,58
In addition to the aforementioned mechanisms, cells can detect mechanical stimuli via growth factor- and primary cilia-mediated mechanisms.62 Mechanical stimuli induce changes in mechanical receptors on the cell membrane and result in a biochemical cascade, often referred to as a signaling pathway,63 leading to the modulation of gene expression in the nucleus. Focal adhesions (FAs) with mechanosensitive qualities are able to convey mechanical forces between the extracellular matrix and the cytoskeletal contractile machinery.64 Mature FAs comprise multiple molecular layers. The layer adjacent to the membrane is known as the integrin signaling layer, housing essential FA signaling proteins like focal adhesion kinase (FAK) and its adapter protein paxillin.65
As shown in Fig. 4, FAK pathway is a model of many avenues through which mechanical signals can be used to regulate adaptive responses.66 FAK can be activated by either integrins or growth factors, and tyrosine phosphorylation of FAK can trigger the assembly and disassembly of focal adhesions and cadherin-based ECM.67 FAK assumes a pivotal role in modulating the FA assembly and disassembly, thereby governing the directional mobility of cells.68 Activation of FAK undergoes autophosphorylation of Tyr-397 (FAK Y397), creating a binding site for the Src homology 2 (SH2) domain of Src and related kinases. This initiates a cascade where Src phosphorylates FAK at numerous other tyrosine residues, serving as docking points for various signaling molecules like Grb2, p130cas, and phosphatidylinositol 3-kinase, ultimately activating extracellular signal-regulated kinases (Erks). Rigid substrates foster heightened and prolonged integrin clustering, with increased clustering density enhancing FAK phosphorylation in a linear fashion. This augmented integrin clustering density prolongs FAK Y397 phosphorylation by facilitating FAK re-binding to integrin via talin, thereby extending the reaction duration.69 Mechanical activation by the compression of Piezo channels induces calcium influx, which stimulates downstream signaling pathways, such as Src and extracellular signal-regulated protein kinase (ERK), finally leading to the formation of actin-based protrusions. Shear flow, substrate stiffness, and enlarged focal adhesions lead to YAP activation in the cytoplasm and its translocation to the nucleus, where YAP triggers gene transcription, resulting in increased cell proliferation and other functions.70,71 Various mechanosensors and signaling pathways may play synergistic roles in ensuring a robust mechanosignalling program.
Fig. 4. Representative schematic showing several typical mechanisms involved in cell mechanotransduction.
The mechano-adaptive changes in morphology of cells can result in the physical extension or contraction of nuclear pores, influencing the transport of molecules between the cytoplasm and nucleus, and subsequently modulating gene expression levels.72,73 Microtubules are recognized as essential regulators of nuclear morphology, which in turn modulates transcriptional activity through chromatin condensation and the expression of downstream genes involved in cellular mechanics.74
This biomechanical and mechanobiological understanding suggests that by optimizing the mechanical properties of biomaterials, it is possible to construct an appropriate mechanical microenvironment for tissue-resident or exogenous cells to regulate their behavior, prevent pathological remodeling, and promote tissue regeneration (Fig. 5). Numerous studies have demonstrated that the mechanical properties of biomaterials and their (stimuli-responsive) programming strategies regulate the mechanical signals perceived by cells that adhere to the biomaterials or within local tissues. The effects of intrinsic and (stimuli-responsive) dynamic mechanical properties on cellular and tissue responses are discussed in the following section to better illustrate the concept and rational design of mechanobiomaterials.
Fig. 5. Illustration of typical mechanotransduction mechanisms steering material- and tissue-resident cell behaviors. IVD: intervertebral disc.

3. Effect of material mechanics on cellular and tissue responses

Numerous studies have revealed the influence of extracellular mechanics on the phenotype and function of various cells (Table 1). In addition, the mechanical properties of implants have been found to regulate adjacent tissue responses.
Table 1. Cellular processes that are potentially affected by matrix mechanical properties.
Stiffness
Matrix Biological processes Ref.
High-density collagen-I gels Maintenance, proliferation, stratification, and survival of limbal epithelial stem cells 105
Polyethylene glycol hydrogel Stemness and proliferative efficiency of skeletal muscle stem cell 107
Alginate-RGD hydrogels Matrix remodelling of bone marrow mesenchymal stem cells 106
Hybrid gold nanoparticle-hyaluronic acid hydrogel Electrical and contractile performance of human induced pluripotent stem cells 108
Trilayer extracellular matrix-based microribbon scaffold Deposition of cartilage-like collagen and synthesis of sGAG of human mesenchymal stem cells 109
Beta-sheet rich silk nanofiber and amorphous silk nanofiber composite hydrogel Chondrogenic-osteogenic gradient differentiation of bone mesenchymal stem cells 189
Proteolytically degradable alginate hydrogels Migration and invasion of human mesenchymal stem cells 111
Alginate microgel In vivo residence time of clonally derived mouse marrow stromal cells after intravenous injection 112
Four-armed poly(ethylene glycol) hydrogel Organoids generation of human pluripotent stem cell 116
Gelatin-hydroxyphenylpropionic acid hydrogel sGAG production and gene expression of chondrocyte 103
γ-PGA-SH/OHA-GMA hydrogel Migration and infiltration of fibroblasts 92
Hyaluronic acid-based microrods Reprograming of myofibroblast 95
Alginate hydrogel Osteogenic differentiation of mesenchymal stem cells 110
Biphasic CAN-PAC hydrogel Viability of cartilage cells and osteoblast cells 96
Viscoelasticity/dynamic network
Matrix Biological processes Ref.
Hyaluronate-alginate hybrid hydrogel Chondrogenic differentiation of mouse chondrocytes (ATDC5 cells) 190
Benzimidazole-based catalyst enhanced hyaluronic acid hydrogel Long-term survival and adhesion of human umbilical vein endothelial cells (HUVEC) 148
Supramolecular gelatin hydrogels Spreading, tension of cytoskeletal and osteogenesis of human mesenchymal stem cells 147
PEG-gellan gum/PEGDA double network hydrogel Proliferation, spreading and chondrogenic differentiation of bone mesenchymal stem cells 157
DEX-UPy dynamic hydrogel Osteogenic differentiation of bone mesenchymal stem cells and chondrogenic expression of chondrocyte 155
Peptide supragel Embedding and spheroid harvesting of cancer cells (MCF-7 and 4T1) 161
PPy-based dynamic conductive hydrogel Differentiation towards astrocytes of neural stem cells 164
Gelatin methacryloyl -fibrinogen interpenetrating network hydrogel 3D myoblast alignment and elongation of C2C12 cells 165
Collagen, alginate, and PEDOT:PSS biohybrid hydrogel Physiological beating rate of human-induced pluripotent stem cell-derived cardiomyocytes 166
3D hydrogel structure based on reversible receptor-ligand interaction between the glycopeptide antibiotic vancomycin and dipeptide D-Ala-D-Ala Adhesion and morphology of L929 cells and HUVECs 167
Alginate hydrogel Osteogenic differentiation and migration of hMSCs 169
GelMA hydrogels Spreading, proliferation, osteogenesis and chondrogenesis of bone mesenchymal stem cells 14
PVA/Glycerol hydrogel Vitality of nucleus pulposus cells 11
Strain-stiffening effect
Matrix Biological processes Ref.
Polyisocyanides gel KCa3.1 channel expression of HepG2 cells 130
Viscosity
Matrix Biological processes Ref.
Polydimethylsiloxane-based substrate Collective movement of epithelial cells 181
Hyaluronic acid hydrogel Cell spreading of human bone marrow-derived mscs 182
Poly(ε-caprolactone)-based substrate Spread area and formation of spheroids of NIH/3T3 fibroblast 183
Supported lipid bilayers Differentiation of C2C12 mouse myoblasts 184
Gelatin solution Adipogenic and osteogenic of human bone marrow-derived mesenchymal stem cells 186
ECM production and expression of collagen type II/aggrecan of chondrocytes 185
Chondrogenic differentiation of human mesenchymal stem cells 187
High-viscosity storage solution (EAS-1587) Storage of red blood cell and inflammatory cytokines expression of human mesenchymal stem cells 188
Numerous mathematical or physical models have been developed to understand and describe the mechanical interactions between biomaterials and biological systems at the tissue, cellular, and subcellular levels, as discussed in several review papers.75,76 These models have helped understand the role of mechanical cues in tissue degeneration for the rational design of mechanobiomaterials.
Among various mechanical properties of materials (Fig. 6), some are related to rate- or time-independent responses to load (not changing with time or strain rate, including stiffness, strain-stiffening behavior, etc.), while others elicit time-dependent responses (changing with time or strain rate, including viscoelasticity, poroelasticity, viscosity, etc.).35,77,78,79,80,81 In this section, the effects of these mechanical properties on cell and tissue responses are summarized, and their biological implications are discussed.
Fig. 6. Schematics of typical constitutive models for describing mechanical properties of biomaterials. (a) The stress-strain curve for a hyperelastic (Neo-Hookean) model. (b) The stress-strain curve of the polyisocyanopeptide hydrogel in a stress ramp. (c) Viscoelastic creep and stress relaxation under an instantaneous and constant applied stress (σ0) and strain (ε0).136 (d) Poroelastic behaviors of biopolymer networks under compression and shear tests.173 (e) Newton's law of viscosity.

3.1. Time-independent mechanical properties

3.1.1. Stiffness/elasticity

When a constant stress or strain is applied to elastic materials, they respond with a strain or stress that also remains constant over time, storing mechanical energy until the force or deformation is released, while the material fully returns to its original configuration. Stiffness is a key parameter describing the response of an elastic material to a load. With a linear elastic isotropic model, the stiffness is usually defined by Young's modulus $=\sigma / \varepsilon$, where $\sigma=F / A$, $\varepsilon=\Delta L / L$, σ is stress, ε is strain, F is force, A is area, and ΔL is change in length. Constitutive modelling of biomaterial stiffness can be challenging because both the nonlinear and anisotropic nature must be properly handled, whereas a linear elastic isotropic model is not always valid. Often, a hyperelastic isotropic incompressible material with a neo-Hookean material model is used to describe the mechanical behavior of biomaterials.82 Usually, the strain energy density of a neo-Hookean material is defined as
$W=C_{10}\left(I_{1}-3\right)$
where I1 is the first deviatoric strain invariant and C10 is a material parameter. A stress-strain curve of a neo-Hookean model material is schematically presented in Fig. 6(a). Models of more complex biomaterials that consider anisotropic properties are increasingly being developed.83,84,85,86,87,88
Cell behaviors, including cell spreading, migration, ECM deposition, proliferation, stem cell lineage commitment, and self-renewal of a variety of cell types are affected by the stiffness of the substrate or material to which they attach (Table 1).78,79,80 Here, the effects of biomaterial stiffness on immune responses, the efficacy of cell therapy, and the activity of tissue-forming cells are briefly discussed for guiding biomaterial design.
Biomaterial stiffness influences the behavior of endogenous cells.89 First, the stiffness of the substrate determines cell migration and tissue integration in synthetic scaffolds.90,91,92,93,94 Secondly, substrate stiffness affects the phenotypic reprogramming of endogenous cells.95,96 Third, the inflammatory responses of innate immune cells, which play a central role in regulating tissue injury repair, are also influenced by the stiffness of the biomaterial, as seen in in vitro97,98 and in vivo studies.93,99 However, the correlation between matrix stiffness and immune responses remains to be fully elucidated as mixed results have been reported. Either increased98,100 or decreased101,102 pro-inflammatory macrophage phenotypes with increasing substrate stiffness have been observed in different studies.
Biomaterial stiffness affects the efficacy of cell therapy.91,103 In particular, biomaterial stiffness modulates the regenerative capacity of stem cells.104 First, the substrate stiffness affects the maintenance, activity, and proliferation of stem cells.105,106,107,108 The specific lineage commitment of stem cells has been found to be controlled by substrate stiffness.109,110 Substrate stiffness significantly affects cell migration from the scaffold to the host tissue. For example, softer alginate-based scaffolds improve MSC invasion into the host tissue and significantly enhance tissue matrix production.111 Substrate stiffness affects the persistence of stem cells in vivo112,113 and their adaptation to the host immune114,115 environments. Finally, the matrix stiffness has been shown to affect the morphogenesis of stem cell spheroids and organoids.116
The stiffness of the implant drastically changes the original mechanical environment of the tissues adjacent to the implantation site, leading to varied biological responses. For example, remodeling of the peri-implant bone around a stiffness-optimized Ti-35Nb-2Ta-3Zr alloy implant was significantly enhanced compared to Ti-6Al-4V alloy implants under similar loading conditions.117 Bone scaffolds (3D-printed) with rationally designed stiffness have demonstrated the ability to promote bone regeneration and ingrowth compared to scaffolds with an arbitrary design of stiffness.118,119 In addition, adjusting the mechanical stress at the muscle injury site using a biomaterial with suitable stiffness can modulate the immune response and reduce fibrosis in the injured muscle.120 Myocardial injection of biomaterials into the infarcted and border zones to improve the mechanical properties of the infarcted myocardium is a promising strategy for reducing pathological remodeling after myocardial infarction,121,122 whose therapeutic effect is highly dependent on the stiffness of the injected material.123,124 The influence of substrate stiffness on cell behavior has been discussed extensively in several recent reviews.77,78,79,80

3.1.2. Strain-stiffening behavior

Native ECM and many natural biomaterials exhibit strain-stiffening behavior (i.e., stiffness is not constant, but increases with strain) when subjected to large deformations.125,126,127 A strain stiffening of a polyisocyanopeptide hydrogel behaviors is displayed in Fig. 6(b). The stress-strain curve of the hydrogel is in a stress ramp. K’ is its modulus of elasticity.128
In fact, many soft tissues also exhibit nonlinear elasticity owing to strain stiffening, which becomes difficult to extend as they are deformed. At the tissue level, the strain-stiffening behavior simultaneously enables tissue deformation under physiological mechanical loads and prevents tissue damage under large loads. For example, in blood vessel walls, distensibility at low strains accommodates pulsatile blood flow, whereas increased stiffness at high strains provides mechanical stability to prevent vessel rupture.129
By mimicking the non-linear mechanical behavior of soft tissues, biomaterials with strain-stiffening behavior can better reconstruct the tissue mechanical environment. Importantly, recent findings imply that the strain-stiffening property of gels influences the expression of KCa3.1 potassium channel by mediating cytoskeletal stress fiber formation and affects the functional response of liver carcinoma cells.130 Therefore, biomaterials with strain-stiffening behavior bear considerable potential for tissue repair by reconstructing the tissue mechanical environment and directing cellular behavior. For example, hydrogels with strain-stiffening properties facilitate wound repair.131,132

3.2. Time-dependent mechanical behavior

Native tissues experience a wide range of loading rates and frequencies, from daily physiological loading to high-rate impact injuries. The time-dependent (or frequency-dependent) mechanical behavior of tissues arising from their intrinsic viscoelasticity and fluid flow-induced poroelasticity is critical to their adaptation to a complicated mechanical environment.133 Viscoelastic processes are often associated with stretching and sliding of collagen fibrils (and other macromolecules) within the ECM, whereas poroelastic processes involve local matrix compression and fluid flow. Tissue poroelasticity strongly depends on the experimental length scale, whereas viscoelasticity is independent of the length scale. These two distinct mechanisms are crucial for the physiological functions of connective tissues, such as load bearing, energy dissipation, and storage in cartilage; protection, regulation, and sensing in the skin; and energy dissipation and force transmission in tendons. The time-dependent mechanical properties of soft tissues are increasingly recognized to be linked to several pathological processes. Recently, the effects of viscoelasticity, poroelasticity, and viscosity of biomaterials on cell and tissue responses have been increasingly demonstrated.35,134

3.2.1. Viscoelasticity

Natural tissues and ECM components behave more like viscoelastic or viscoplastic materials, which exhibit prominent stress relaxation behaviors and may undergo permanent deformation (viscoplastic materials) in response to loads.34,35,135 Viscoelasticity models examine how materials exhibit both elastic and viscous behaviors by analyzing how stress and strain are distributed within the material and how they interact with external forces. Initial attempts to model viscous elastic effects commonly employ springs and dashpots. The Maxwell and Kelvin models are fundamental models used to describe viscoelastic materials. Typical constitutive models for viscoelastic creep and stress relaxation under an instantaneous and constant applied stress (σ0) and strain (ε0) include Spring element, Dashpot element, Maxwell model, and Kelvin/Voigt model. A schematic presentation of common viscoelastic models is displayed in Fig. 6(c),136 where t represents time, E denotes the elastic modulus of spring elements, η signifies the viscosity of dashpot elements, and τ denotes the relaxation time.
The Maxwell model combines elements of elasticity and viscosity in series, where the total strain experienced by the material is the cumulative effect of both its elastic and viscous strains, corresponding to the following constitutive equations:
$\dot{\varepsilon}=\frac{\sigma}{\mu}+\frac{\dot{\sigma}}{E}$
where $\cdot \omega$ is the strain rate, that is, the derivative of the strain ε with respect to time; μ is the viscosity coefficient of the viscous element; E is the modulus of elasticity of the elastic element; and $\omega $ and $\cdot \omega$ are the stress and stress rate, respectively.
The Kelvin model corresponds to a parallel connection of elastic and viscous elements with the following constitutive equation:
$ \sigma=E \varepsilon+\mu \dot{\varepsilon} $
The viscoelastic behavior of a material is usually measured using a rheology test and is described by the following equations:
$ G^{*}(\omega)=G_{e}+i G_{N}^{0} \tau_{0} \omega $
$ G^{\prime}=G_{e}, G^{\prime \prime}=G_{N}^{0} \tau_{0} \omega $
where G is complex modulus, ω is oscillation frequency, $G_{N}^{0}$ is the plateau modulus, Ge is equilibrium modulus, and τ0 is relaxation time. Stress relaxation and creep tests were also widely used to evaluate the viscoelastic properties of a material.35,134 There are also many other sophisticated viscoelastic models to describe behaviors of biomaterials. For example, a two-layer structural model for healthy young arterial walls was proposed to describe the viscoelastic behaviors.137 An orthotropic viscoelastic model was proposed for describing the passive mechanical property of myocardium138. A novel fractional anisotropic nonlinear viscoelastic model has been suggested to capture observed phenomena in the heart.57 This model marks the first instance of a nonlinear anisotropic viscoelastic model for human myocardium that has been shown to accurately match the biomechanical behavior of myocardial tissue and provide credible predictions of tissue response.139 A recent study has investigated the use of multiscale mechanical markers to monitor disease progression and treatment. The research revealed a consistent two-stage power-law rheological pattern across various time scales in healthy livers, fibrotic livers, and fibrotic livers treated with mesenchymal stem cells.140
Recently, viscoelastic and viscoplastic hydrogels containing dynamic network bonds have attracted considerable attention for tissue repair applications.126,127 The debonding and rebonding processes allow the polymer matrix to flow under the applied stress. Therefore, viscoelastic and viscoplastic polymers can respond to tiny mechanical changes at the cell-material interface to make rearrangements in the polymer network, which could create a self-adaptive mechanical microenvironment facilitating complex cellular functions.126,127 Many cellular behaviors, including focal adhesion formation, cell spreading, maintenance or differentiation of stem cells, cell migration, and ECM synthesis, have been found to be regulated by the matrix viscoelasticity and viscoplasticity rather than by the initial elastic modulus (Table 1).34,35,45,134,141,142,143,144 Recent studies have indicated that substrate viscoelasticity influences immune responses.145,146,147 These findings provide a basis for the application of viscoelastic and viscoplastic biomaterials as cell delivery vesicles in tissue repair,148 3D cell/organoid culture platform,149,150 and TE scaffolds for guided tissue regeneration.151,152
The viscoelasticity of delivery vesicles affects cell survival during cell delivery with injection approaches by modulating the shear stress on the cells during delivery by injection. For example, injectable viscoelastic hydrogels with stress dissipation capabilities have been developed for cell transplantation to provide mechanical protection to cells during injection and improve cell viability and retention at the transplantation site post-injection.153,154 In addition, the viscoelasticity of delivery vesicles can regulate cell migration155 and differentiation156,157,158 and improve the efficacy of cell therapy. Recently, viscoelastic hydrogels have been used to guide the formation of mature organoids,150,159,160,161 which are better representatives of native tissue than cells. Thus, injectable viscoelastic hydrogels can be used for organoid therapy and have great potential in regenerative medicine.162
Viscoelastic biomaterials can regulate response of host cells during the repair of tissues, such as nerve,163,164 skeletal muscle,165 myocardium,166 skin 167,168, bone,169 osteochondral tissue,14 and intervertebral disc11 injuries. Viscoelastic hydrogels are preferred over their elastic counterparts for tissue repair because they direct appropriate cell differentiation14,164,169 and enable better cell infiltration and tissue integration.170
Viscoelastic properties of biomaterials are important for reconstructing the dynamic mechanical environment of load-bearing tissues. For example, when OA occurs, the viscoelastic property of the synovial fluid in joints changes such that its efficacy as a lubricating, shock-absorbing, and filtering agent is diminished,171 causing an aberrant mechanoenvironment in the cartilage. In contrast, biomaterials with appropriate viscoelastic behaviors have demonstrated the ability to restore the lubricating and shock-absorbing properties of synovial fluids.172 Our recent study revealed that viscoelasticity plays a critical role in allowing epicardial patches to reconstruct the mechanical environment of the myocardium and facilitate tissue restoration after infarction.10 The therapeutic effects might due to the protection of mitochondrial function through mitigating the mechanical challenges.

3.2.2. Poroelasticity

Poroelasticity refers to the mechanics of interactions between fluid flow, stress and solid deformation within a porous medium. Porous materials are solid structures that contain empty spaces or voids. Some biopolymer gels and tissues, such as cartilage, intervertebral disc tissue, are poroelastic materials saturated with fluid that flows relative to a deforming solid matrix. These materials exhibit time-dependent responses to mechanical loads, mainly derived from the solid and fluid phase interactions across multiple length scales. Typical poroelastic behaviors of biopolymer networks are schematically presented in Fig. 6(d). When a fluid is extruded within a polymer, it induces a time-dependent normal force along the axial direction. Compressive deformations altering the volume of porous material will inevitably trigger fluid flow through the network due to the incompressibility of fluids. When deformation occurs rapidly, the system behaves akin to an incompressible material due to the load being predominantly sustained by the incompressibility of the interstitial fluid. Conversely, when deformation happens slowly enough to permit fluid outflow, the system behaves like a sponge-like material devoid of fluid, where the compression of voids notably contributes to the reduction in volume. Upon applying a constant shear stress at t ​= ​0, the subsequent decay of normal stress follows an exponential pattern over time. The characteristic time governing this decay is intricately linked to the pore size/shape/connectedness within the network as well as the friction between fluids and solid matrix (permeability of the porous medium). Notably, the biopolymer gels exhibit significantly smaller characteristic time (seconds to minutes) compared to their synthetic counterparts (hours to days), which means that the characteristic time can be sophisticatedly regulated in a rather wide range. For more details, the reader could be referred to 173.
The Biot formulation of the constitutive equations for a porous material filled with a fluid is used to describe these responses. The Biot formulation relies on certain key assumptions, including the presumption of linearity in the relationship between stress σij and strain εij, as well as the notion of reversibility in the deformation process, meaning that no energy is dissipated during a closed loading cycle. The most general form of the isotropic material response is.
$ \varepsilon_{i j}=\frac{\sigma_{i j}}{2 G}-\left(\frac{1}{6 G}-\frac{1}{9 K}\right) \delta_{i j} \sigma_{k k}+\frac{1}{3 H^{\prime}} \delta_{i j} p $
$ \zeta=\frac{\sigma_{k k}}{3 H H^{\prime}}+\frac{p}{R^{\prime}} $
where K and G are thus identified as the bulk and shear moduli of the drained elastic solid, respectively. The additional constitutive constants H ′, H ″ and R ′ characterize the coupling between the solid and fluid stresses and strains, respectively. For the material parameters, refer to 174.
The poroelasticity-enabled time-dependent mechanical responses of tissues are critical to their physiological functions. For example, energy dissipation due to poroelasticity occurs under tension or compression resulting from volume changes caused by water flowing into or out of the network. A recent study revealed that poroelastic behavior, together with viscoelastic behavior, play an important role in the physiological functions of connective tissues when adapting to loads across different frequency regimes.133
The poroelastic properties of tissues also play essential roles in regulating the homeostasis and function of tissue-resident cells. First, the deformation of the poroelastic tissue results in the flow of interstitial fluid, generating fluid shear stress that affects tissue cell behavior.175 Recently, a poroelastic model that associates the relaxation process with fluid diffusion through a fibrous matrix was built to describe local stresses, strains, and velocities sensed by stem cells for a more accurate prediction of stem cell differentiation.176 Second, water flowing into or out of the network of poroelastic tissues causes osmotic changes in the cell microenvironment, which is an important stimulus for cell mechanosignaling.177 Third, the transport of nutrients from the periphery to the poroelastic tissues and the proper elimination of waste products from the tissues are necessary for tissue survival, which in turn depends on the poroelastic properties of the tissues. This mechanical regulation of tissue nutrition is an indirect mechanism of mechanotransduction.178 Fourth, poroelastic tissues such as hyaline cartilage and nucleus pulposus have high thermodissipative capabilities during cyclic deformations, thus providing a dynamic thermal milieu for the residing cells. A recent study revealed that the self-heating property of poroelastic materials is important for cells to effectively perceive and translate the applied mechanical loading,179 which could be viewed as another indirect mechanotransduction mechanism.

3.2.3. Viscosity

Viscous liquids “flow” in response to stress, with strain increasing linearly (and irreversibly) with time when a constant force is applied. The viscosity is a key property of viscous liquids. A schematic diagram of Newton's law of viscosity can be found in Fig. 6(e). The shear stress can be expressed as τ ​= ​μ (du/dy), where τ is the shear stress, μ is the kinematic viscosity of the liquid, and du/dy is the rate of shear deformation.
Various models have been proposed to describe the viscosities of biomaterials.180 The generalized form of Newton's law of viscosity can be written as
$ \tau=\mu\left[\nabla v+(\nabla v)^{*}\right]-\left(\frac{2}{3} \mu-\kappa\right)(\nabla \cdot v) \delta $
where τ is the viscous stress tensor, flow velocity is denoted as v, the dagger ∗ denotes the transpose, δ is the unit tensor, μ and κ denote the shear and bulk viscosities, respectively.
The influence of matrix viscosity on cellular behavior has also been revealed (Table 1).134,181,182,183,184 For example, substrate viscosity regulates the collective movement of epithelial cells.181 Viscosity-dependence of cell spreading and spheroid formation was observed on poly(ε-caprolactone)-based substrate.183 More viscous surfaces increase the amount of mechanosensitive YAP in the nucleus and enhance myoblast differentiation.184
The effect of viscosity on cell behavior in 3D cultures has been investigated by embedding a viscous gelatin solution in chemically crosslinked gelatin hydrogels.185,186,187 Chondrocytes have a more spread morphology, higher proliferation, lower ECM production, and decreased expression of collagen type II and aggrecan in a low-viscosity gelatin solution than in a high-viscosity solution, suggesting that high viscosity is more beneficial for the maintenance of the chondrocyte phenotype, whereas low viscosity is more beneficial for cell expansion.185,186,187 Using a similar system, the authors revealed that a gelatin solution promoted the proliferation of MSCs, but its promotive effect decreased with an increase in viscosity.186 Importantly, the results showed that high viscosity is beneficial for the osteogenic differentiation of MSCs, whereas low viscosity is beneficial for adipogenic differentiation.186 High-viscosity gelatin solution is also beneficial for chondrogenic differentiation of hMSCs under the synergistic stimulation of chondrogenic induction factors.187 Recently, a high-viscosity buffered red blood cell storage solution was demonstrated to mitigate many aspects of red blood cell storage lesions and the inflammatory response to resuscitation after hemorrhage.188
These studies suggest that viscosity should be considered as one of the key mechanical cues affecting cell-material interactions and is an important factor in engineering materials that control cell behavior.

3.3. Programming mechanical properties of biomaterials to control cell/tissue responses

Biomaterials that exhibit stimuli-responsive changes in mechanical characteristics can temporally exert dynamic mechanical stimuli to adherent cells or restore tissue mechanical environment adaptively,80,191 offering an opportunity to emulate the dynamics of physiological mechanical environment for stimulating tissue repair.126,127 Biomaterials that exhibit stimuli-responsive changes in their mechanical characteristics can temporally exert dynamic mechanical stimuli on adherent cells or adaptively restore the tissue mechanical environment,80,191 offering an opportunity to emulate the dynamics of the physiological mechanical environment to stimulate tissue repair. This type of biomaterial has been explored as a novel platform for steering cell behavior,190,192,193,194,195 revealing cellular mechanobiological responses that are not captured by static biomaterial platforms. First, the dynamic substrate stiffness reversibly controls cell morphology.196 Second, the dynamic substrate affects the commitment and differentiation of stem cells and their spheroids.197,198 Third, the cell phenotype may change with alterations in substrate stiffness.199 Thus, biomaterials offer a unique opportunity for tissue engineering applications, where in situ dynamic alterations of matrix mechanical properties can be programmed to adaptively emulate the mechanical environment of the native extracellular matrix.200 At load-bearing sites, this type of biomaterial can consistently restore the dynamic tissue mechanical environment in adaptation to the tissue healing process, thereby optimizing the dynamic mechanical signals transmitted to biomaterial- or tissue-resident cells.201,202
Biomaterials with stimuli-responsive alternating geometries have also been developed.203,204 Under the actuation of external stimuli, the geometries of these materials change in a programmed manner, actively exerting mechanical forces on the adherent cells and contacting tissues. Such biomaterials provide opportunities to exert mechanical stimuli on tissues/cells that cannot be realized by physiological activities. Efforts have been made to elucidate the effects of biomaterial-programmed dynamic mechanical stimuli on cells.205,206,207 For example, the remote stimulation of an ultrasound-responsive Janus scaffold resulted in mechanical nanovibration that transmitted dynamic mechanical signals to adherent hBMSCs, which enhanced cell proliferation, matrix deposition, and osteogenic differentiation via the formation and activation of voltage-gated calcium ion channels.207 Acoustic radiation force was successfully delivered to cells in gelatin methacryloyl hydrogels using acoustic waves, which induced cell patterning.208 These findings have inspired researchers to develop advanced materials and devices that deliver dynamic mechanical stimuli to cells. This type of biomaterial is capable of exerting mechanical stimuli on the surrounding tissue in a controlled and programmed manner to enhance tissue regeneration. For example, alginate-based hydrogels impregnated with iron oxide nanoparticles enhanced healing in a mouse muscle injury model by applying cyclic mechanical loading to the tibialis anterior under an oscillating magnetic field.209 Similarly, periodic axial muscle stretching applied by an injectable, biodegradable magnetic hydrogel system was superior to massage-like compression in maintaining muscle mass and structure in an animal model, suggesting pathways for combating muscle disuse atrophy.15 A mechanically active gel-elastomer-nitinol tissue adhesive (MAGENTA) was developed to generate and deliver muscle contraction-mimicking stimulation to target tissues with programmed strength and frequency,210 which activated mechanosensing pathways involving yes-associated protein and myocardin-related transcription factor A, and increases the rate of muscle protein synthesis. Disused muscles treated with MAGENTA exhibit greater size and weight and generate higher forces than untreated muscles, demonstrating their potential for atrophy mitigation.210 In addition, dressings that can actively contract wounds in response to skin temperature accelerate skin wound healing.16 Similarly, injectable self-healing hydrogels with thermoresponsive self-contraction and tissue adhesion properties have been shown to assist in wound closure by actively contracting wounds.211 These studies demonstrate the great promise of programming dynamic mechanical stimuli from biomaterials for tissue repair and regeneration. However, the mechanobiological mechanisms behind the therapeutic effects remains to be revealed.4

4. Theoretical tools for deciphering mechanical interactions and designing optimal mechanics and geometry

The development of mechanobiomaterials relies on better prediction of the mechanical interaction between the biomaterial and host tissue/cell to characterize the mechanical cues perceived by cells. Numerical simulation approaches have demonstrated great promise for deepening the mechanistic understanding and guiding the rational design of mechanobiomaterials.
Numerical simulation approaches have been widely used to determine the mechanical properties of implants. Finite element analysis (FEA) is typically used to investigate material/tissue interactions to optimize the mechanical and structural properties of medical devices, including bone screws,212 tissue engineering scaffolds,213,214 stents,215,216 and epicardial patches.10 For example, a strain-programmed patch that implemented a hydration-based shape-memory mechanism was developed to mechanically contract diabetic wounds in a programmable manner (Fig. 7a).217 The optimal mechanical and geometric parameters of the patch were determined using FEA. The patch enhanced the healing of diabetic wounds by promoting faster re-epithelialization and angiogenesis as well as the enrichment of fibroblast populations with a pro-regenerative phenotype in various animal models.217 FEA has also been used to explain or predict the biological responses by coupling established mechanobiological mechanisms with tissue or cell stress/strain.2, 218
Fig. 7. Application of simulation approaches in the development of mechanobiomaterials. (a) With the aid of FEA, a strain-programmed patch was developed, which could mechanically contract diabetic wounds in a programmable manner to enhance the healing of diabetic wounds.217 (b) Guided by coarse-grained MD simulations, peptide amphiphiles (PAs) based assemblies with high internal dynamics was developed for spinal cord injury repair.163
FEA aims to employ and solve relevant partial differential equations to describe physical phenomena within materials and structures on a continuum scale (usually on a millimeter scale for biological systems). As such, it does not model the interactions between individual molecules. In contrast, molecular dynamics (MD) is used for microscopic simulations of interatomic and intermolecular interactions. MD simulations have been used to study material/cell interactions at the subcellular or molecular level and the dynamics of hydrogel networks.163,219,220 For example, in the development of scaffolds for spinal cord injury repair, the internal dynamics of various peptide amphiphilic (PAs)-based assemblies were compared using coarse-grained MD simulations (Fig. 7b).163 In vitro and in vivo studies have revealed that the degree of internal dynamics is critical to the bioactivity of scaffolds,163 which resulted in notable differences in vascular growth, axonal regeneration, myelination, survival of motor neurons, reduced gliosis, and functional recovery.163 Recently, using coarse-grained MD simulations, the receptor-mediated endocytosis of elastic nanoparticles of different sizes and shapes was systematically investigated,220 providing a detailed mechanistic understanding of the influence of nanoparticle size, shape, and elasticity on membrane-wrapping efficiency, which serves as a foundation for the rational design of nanoparticle-based drug carriers.

5. Rational design of mechanobiomaterials

A thorough understanding of the mechanobiological effects of materials on cell and tissue functions is the foundation for the rational design of mechanobiomaterials to promote tissue repair and regeneration. One of the basic considerations of mechanobiomaterials is the optimization of the dynamic mechanical interactions between tissues and biomaterials to facilitate tissue repair and regeneration. When exogenous cells are incorporated into biomaterials for cell-based therapy, another basic consideration is the optimization of the mechanical environment of the cells to enable their ability to mediate tissue regeneration. Based on these two considerations, a typical workflow for the inverse design of mechanobiomaterials was proposed (Fig. 8). The rational integration of a material's mechanical properties and geometry/structure to achieve targeted effects on biological functions is important, but also challenging. In particular, physical models and simulation approaches that can decipher the mechanical interactions and mechanisms between materials and biological systems are necessary.
Fig. 8. A closed-loop workflow of designing mechanobiomaterials.

5.1. Function-oriented optimization of mechanics and geometry

The design of mechanobiomaterials can sometimes be deemed as the reverse engineering of the material's mechanical properties and geometries, according to the established knowledge of mechanobiology and analysis of multiscale mechanical interactions between biological systems and biomaterials. A typical workflow is as follows:
Computational models allow the simulation of biomechanical processes under normal and pathological conditions based on the quantitative analysis of changes in tissue mechanics and geometry after tissue lesions. Biomaterials with different mechanical properties and geometries were included in the computational model to capture the dynamics of the mechanical environment in biological systems and biomaterials. Computational models based on mechanobiology should aim to predict how biomaterials affect the mechanobiological processes of tissue or biomaterial-resident cells through mechanical interactions. After assessing the biological impact of biomaterials on target tissues/cells, mechanics-geometry-function (MGF) relationships can be established at multiple scales, and key mechanical/geometric parameters can be identified and optimized to maximize the therapeutic efficacy of biomaterials. Then, the key mechanical/geometrical parameters and their optimal values can be used to design biomaterials and their fabrication processes with the aid of machine learning and/or other novel approaches based on material databases (e.g., mechanical properties, biodegradability, and biocompatibility). Finally, the designed biomaterials were prepared and investigated using material characterization and biological and preclinical/clinical evaluations to verify the success of the design. The closed-loop workflow is shown in Fig. 8. Our research team used this closed-loop workflow and developed a viscoelastic epicardial patch with an optimized shear modulus and loss factor to restore the dynamic mechanical environment of the infarcted myocardium for MI treatment (Fig. 9).10 In this study, the mechanical properties and geometry of the epicardial patch were optimized by deciphering the mechanical interactions between the biomaterial and the myocardium through computational modelling. The optimal epicardial patch can restore the systolic function of left ventricle to nearly normal levels.10
Fig. 9. Development of a mechanobiologically optimized epicardial patch using a closed-loop workflow of reverse design.10
Such a reverse design strategy inverts the conventional “trial and error” paradigm of biomaterial development by starting with a set of desirable functionalities and searching for the optimal designs and fabrication routes. The essence of mechanobiomaterials is to achieve the reverse design of biomaterials through the computationally aided optimization of mechanics and geometry. This approach not only assists in the design of biomaterials with improved functionalities, but also represents a new paradigm compared to current practices in biomaterial development, where the mechanical properties and geometry of the biomaterials are often adjusted according to large quantities of tests and data analysis. A reverse-design approach for mechanobiomaterials could help reduce technological and experimental efforts to produce biomedical devices for regenerative medicine.

5.2. Integration and synergy of material mechanics and biology

Native tissues exhibit a combination of spatiotemporally varying mechanical characteristics at multiple length scales to adapt to their dynamic mechanical environments, maintain physiological functions, and deliver appropriate mechanical stimuli to tissue cells. Inspired by native tissues, the integration of mechanical characteristics with different time dependences at different length scales in a spatiotemporally controlled manner is critical to the design of mechanobiomaterials that are capable of simultaneously re-constructing the mechanical environment for both tissue- and biomaterial-resident cells.
For re-constructing the mechanical environment of constantly deformed tissues, both the stiffness and stress-relaxation behaviors of biomaterials should be tailored to accommodate the time-varying tissue deformation. A prominent example is our viscoelastic epicardial patch with optimized stiffness and stress-relaxation behavior to restore the dynamic mechanical environment of the infarcted myocardium for MI treatment;10 which was identified to effectively prevent adverse left ventricular remodeling, whereas the stress-relaxation behavior was designed to maintain the pumping function of left ventricle. Recently, we developed a glycerol-crosslinked PVA gel (GPG) for the treatment of IVDD,11 which has a nucleus pulposus (NP)-matched shear modulus to restore the mechanical environment of IVD tissues under static loading while possessing a prominent energy dissipation capability to accommodate the dynamic loading environment. The injection of GPG efficiently restored IVD height in a needle puncture model and prevented IVD pathological remodeling.11
Natural ECM in load-bearing tissues is both macroscopically stable to withstand repeated biomechanical challenges and microscopically dynamic to allow for cellular activities.173,221 Recapitulating the mechanical characteristics of the ECM in synthetic hydrogels is critical for many biomedical applications. Recently, cartilage-like tissue was fabricated by 3D printing using two biomaterials with different mechanical properties: a hard biomaterial (with a compressive modulus on the order of MPa) to achieve the macro-mechanical properties of cartilage, and a soft biomaterial (with a compressive modulus on the order of kPa) to create a chondrogenic micromechanical environment (Fig. 10).222 The hierarchical mechanical design ensured appropriate mechanical stimuli for the hMSC spheroids in the scaffold, resulting in high viability and chondrogenic-like behavior.222 It is worth noting that cells in scaffold would also affect the macro-mechanical characteristics of scaffolds under external load via the “tensional homeostasis” mechanism,59 which should be taken into consideration for rational design of cell-laden mechanobiomaterials for repair of load-bearing tissues.
Fig. 10. Schematic of a 3D bioprinting approach for engineering cartilage-like mechanically hierarchical constructs with a hard biomaterial to mimic the macro-mechanical properties of native cartilage, and a soft biomaterial to create a chondrogenic microenvironment.222

Spatiotemporal control of mechanical properties is also critical for the precisely reconstructing cell/tissue mechanical environment. An example is the development of a bilayer osteochondral scaffold with a 3D-printed Ti alloy subchondral bone compartment and a freeze-dried poly-lactic-co-glycolic acid (PLGA) reinforced collagen sponge as cartilage compartment was developed.223 The mechanical support provided by the 3D-printed Ti alloy layer facilitates long-term regeneration of cartilage by accelerating osteochondral formation and its integration with the adjacent host tissue. Recently, osteochondral scaffolds with spatially heterogeneous stiffness demonstrated the stiffness-specific induction of chondrogenic and osteogenic differentiation of stem cells for cartilage and bone regeneration, respectively.189 Most recently, the multidirectional stiffness design of artificial IVD demonstrated the ability to simulate the physiological kinematic behaviors of native IVD, bearing great potential in reconstructing the mechanical environment of adjacent tissues.224

5.3. Structural and geometrical optimization

By resorting to a microgeometry-driven approach, the design domain of biomaterials is becoming much wider, which also enables a more accurate modulation of tissue/cell-material mechanical interactions. By tailoring the geometrical parameters of the unit cells as basic building blocks, desirable mechanical properties can be obtained,225,226 such as tunable Poisson's ratio,227 chirality,228 and high specific stiffness.229 Certain tissues in the human body, such as tendons, skin, annulus fibrosus disks, arteries, and cancellous bone, exhibit auxetic characteristics when subjected to tensile forces. Auxetic materials can serve as scaffolds in tissue engineering. These materials not only demonstrate unique mechanical attributes but can also enhance cell proliferation when subjected to external loads.230
Numerous studies have highlighted the potential of auxetic materials for tissue engineering and biomedical devices. In particular, auxetic materials have great potential for the design and fabrication of mechanobiomaterials231 because their effective mechanical properties can be tuned by modifying the underlying microstructures, thereby reproducing the mechanical behaviors of skin tissue, arteries, tendons, IVD, and cancellous bone.230,232,233 Numerous auxetic structures have been successfully fabricated and investigated based on various microstructure design motifs including “re-entrant”,234 “chiral”,235 and “rotating”231 types of unit cell.
Auxetic materials have also gained substantial attention for biomedical device applications (Fig. 11). In particular, artificial auxetic polyurethane IVDs possess a remarkable ability to bend and twist, potentially offering superior biomechanical performance compared with conventional disc replacement options. Because of their anisotropic negative Poisson's ratio, these discs effectively prevent the occurrence of bulges that may harm nearby nerve endings.233 Importantly, these discs closely replicated the behavior exhibited by natural lumbar IVDs.233 An auxetic pedicle screw was created using a re-entrant unit cell structure,236 achieving outstanding screw-bone fixation performance in spinal surgery. It was found that implant-bone contact and implant longevity were improved by using auxetic-based hybrid implants when pressing onto the bones from both the medial and lateral sides.237 An innovative auxetic coronary stent enables the stent maintains a specific lumen volume by simultaneously expanding in two directions when the balloon is inflated.238 Analysis of the stent diameter and length before and after tension revealed that the auxetic coronary stent exhibited radial and longitudinal expansion. This expandability is advantageous because it allows the stents to adapt to different vessel dimensions. Other applications include an auxetic material patch for the treatment of myocardial infarction,239 where the unique property of having a negative Poisson's ratio allows the patch to effectively adapt to the intricate mechanical requirements of the heart, as well as auxetic materials used in the neck240 and ankle braces.241
Fig. 11. Various auxetic materials for biomedical applications.239,246,247,248,249,251,252,253,254,255
The presence of auxetic properties in several biological tissues has sparked interest in the use of auxetic materials for tissue engineering.230 Studies have indicated that auxetic structure-based scaffolds establish a biophysical environment that influences the organization of actin and vascular differentiation of pluripotent stem cells.242,243 Several studies have examined the effects of loading cells within auxetic scaffolds in a dynamic loading environment mimicking physiological conditions.244,245
Biomaterials with programmable mechanical properties and structures/geometries can actively exert mechanical stimuli on tissues and cells. Various modelling approaches have shown great promise for describing the mechanical interactions between materials and biological systems at the tissue, cellular, and subcellular levels. In the proposed closed-loop reverse design of mechanobiomaterials, mathematical/physical modeling and simulation approaches play a critical role in establishing MGF relationships at multiple scales to identify and optimize the key mechanical/geometrical parameters of biomaterials with the aim of maximizing their therapeutic efficacy. Data-driven machine-learning approaches will play increasingly important roles in the design and fabrication of mechanobiomaterials.250

6. Summary and outlook

Recently, mechanobiology of living tissues has been increasingly incorporated into the design of biomaterials with desirable biological functions. Programmed mechanical interactions between biomaterials and host tissues/cells can attenuate inflammatory responses, regulate the activity of regeneration-related cells, and reconstruct physiologically relevant mechanical environments that are crucial for tissue repair and regeneration. Based on an increasing number of studies devoted to the development of mechanobiologically optimized materials with enhanced tissue repair and regeneration capabilities, a new field of mechanobiomaterials is emerging. This field represents a paradigm shift in biomaterial engineering from passive adaptation to tissue mechanical environment to proactive regulation of the tissue mechanical environment. This article aims to profile this newly emerging field by summarizing the existing theoretical and experimental underpinnings, development, and design tools, as well as selective examples of success in this field.
Despite rapid development, the field of mechanobiomaterials remains at a very early stage and their potential for tissue repair and regeneration is yet to be fully realized. The design of mechanobiomaterials is still limited by a lack of quantitative guidelines for optimizing the mechanical parameters. Although some basic guidelines have been established, as summarized in this review, a systematic framework with qualitative and quantitative understanding of MGF relationships is still lacking. Further understanding of MGF relationships relies on novel biomaterial fabrication techniques, a better description of the mechanical interactions between the implant and host tissues, and clarification of cell response mechanisms to mechanical stimuli. Advances in the following aspects are expected to significantly promote the development of mechanobiomaterials and their translation from bench to bed. (1) The role of material mechanics in regulating the biology of material-resident cells needs to be delineated. (2) The connections between the tissue mechanical environment and tissue cell signaling, transcription factor activation, and the epigenome require further exploration. (3) Synergistic effects of mechanical and other physicochemical cues must be elucidated. Multidisciplinary collaborations across biology, materials science, biomedical science, engineering, and other sister disciplines are indispensable to address these challenges. In addition, data-driven machine learning and artificial intelligence tools can be used to accelerate the progress in this field. Current MGF relationships can be utilized in both the design and fabrication of mechanobiomaterials that start from target functionalities and generate the corresponding design parameters for subsequent fabrication.
Ideally, the mechanical properties of the implant material should dynamically match those of native tissue during the entire healing process. However, achieving dynamic mechanical regulation of matching between the implant and tissue during healing progression remains impossible. Biomaterials that have dynamic mechanics in response to time-varied physiological environment can shed light on developing better mechanobiomaterials that consistently adapt to the local mechanical environment for reconstructing tissue homoeostasis or promoting tissue regeneration. Therefore, advances in this type of biomaterials are expected to provide essential technological basis for fast development of mechanobiomaterials. Generally, biomaterials with dynamic mechanics could be achieved by stimuli-responsive property design and self-regulation ability design.
Despite many outstanding challenges, the emerging field of mechanobiomaterials is creating new strategic avenues for tissue regeneration and repair. Further understanding and application of MGF relationships would not only allow the customization of next-generation functional biomaterials, but also provide important guidelines for the development of novel biomedical devices and therapeutics. Joint efforts by materials engineers, mechanical scientists, biologists, and clinicians are required to advance this field.

CRediT authorship contribution statement

Xiao Lin: Funding acquisition, Writing - original draft. Hua Yang: Writing - original draft. Yi Xia: Writing - original draft. Kang Wu: Writing - original draft. Fengcheng Chu: Writing - original draft. Huan Zhou: Writing - original draft, Funding acquisition. Huajian Gao: Supervision, Writing - review & editing. Lei Yang: Funding acquisition, Supervision, Writing - review & editing.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

The authors thank Mr. Shubo Zhang from Nanyang Technological University for valuable discussion on poroelastic materials. This work was supported by the National Natural Science Foundation of China (grant numbers 82025025, U23A6008, 32171321, 32371385, 81802155, 51672184 and 81622032); National Key Research and Development Project of China (grant number 2023YFC2412300); Natural Science Foundation of Hebei Province of China (grant number H2022202007); Full-time Talents Program of Hebei Province of China (grant number 2020HBQZYC012); Natural Science Foundation of Tianjin of China (grant number 21JCZDJC01110, 21JCYBJC01030); Academician Expert Workstation of Yunnan Province of China (202205AF150025); Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD); Key Laboratory of Orthopaedics of Suzhou (SZS2022017); Scientific Research Project of Tianjin Education Commission (2022KJ096); and Science and Technology Innovation Project of Foshan City (1920001000025).
1.
Chen C, Tambe DT, Deng L, Yang L. Biomechanical properties and mechanobiology of the articular chondrocyte. Am J Physiol Cell Physiol. 2013; 305(12):1202-1208.

2.
Zhen G, Guo Q, Li Y, et al. Mechanical stress determines the configuration of TGFβ activation in articular cartilage. Nat Commun. 2021; 12(1):1706.

3.
Hodgkinson T, Kelly DC, Curtin CM, O'Brien FJ. Mechanosignalling in cartilage: an emerging target for the treatment of osteoarthritis. Nat Rev Rheumatol. 2022; 18(2): 67-84.

4.
Sun Q, Wei Q, Zhao C. How do the cells sense and respond to the microenvironment mechanics? Chin Sci Bull. 2021; 66(18):2303-2311.

5.
Agarwal P, Lee H-p, Smeriglio P, et al. A dysfunctional TRPV4-GSK3β pathway prevents osteoarthritic chondrocytes from sensing changes in extracellular matrix viscoelasticity. Nat Biomed Eng. 2021; 5(12):1472-1484.

6.
Deng Y, Lu J, Li W, et al. Reciprocal inhibition of YAP/TAZ and NF-κB regulates osteoarthritic cartilage degradation. Nat Commun. 2018; 9(1):4564.

7.
Donderwinkel I, Tuan RS, Cameron NR, Frith JE. Tendon tissue engineering: current progress towards an optimized tenogenic differentiation protocol for human stem cells. Acta Biomater. 2022; 145:25-42.

8.
Dance A. The secret forces that squeeze and pull life into shape. Nature. 2021; 589(7841):186-188.

9.
Lin X, Bai YJ, Zhou H, Yang L. Mechano-active biomaterials for tissue repair and regeneration. J Mater Sci Technol. 2020; 59:227-233.

10.
Lin X, Liu Y, Bai AB, et al. A viscoelastic adhesive epicardial patch for treating myocardial infarction. Nat Biomed Eng. 2019; 3(8):632-643.

11.
Jia H, Lin X, Wang D, et al. Injectable hydrogel with nucleus pulposus-matched viscoelastic property prevents intervertebral disc degeneration. J Orthop Translat. 2022; 33:162-173.

12.
Zhang Y, Habibovic P. Delivering mechanical stimulation to cells: state of the art in materials and devices design. Adv Mater. 2022; 34(32):2110267.

13.
Cai P, Wang C, Gao H, Chen X. Mechanomaterials: a rational deployment of forces and geometries in programming functional materials. Adv Mater. 2021; 33(46): 2007977.

14.
Liu C, Yu Q, Yuan Z, et al. Engineering the viscoelasticity of gelatin methacryloyl (GelMA) hydrogels via small “dynamic bridges” to regulate BMSC behaviors for osteochondral regeneration. Bioact Mater. 2022; 25:445-459.

15.
Chang L, Li Y, Li M, et al. An Injectable, biodegradable magnetic hydrogel system for exogenous promotion of muscle mass and regeneration. Chem Eng J. 2021; 420: 130398.

16.
Blacklow SO, Li J, Freedman BR, Zeidi M, Chen C, Mooney DJ. Bioinspired mechanically active adhesive dressings to accelerate wound closure. Sci Adv. 2019; 5(7):eaaw3963.

17.
Yu B, Kang SY, Akthakul A, et al. An elastic second skin. Nat Mater. 2016; 15(8): 911-918.

18.
Selvaprithiviraj V, Vaquette C, Ivanovski S. Hydrogel based soft tissue expanders for orodental reconstruction. Acta Biomater. 2023; 172:53-66.

19.
Coletta R, Olivieri C, Persano G, Solari V, Inserra A, Morabito A. Expanding intestinal segment using osmotic hydrogel: an in vivo study. J Biomed Mater Res B. 2019; 107(4):1304-1309.

20.
Lavrador P, Esteves MR, Gaspar VM, Mano JF. Stimuli-responsive nanocomposite hydrogels for biomedical applications. Adv Funct Mater. 2021; 31(8):2005941.

21.
Uto K, Tsui JH, DeForest CA, Kim DH. Dynamically tunable cell culture platforms for tissue engineering and mechanobiology. Prog Polym Sci. 2017; 65:53-82.

22.
Lampi MC, Reinhart-King CA. Targeting extracellular matrix stiffness to attenuate disease: from molecular mechanisms to clinical trials. Sci Transl Med. 2018; 10(422): eaao0475.

23.
Sun Q, Pei F, Zhang M, et al. Curved nanofiber network induces cellular bridge formation to promote stem cell mechanotransduction. Adv Sci. 2023; 10(3): 2204479.

24.
Liu W, Sun Q, Zheng Z-L, et al. Topographic cues guiding cell polarization via distinct cellular mechanosensing pathways. Small. 2022; 18(2):2104328.

25.
Xi W, Saw TB, Delacour D, Lim CT, Ladoux B. Material approaches to active tissue mechanics. Nat Rev Mater. 2019; 4(1):23-44.

26.
Janaszak MM, Wolfe RP, Ahsan T. Chapter 2: Biomechanics in Stem Cell Manufacturing. Stem Cell Manufacturing. 2016:27-ISBN: 978-0-444-63265-4.

27.
Schvartzman C, Zhao H, Ibarboure E, Ibrahimova V, Garanger E, Lecommandoux S. Control of enzyme reactivity in response to osmotic pressure modulation mimicking dynamic assembly of intracellular organelles. Adv Mater. 2023; 35(33):2301856.

28.
Guo Y, Li D, Zhang S, et al. Visualizing intracellular organelle and cytoskeletal interactions at nanoscale resolution on millisecond timescales. Cell. 2018; 175(5): 1430-1442.e17.

29.
Zhao J, Meng F, Qian J, Huang Y, Fan Y. In vitro cell stretching devices and their applications: from cardiomyogenic differentiation to tissue engineering. Medicine in Novel Technology and Devices. 2023; 18:100220.

30.
Onal S, Alkaisi MM, Nock V. Microdevice-based mechanical compression on living cells. iScience. 2022; 25(12):105518.

31.
Galie PA, van Oosten A, Chen CS, Janmey PA. Application of multiple levels of fluid shear stress to endothelial cells plated on polyacrylamide gels. Lab Chip. 2015; 15(4):1205-1212.

32.
Rodriguez ML, McGarry PJ, Sniadecki NJ. Review on cell mechanics: experimental and modeling approaches. Appl Mech Rev. 2013; 65(6):060801.

33.
T. Sadhanasatish, K. Augustin, L. Windgasse, A. Chrostek-Grashoff, M. Rief, C. Grashoff, A molecular optomechanics approach reveals functional relevance of force transduction across talin and desmoplakin. Sci Adv. 9(25) eadg3347.

34.
Vining KH, Marneth AE, Adu-Berchie K, et al. Mechanical checkpoint regulates monocyte differentiation in fibrotic niches. Nat Mater. 2022; 21:939-950.

35.
Chaudhuri O, Cooper-White J, Janmey PA, Mooney DJ, Shenoy VB. Effects of extracellular matrix viscoelasticity on cellular behaviour. Nature. 2020; 584(7822): 535-546.

36.
Cai Y, Rossier O, Gauthier NC, et al. Cytoskeletal coherence requires myosin-IIA contractility. J Cell Sci. 2010; 123(3):413-423.

37.
Oria R, Wiegand T, Escribano J, et al. Force loading explains spatial sensing of ligands by cells. Nature. 2017; 552(7684):219-224.

38.
Houk Andrew R, Jilkine A, Mejean Cecile O, et al. Membrane tension maintains cell polarity by confining signals to the leading edge during neutrophil migration. Cell. 2012; 148(1):175-188.

39.
De R, Safran SA. Dynamical theory of active cellular response to external stress. Phys Rev E. 2008; 78(3):031923.

40.
Keren K. Membrane tension leads the way. Proc Natl Acad Sci USA. 2011; 108(35): 14379-14380.

41.
Ladoux B, M_ege R-M. Mechanobiology of collective cell behaviours. Nat Rev Mol Cell Biol. 2017; 18(12):743-757.

42.
Gupta M, Sarangi BR, Deschamps J, et al. Adaptive rheology and ordering of cell cytoskeleton govern matrix rigidity sensing. Nat Commun. 2015; 6(1):7525.

43.
Batchelder EL, Hollopeter G, Campillo C, et al. Membrane tension regulates motility by controlling lamellipodium organization. Proc Natl Acad Sci USA. 2011; 108(28): 11429-11434.

44.
Zhou EH, Trepat X, Park CY, et al. Universal behavior of the osmotically compressed cell and its analogy to the colloidal glass transition. Proc Natl Acad Sci USA. 2009; 106(26):10632-10637.

45.
Chaudhuri O, Gu L, Klumpers D, et al. Hydrogels with tunable stress relaxation regulate stem cell fate and activity. Nat Mater. 2016; 15(3):326-334.

46.
Coste B, Mathur J, Schmidt M, et al. Piezo1 and Piezo 2 are essential components of distinct mechanically activated cation channels. Science. 2010; 330(6000):55-60.

47.
Olesen S-P, Claphamt D, Davies P. Haemodynamic shear stress activates a Kt current in vascular endothelial cells. Nature. 1988; 331(6152):168-170.

48.
Gerhold KA, Schwartz MA. Ion channels in endothelial responses to fluid shear stress. Physiology. 2016; 31(5):359-369.

49.
Basbaum AI, Bautista DM, Scherrer G, Julius D. Cellular and molecular mechanisms of pain. Cell. 2009; 139(2):267-284.

50.
Cox CD, Bae C, Ziegler L, et al. Removal of the mechanoprotective influence of the cytoskeleton reveals PIEZO 1 is gated by bilayer tension. Nat Commun. 2016; 7(1): 10366.

51.
Kefauver JM, Ward AB, Patapoutian A. Discoveries in structure and physiology of mechanically activated ion channels. Nature. 2020; 587(7835):567-576.

52.
Cox CD, Bavi N, Martinac B. Biophysical principles of ion-channel-mediated mechanosensory transduction. Cell Rep. 2019; 29(1):1-12.

53.
Morachevskaya EA, Sudarikova AV. Actin dynamics as critical ion channel regulator: ENaC and Piezo in focus. Am J Physiol Cell Physiol. 2021 ;320(5):C696-C702.

54.
Botello-Smith WM, Jiang W, Zhang H, et al. A mechanism for the activation of the mechanosensitive Piezo 1 channel by the small molecule Yoda1. Nat Commun. 2019; 10(1):4503.

55.
Chen X, Wanggou S, Bodalia A, et al. A feedforward mechanism mediated by mechanosensitive ion channel PIEZO1 and tissue mechanics promotes glioma aggression. Neuron. 2018; 100(4):799-815.e7.

56.
Lee W, Leddy HA, Chen Y, et al. Synergy between Piezo1 and Piezo2 channels confers high-strain mechanosensitivity to articular cartilage. Proc Natl Acad Sci USA. 2014; 111(47):E5114-E5122.

57.
Zhang Y, Jiang L, Huang T, et al. Mechanosensitive cation channel Piezo1 contributes to ventilator-induced lung injury by activating RhoA/ROCK1 in rats. Respir Res. 2021; 22(1):250.

58.
He J, Fang B, Shan S, et al. Mechanical stretch promotes hypertrophic scar formation through mechanically activated cation channel Piezo1. Cell Death Dis. 2021; 12(3):226.

59.
Stamenović D, Smith ML. Tensional homeostasis at different length scales. Soft Matter. 2020; 16(30):6946-6963.

60.
Gauthier NC, Masters TA, Sheetz MP. Mechanical feedback between membrane tension and dynamics. Trends Cell Biol. 2012; 22(10):527-535.

61.
Liu H, Hu J, Zheng Q, et al. Piezo1 channels as force sensors in mechanical forcerelated chronic inflammation. Front Immunol. 2022; 13:816149.

62.
Vining KH, Mooney DJ. Mechanical forces direct stem cell behaviour in development and regeneration. Nat Rev Mol Cell Biol. 2017; 18(12):728-742.

63.
Dupont S, Morsut L, Aragona M, et al. Role of YAP/TAZ in mechanotransduction. Nature. 2011; 474(7350):179-183.

64.
Swaminathan V, Waterman CM. The molecular clutch model for mechanotransduction evolves. Nat Cell Biol. 2016; 18(5):459-461.

65.
Kanchanawong P, Shtengel G, Pasapera AM, et al. Nanoscale architecture of integrin-based cell adhesions. Nature. 2010; 468(7323):580-584.

66.
Xia Y, Fan X, Yang H, et al. ZnO/Nanocarbons-Modified fibrous scaffolds for stem cell-based osteogenic differentiation. Small. 2020; 16(38):2003010.

67.
Mitra SK, Hanson DA, Schlaepfer DD. Focal adhesion kinase: in command and control of cell motility. Nat Rev Mol Cell Biol. 2005; 6(1):56-68.

68.
Zhao G, Zhang X, Lu TJ, Xu F. Recent advances in electrospun nanofibrous scaffolds for cardiac tissue engineering. Adv Funct Mater. 2015; 25(36):5726-5738.

69.
B. Cheng, W. Wan, G. Huang, Y. Li, G.M. Genin, M.R.K. Mofrad, T.J. Lu, F. Xu, M. Lin, Nanoscale integrin cluster dynamics controls cellular mechanosensing via FAKY397 phosphorylation. Sci Adv. 6(10) eaax1909.

70.
Ege N, Dowbaj AM, Jiang M, et al. Quantitative analysis reveals that actin and srcfamily kinases regulate nuclear YAP1 and its export. Cell Syst. 2018; 6(6): 692-708.e13.

71.
Xia Y, Yang H, Li S, et al. Multivalent polyanionic 2D nanosheets functionalized nanofibrous stem cell-based neural scaffolds. Adv Funct Mater. 2021; 31(20): 2010145.

72.
Dahl KN, Ribeiro AJS, Lammerding J. Nuclear shape, mechanics, and mechanotransduction. Circulation Res. 2008; 102(11):1307-1318.

73.
Wang P, Geng J, Tan C, Wei Q. Substrate stiffness dominants cell gene expression via regulation of HDAC3 subcellular localization. Colloid and Interface Science Communications. 2023; 55:100719.

74.
Geng J, Kang Z, Sun Q, et al. Microtubule assists actomyosin to regulate cell nuclear mechanics and chromatin accessibility. Research. 2023; 6:54.

75.
Vernerey FJ, Lalitha Sridhar S, Muralidharan A, Bryant SJ. Mechanics of 3D cell-hydrogel interactions: experiments, models, and mechanisms. Chem Rev. 2021; 121(18):11085-11148.

76.
Peng Z, Resnick A, Young Y-N. Primary cilium: a paradigm for integrating mathematical modeling with experiments and numerical simulations in mechanobiology. Math Biosci Eng MBE. 2021; 18(2):1215.

77.
Guimar-aes CF, Gasperini L, Marques AP, Reis RL. The stiffness of living tissues and its implications for tissue engineering. Nat Rev Mater. 2020; 5(5):351-370.

78.
Holle AW, Young JL, Van Vliet KJ, et al. Cell-extracellular matrix mechanobiology: forceful tools and emerging needs for basic and translational research. Nano Lett. 2018; 18(1):1-8.

79.
Han FX, Zhu CH, Guo QP, Yang HL, Li B. Cellular modulation by the elasticity of biomaterials. J Mater Chem B. 2016; 4(1):9-26.

80.
Wan X, Liu Z, Li L. Manipulation of stem cells fates: the master and multifaceted roles of biophysical cues of biomaterials. Adv Funct Mater. 2021; 31(23):2010626.

81.
Huang D, Huang Y, Xiao Y, et al. Viscoelasticity in natural tissues and engineered scaffolds for tissue reconstruction. Acta Biomater. 2019; 97:74-92.

82.
Holzapfel GA. Nonlinear Solid Mechanics: A Continuum Approach for Engineering Science. Dordrecht: Kluwer Academic Publishers; 2002.

83.
Cheng J, Jia Z, Li T. A constitutive model of microfiber reinforced anisotropic hydrogels: with applications to wood-based hydrogels. J Mech Phys Solid. 2020; 138: 103893.

84.
Meng Q, Shi X. A mechanistically motivated constitutive model of biopolymer hydrogels with structural evolution. J Mech Phys Solid. 2023; 173:105205.

85.
Y.-c. Fung, Biomechanics: Mechanical Properties of Living Tissues, Springer Science & Business Media2013.

86.
Holzapfel GA, Ogden RW. An arterial constitutive model accounting for collagen content and cross-linking. J Mech Phys Solid. 2020; 136:103682.

87.
Holzapfel GA, Ogden RW. A damage model for collagen fibres with an application to collagenous soft tissues. Proceedings of the Royal Society A. 2020; 476(2236): 20190821.

88.
Kuhl E, Garikipati K, Arruda EM, Grosh K. Remodeling of biological tissue: mechanically induced reorientation of a transversely isotropic chain network. J Mech Phys Solid. 2005; 53(7):1552-1573.

89.
Ma YH, Shi HJ, Wei QS, et al. Developing a mechanically matched decellularized spinal cord scaffold for the in situ matrix-based neural repair of spinal cord injury. Biomaterials. 2021; 279:121192.

90.
Hu Q, Liu M, Chen G, Xu Z, Lv Y. Demineralized bone scaffolds with tunable matrix stiffness for efficient bone integration. ACS Appl Mater Interfaces. 2018; 10(33): 27669-27680.

91.
Schweller RM, Wu ZJ, Klitzman B, West JL. Stiffness of protease sensitive and cell adhesive PEG hydrogels promotes neovascularization in vivo. Ann Biomed Eng. 2017; 45(6):1387-1398.

92.
Yang R, Huang J, Zhang W, et al. Mechanoadaptive injectable hydrogel based on poly(γ-glutamic acid) and hyaluronic acid regulates fibroblast migration for wound healing. Carbohydr Polym. 2021; 273:118607.

93.
Basurto IM, Passipieri JA, Gardner GM, et al. Photoreactive hydrogel stiffness influences volumetric muscle loss repair. Tissue Eng. 2022; 28(7-8):312-329.

94.
Xu P, Yang Q, Zhang L, et al. A multi-functional SiO 3 2_-releasing hydrogel with bioinspired mechanical properties and biodegradability for vascularized skeletal muscle regeneration. J Mater Chem B. 2022; 10(37):7540-7555.

95.
Le LV, Mohindra P, Fang QZ, et al. Injectable hyaluronic acid based microrods provide local micromechanical and biochemical cues to attenuate cardiac fibrosis after myocardial infarction. Biomaterials. 2018; 169:11-21.

96.
Liao JF, Tian TR, Shi SR, et al. The fabrication of biomimetic biphasic CAN-PAC hydrogel with a seamless interfacial layer applied in osteochondral defect repair. Bone Res. 2017; 5:17018.

97.
Abaricia JO, Farzad N, Heath TJ, Simmons J, Morandini L, Olivares-Navarrete R. Control of innate immune response by biomaterial surface topography, energy, and stiffness. Acta Biomater. 2021; 133:58-73.

98.
Sridharan R, Cavanagh B, Cameron AR, Kelly DJ, O'Brien FJ. Material stiffness influences the polarization state, function and migration mode of macrophages. Acta Biomater. 2019; 89:47-59.

99.
Woods I, O'Connor C, Frugoli L, et al. Biomimetic scaffolds for spinal cord applications exhibit stiffness-dependent immunomodulatory and neurotrophic characteristics. Adv Healthcare Mater. 2022; 11(3):2101663.

100.
Jansen LE, Amer LD, Chen EYT, et al. Zwitterionic PEG-PC hydrogels modulate the foreign body response in a modulus-dependent manner. Biomacromolecules. 2018; 19(7):2880-2888.

101.
Chen M, Zhang Y, Zhou P, et al. Substrate stiffness modulates bone marrow-derived macrophage polarization through NF-κB signaling pathway. Bioact Mater. 2020; 5(4):880-890.

102.
Kim J-K, Han S-B, Park SI, Kim I-S, Kim D-H. Nuclear transport of STAT6 determines the matrix rigidity dependent M2 activation of macrophages. Biomaterials. 2022: 121859.

103.
Wang LS, Du C, Toh WS, Wan ACA, Gao SJ, Kurisawa M. Modulation of chondrocyte functions and stiffness-dependent cartilage repair using an injectable enzymatically crosslinked hydrogel with tunable mechanical properties. Biomaterials. 2014; 35(7): 2207-2217.

104.
Chen C, Song J, Qiu J, Zhao J. Repair of a meniscal defect in a rabbit model through use of a thermosensitive, injectable, in situ crosslinked hydrogel with encapsulated bone mesenchymal stromal cells and transforming growth factor β1. Am J Sports Med. 2020; 48(4):884-894.

105.
Gouveia RM, Lepert G, Gupta S, Mohan RR, Paterson C, Connon CJ. Assessment of corneal substrate biomechanics and its effect on epithelial stem cell maintenance and differentiation. Nat Commun. 2019; 10:1496.

106.
Wong SW, Tamatam CR, Cho IS, et al. Inhibition of aberrant tissue remodelling by mesenchymal stromal cells singly coated with soft gels presenting defined chemomechanical cues. Nat Biomed Eng. 2022; 6(1):54-66.

107.
Gilbert PM, Havenstrite KL, Magnusson KEG, et al. Substrate elasticity regulates skeletal muscle stem cell self-renewal in culture. Science. 2010; 329(5995): 1078-1081.

108.
Li HK, Yu B, Yang PZ, et al. Injectable AuNP-HA matrix with localized stiffness enhances the formation of gap junction in engrafted human induced pluripotent stem cell-derived cardiomyocytes and promotes cardiac repair. Biomaterials. 2021; 279:121231.

109.
Gegg C, Yang F. Spatially patterned microribbon-based hydrogels induce zonallyorganized cartilage regeneration by stem cells in 3D. Acta Biomater. 2020; 101: 196-205.

110.
Huebsch N, Lippens E, Lee K, et al. Matrix elasticity of void-forming hydrogels controls transplanted-stem-cell-mediated bone formation. Nat Mater. 2015; 14(12): 1269-1277.

111.
Fonseca KB, Gomes DB, Lee K, et al. Injectable MMP-sensitive alginate hydrogels as hMSC delivery systems. Biomacromolecules. 2014; 15(1):380-390.

112.
Mao AS, Shin JW, Utech S, et al. Deterministic encapsulation of single cells in thin tunable microgels for niche modelling and therapeutic delivery. Nat Mater. 2017; 16(2):236-243.

113.
Mao AS, Ozkale B, Shah NJ, et al. Programmable microencapsulation for enhanced mesenchymal stem cell persistence and immunomodulation. P Natl Acad Sci USA. 2019; 116(31):15392-15397.

114.
Wang Y, Chen XD, Cao W, Shi YF. Plasticity of mesenchymal stem cells in immunomodulation: pathological and therapeutic implications. Nat Immunol. 2014; 15(11):1009-1016.

115.
Wong SW, Lenzini S, Cooper MH, Mooney DJ, Shin JW. Soft extracellular matrix enhances inflammatory activation of mesenchymal stromal cells to induce monocyte production and trafficking. Sci Adv. 2020; 6(15):eaaw0158.

116.
Cruz-Acu-na R, Quirós M, Huang S, et al. PEG-4MAL hydrogels for human organoid generation, culture, and in vivo delivery. Nat Protoc. 2018; 13(9):2102-2119.

117.
Mao C, Yu W, Jin M, et al. Mechanobiologically optimized Ti-35Nb-2Ta-3Zr improves load transduction and enhances bone remodeling in tilted dental implant therapy. Bioact Mater. 2022; 16:15-26.

118.
Nasello G, Vautrin A, Pitocchi J, et al. Mechano-driven regeneration predicts response variations in large animal model based on scaffold implantation site and individual mechano-sensitivity. Bone. 2021; 144:115769.

119.
Xu C, Yu SN, Wu WZ, Liu QP, Ren LQ. Direct ink writing of Fe bone implants with independently adjustable structural porosity and mechanical properties. Addit Manuf. 2022; 51:102589.

120.
Wang SY, Yan HG, Fang B, et al. A myogenic niche with a proper mechanical stress environment improves abdominal wall muscle repair by modulating immunity and preventing fibrosis. Biomaterials. 2022; 285:121519.

121.
Morita M, Eckert CE, Matsuzaki K, et al. Modification of infarct material properties limits adverse ventricular remodeling. Ann Thorac Surg. 2011; 92(2):617-624.

122.
Berry MF, Engler AJ, Woo YJ, et al. Mesenchymal stem cell injection after myocardial infarction improves myocardial compliance. Am J Physiol Heart Circ Physiol. 2006 ;290(6):H2196-H2203.

123.
Ifkovits JL, Tous E, Minakawa M, et al. Injectable hydrogel properties influence infarct expansion and extent of postinfarction left ventricular remodeling in an ovine model. Proc Natl Acad Sci U S A. 2010; 107(25):11507-11512.

124.
Ghanta RK, Pugazenthi A, Zhao Y, et al. Influence of supraphysiologic biomaterial stiffness on ventricular mechanics and myocardial infarct reinforcement. Acta Biomater. 2022:30-39.

125.
Storm C, Pastore JJ, MacKintosh FC, Lubensky TC, Janmey PA. Nonlinear elasticity in biological gels. Nature. 2005; 435(7039):191-194.

126.
Xie W, Wei X, Kang H, et al. Static and dynamic: evolving biomaterial mechanical properties to control cellular mechanotransduction. Adv Sci. 2023; 10(9):2204594.

127.
Sun Q, Hou Y, Chu Z, Wei Q. Soft overcomes the hard: flexible materials adapt to cell adhesion to promote cell mechanotransduction. Bioact Mater. 2022; 10: 397-404.

128.
Jaspers M, Dennison M, Mabesoone MF, MacKintosh FC, Rowan AE, Kouwer PH. Ultra-responsive soft matter from strain-stiffening hydrogels. Nat Commun. 2014; 5(1):5808.

129.
Shadwick RE. Mechanical design in arteries. J Exp Biol. 1999; 202(23):3305-3313.

130.
Liu Z, Fu J, Yuan H, et al. Polyisocyanide hydrogels with tunable nonlinear elasticity mediate liver carcinoma cell functional response. Acta Biomater. 2022; 148:152-162.

131.
op ‘t Veld RC, van den Boomen OI, Lundvig DMS, et al. Thermosensitive biomimetic polyisocyanopeptide hydrogels may facilitate wound repair. Biomaterials. 2018; 181: 392-401.

132.
Op 't Veld RC, Joosten L, van den Boomen OI, et al. Monitoring(111)In-labelled polyisocyanopeptide (PIC) hydrogel wound dressings in full-thickness wounds. Biomater Sci. 2019; 7(7):3041-3050.

133.
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. 2018; 70:249-259.

134.
Cantini M, Donnelly H, Dalby MJ, Salmeron-Sanchez M. The plot thickens: the emerging role of matrix viscosity in cell mechanotransduction. Adv Healthcare Mater. 2020; 9(8):1901259.

135.
Chaudhuri O, Gu L, Darnell M, et al. Substrate stress relaxation regulates cell spreading. Nat Commun. 2015; 6:6365.

136.
Roeder RK. Mechanical characterization of biomaterials. Characterization of Biomaterials. 2013:49-104.

137.
Holzapfel GA, Gasser TC, Stadler M. A structural model for the viscoelastic behavior of arterial walls: continuum formulation and finite element analysis. Eur J Mech Solid. 2002; 21(3):441-463.

138.
Gültekin O, Sommer G, Holzapfel GA. An orthotropic viscoelastic model for the passive myocardium: continuum basis and numerical treatment. Comput Methods Biomech Biomed Eng. 2016; 19(15):1647-1664.

139.
Nordsletten D, Capilnasiu A, Zhang W, et al. A viscoelastic model for human myocardium. Acta Biomater. 2021; 135:441-457.

140.
Chang Z, Zhang L, Hang J-T, Liu W, Xu G-K. Viscoelastic multiscale mechanical indexes for assessing liver fibrosis and treatment outcomes. Nano Lett. 2023; 23(20): 9618-9625.

141.
Webber MJ, Tibbitt MW. Dynamic and reconfigurable materials from reversible network interactions. Nat Rev Mater. 2022; 7:541-556.

142.
Clark AG, Maitra A, Jacques C, et al. Self-generated gradients steer collective migration on viscoelastic collagen networks. Nat Mater. 2022; 21:1200-1210.

143.
Ma YF, Han T, Yang QX, et al. Viscoelastic cell microenvironment: hydrogel-based strategy for recapitulating dynamic ECM mechanics. Adv Funct Mater. 2021; 31(24): 2100848.

144.
Madl CM, LeSavage BL, Dewi RE, et al. Maintenance of neural progenitor cell stemness in 3D hydrogels requires matrix remodelling. Nat Mater. 2017; 16(12): 1233-1242.

145.
Zhou YW, Wu Y. Substrate viscoelasticity amplifies distinctions between transient and persistent LPS-induced signals. Adv Healthcare Mater. 2022; 11(8):2102271.

146.
Sadtler K, Wolf MT, Ganguly S, et al. Divergent immune responses to synthetic and biological scaffolds. Biomaterials. 2019; 192:405-415.

147.
Feng Q, Wei KC, Lin SE, et al. Mechanically resilient, injectable, and bioadhesive supramolecular gelatin hydrogels crosslinked by weak host-guest interactions assist cell infiltration and in situ tissue regeneration. Biomaterials. 2016; 101: 217-228.

148.
Lou JZ, Liu F, Lindsay CD, Chaudhuri O, Heilshorn SC, Xia Y. Dynamic hyaluronan hydrogels with temporally modulated high injectability and stability using a biocompatible catalyst. Adv Mater. 2018; 30(22):1705215.

149.
Baker AEG, Bahlmann LC, Tam RY, et al. Benchmarking to the gold standard: hyaluronan-oxime hydrogels recapitulate xenograft models with in vitro breast cancer spheroid culture. Adv Mater. 2019; 31(36):1901166.

150.
Chrisnandy A, Blondel D, Rezakhani S, Broguiere N, Lutolf MP. Synthetic dynamic hydrogels promote degradation-independent in vitro organogenesis. Nat Mater. 2022; 21(4):479-487.

151.
Richardson BM, Wilcox DG, Randolph MA, Anseth KS. Hydrazone covalent adaptable networks modulate extracellular matrix deposition for cartilage tissue engineering. Acta Biomater. 2019; 83:71-82.

152.
Tseng TC, Hsieh FY, Theato P, Wei Y, Hsu SH. Glucose-sensitive self-healing hydrogel as sacrificial materials to fabricate vascularized constructs. Biomaterials. 2017; 133:20-28.

153.
Correa S, Grosskopf AK, Lopez Hernandez H, et al. Translational applications of hydrogels. Chem Rev. 2021; 121(18):11385-11457.

154.
An P, Wei H, Zhang YS, et al. A mechanically adaptive "all-sugar" hydrogel for cellladen injection. Eur Polym J. 2022; 174:111328.

155.
Hou S, Wang XF, Park S, Jin XB, Ma PX. Rapid self-integrating, injectable hydrogel for tissue complex regeneration. Adv Healthcare Mater. 2015; 4(10):1491-1495.

156.
Feng Q, Li D, Li Q, et al. Dynamic nanocomposite microgel assembly with microporosity, injectability, tissue-adhesion, and sustained drug release promotes articular cartilage repair and regeneration. Adv Healthcare Mater. 2022; 11(8): e2102395.

157.
Li WQ, Wu DW, Hu D, et al. Stress-relaxing double-network hydrogel for chondrogenic differentiation of stem cells. Mat Sci Eng C-Mater. 2020; 107:110333.

158.
Hong H, Seo YB, Kim DY, et al. Digital light processing 3D printed silk fibroin hydrogel for cartilage tissue engineering. Biomaterials. 2020; 232:119679.

159.
Yavitt FM, Brown TE, Hushka EA, et al. The effect of thiol structure on allyl sulfide photodegradable hydrogels and their application as a degradable scaffold for organoid passaging. Adv Mater. 2020; 32(30):1905366.

160.
Glorevski N, Sachs N, Manfrin A, et al. Designer matrices for intestinal stem cell and organoid culture. Nature. 2016; 539(7630):560-564.

161.
Ai SF, Li H, Zheng H, et al. A SupraGel for efficient production of cell spheroids. Sci China Mater. 2022; 65(6):1655-1661.

162.
Kim W, Gwon Y, Park S, Kim H, Kim J. Therapeutic strategies of three-dimensional stem cell spheroids and organoids for tissue repair and regeneration. Bioact Mater. 2023; 19:50-74.

163.
Alvarez Z, Kolberg-Edelbrock AN, Sasselli IR, et al. Bioactive scaffolds with enhanced supramolecular motion promote recovery from spinal cord injury. Science. 2021; 374(6569):848-856.

164.
Zhou L, Fan L, Yi X, et al. Soft conducting polymer hydrogels cross-linked and doped by tannic acid for spinal cord injury repair. ACS Nano. 2018; 12(11):10957-10967.

165.
Li T, Hou J, Wang L, et al. Bioprinted anisotropic scaffolds with fast stress relaxation bioink for engineering 3D skeletal muscle and repairing volumetric muscle loss. Acta Biomater. 2022; 156:21-36.

166.
Roshanbinfar K, Vogt L, Greber B, et al. Electroconductive biohybrid hydrogel for enhanced maturation and beating properties of engineered cardiac tissues. Adv Funct Mater. 2018; 28(42):1803951.

167.
He W, Bai J, Chen X, et al. Reversible dougong structured receptor-ligand recognition for building dynamic extracellular matrix mimics. Proc Natl Acad Sci USA. 2022; 119(8):e2117221119.

168.
Zhou ST, Hokugo A, McClendon M, et al. Bioactive peptide amphiphile nanofiber gels enhance burn wound healing. Burns. 2019; 45(5):1112-1121.

169.
Darnell M, Young S, Gu L, et al. Substrate stress-relaxation regulates scaffold remodeling and bone formation in vivo. Adv Healthcare Mater. 2017; 6(1): 1601185.

170.
Zou L, Braegelman AS, Webber MJ. Dynamic supramolecular hydrogels spanning an unprecedented range of host-guest affinity. ACS Appl Mater Interfaces. 2019; 11(6): 5695-5700.

171.
Pontes-Quero GM, Garcia-Fernandez L, Aguilar MR, Roman JS, Cano JP, Vazquez- Lasa B. Active viscosupplements for osteoarthritis treatment. Semin Arthritis Rheum. 2019; 49(2):171-183.

172.
Cai ZX, Zhang HB, Wei Y, Wu M, Fu AL. Shear-thinning hyaluronan-based fluid hydrogels to modulate viscoelastic properties of osteoarthritis synovial fluids. Biomater Sci-Uk. 2019; 7(8):3143-3157.

173.
Burla F, Mulla Y, Vos BE, Aufderhorst-Roberts A, Koenderink GH. From mechanical resilience to active material properties in biopolymer networks. Nature Reviews Physics. 2019; 1(4):249-263.

174.
Wang M, Liu S, Xu Z, et al. Characterizing poroelasticity of biological tissues by spherical indentation: an improved theory for large relaxation. J Mech Phys Solid. 2020; 138:103920.

175.
Buschmann MD, Kim Y-J, Wong M, Frank E, Hunziker EB, Grodzinsky AJ. Stimulation of aggrecan synthesis in cartilage explants by cyclic loading is localized to regions of high interstitial fluid Flow1. Archives Biochem Biophysics. 1999; 366(1): 1-7.

176.
Yuan D, Somers SM, Grayson WL, Spector AA. A poroelastic model of a fibrousporous tissue engineering scaffold. Sci Rep. 2018; 8(1):5043.

177.
Esteki MH, Malandrino A, Alemrajabi AA, Sheridan GK, Charras G, Moeendarbary E. Poroelastic osmoregulation of living cell volume. iScience. 2021; 24(12):103482.

178.
Malandrino A, Jackson AR, Huyghe JM, Noailly J. Poroelastic modeling of the intervertebral disc: a path toward integrated studies of tissue biophysics and organ degeneration. MRS Bull. 2015; 40(4):324-332.

179.
Stampoultzis T, Guo Y, Nasrollahzadeh N, Karami P, Pioletti DP. Mimicking loading-induced cartilage self-heating in vitro promotes matrix formation in chondrocyte-laden constructs with different mechanical properties. ACS Biomaterials Sci Eng. 2023; 9(2):651-661.

180.
Yu Z, Liu J, Wei X. Achieving outstanding damping performance through bioinspired sutural tessellations. J Mech Phys Solid. 2020; 142:104010.

181.
Murrell M, Kamm R, Matsudaira P. Substrate viscosity enhances correlation in epithelial sheet movement. Biophysical J. 2011; 101(2):297-306.

182.
Gong Z, Szczesny SE, Caliari SR, et al. Matching material and cellular timescales maximizes cell spreading on viscoelastic substrates. Proc Natl Acad Sci USA. 2018 ; 115(12):E2686-E2695.

183.
Uto K, Mano SS, Aoyagi T, Ebara M. Substrate fluidity regulates cell adhesion and morphology on poly(ε-caprolactone)-based materials. ACS Biomaterials Sci Eng. 2016; 2(3):446-453.

184.
Bennett M, Cantini M, Reboud J, Cooper JM, Roca-Cusachs P, Salmeron-Sanchez M. Molecular clutch drives cell response to surface viscosity. Proc Natl Acad Sci USA. 2018; 115(6):1192-1197.

185.
Lee K, Chen YZ, Li XM, et al. Solution viscosity regulates chondrocyte proliferation and phenotype during 3D culture. J Mater Chem B. 2019; 7(48):7713-7722.

186.
Lee K, Chen YZ, Yoshitomi T, Kawazoe N, Yang YN, Chen GP. Osteogenic and adipogenic differentiation of mesenchymal stem cells in gelatin solutions of different viscosities. Adv Healthcare Mater. 2020; 9(23):2000617.

187.
Lee K, Chen YZ, Li XM, Kawazoe N, Yang YN, Chen GP. Influence of viscosity on chondrogenic differentiation of mesenchymal stem cells during 3D culture in viscous gelatin C solution-embedded hydrogels. J Mater Sci Technol. 2021; 63:1-8.

188.
Pulliam KE, Joseph B, Makley AT, et al. Improving packed red blood cell storage with a high-viscosity buffered storage solution. Surgery. 2022; 171(3):833-842.

189.
Xu G, Ding ZZ, Lu Q, et al. Electric field-driven building blocks for introducing multiple gradients to hydrogels. Protein Cell. 2020; 11(4):267-285.

190.
Rizwan M, Baker AEG, Shoichet MS. Designing hydrogels for 3D cell culture using dynamic covalent crosslinking. Adv Healthcare Mater. 2021; 10(12):2100234.

191.
Dong YX, Ramey-Ward AN, Salaita K. Programmable mechanically active hydrogelbased materials. Adv Mater. 2021; 33(46):2006600.

192.
Gunay KA, Ceccato TL, Silver JS, et al. PEG-anthracene hydrogels as an on- demand stiffening matrix to study mechanobiology. Angew Chem, Int Ed. 2019; 58(29): 9912-9916.

193.
Humphries BA, Aliabouzar M, Quesada C, et al. Ultrasound-induced mechanical compaction in acoustically responsive scaffolds promotes spatiotemporally modulated signaling in triple negative breast cancer. Adv Healthcare Mater. 2022; 11(10):2101672.

194.
Rosales AM, Vega SL, DelRio FW, Burdick JA, Anseth KS. Hydrogels with reversible mechanics to probe dynamic cell microenvironments. Angew Chem, 2017; 56(40):12132-12136.

195.
He QL, Lin JX, Zhou FH, et al. Musical dish efficiently induces osteogenic differentiation of mesenchymal stem cells through music derived microstretch with variable frequency. Bioeng Transl Med. 2022; 7(2):e10291.

196.
Jeon O, Kim TH, Alsberg E. Reversible dynamic mechanics of hydrogels for regulation of cellular behavior. Acta Biomater. 2021; 136:88-98.

197.
Das RK, Gocheva V, Hammink R, Zouani OF, Rowan AE. Stress-stiffening-mediated stem-cell commitment switch in soft responsive hydrogels. Nat Mater. 2016; 15(3): 318-325.

198.
Zhang JG, Yang H, Abali BE, Li MJ, Xia Y, Haag R. Dynamic mechanics-modulated hydrogels to regulate the differentiation of stem-cell spheroids in soft microniches and modeling of the nonlinear behavior. Small. 2019; 15(30):1901920.

199.
Farrell E, Aliabouzar M, Quesada C, et al. Spatiotemporal control of myofibroblast activation in acoustically-responsive scaffolds via ultrasound-induced matrix stiffening. Acta Biomater. 2022; 138:133-143.

200.
Miao S, Cui H, Esworthy T, et al. 4D self-morphing culture substrate for modulating cell differentiation. Adv Sci. 2020; 7(6):1902403.

201.
Tsang KMC, Annabi N, Ercole F, et al. Facile one-step micropatterning using photodegradable gelatin hydrogels for improved cardiomyocyte organization and alignment. Adv Funct Mater. 2015; 25(6):977-986.

202.
Ji H, Xi K, Zhang Q, Jia X. Photodegradable hydrogels for external manipulation of cellular microenvironments with real-time monitoring. RSC Adv. 2017; 7(39): 24331-24337.

203.
Lim JW, Kim H-j, Kim Y, et al. An active and soft hydrogel actuator to stimulate live cell clusters by self-folding. Polymers. 2020; 12(3):583.

204.
Li Y, Huang G, Gao B, et al. Magnetically actuated cell-laden microscale hydrogels for probing strain-induced cell responses in three dimensions. NPG Asia Mater. 2016; 8(1). e238-e238.

205.
Filippi M, Dasen B, Guerrero J, et al. Magnetic nanocomposite hydrogels and static magnetic field stimulate the osteoblastic and vasculogenic profile of adiposederived cells. Biomaterials. 2019; 223:119468.

206.
Tomas AR, Goncalves AI, Paz E, Freitas P, Domingues RMA, Gomes ME. Magnetomechanical actuation of magnetic responsive fibrous scaffolds boosts tenogenesis of human adipose stem cells. Nanoscale. 2019; 11(39):18255-18271.

207.
Camarero-Espinosa S, Moroni L. Janus 3D printed dynamic scaffolds for nanovibration-driven bone regeneration. Nat Commun. 2021; 12(1):1031.

208.
Armstrong JPK, Puetzer JL, Serio A, et al. Engineering anisotropic muscle tissue using acoustic cell patterning. Adv Mater. 2018; 30(43):1802649.

209.
Cezar CA, Roche ET, Vandenburgh HH, Duda GN, Walsh CJ, Mooney DJ. Biologicfree mechanically induced muscle regeneration. Proc Natl Acad Sci USA. 2016; 113(6):1534-1539.

210.
Nam S, Seo BR, Najibi AJ, McNamara SL, Mooney DJ. Active tissue adhesive activates mechanosensors and prevents muscle atrophy. Nat Mater. 2023; 22(2): 249-259.

211.
Li M, Liang YP, He JH, Zhang HL, Guo BL. Two-pronged strategy of biomechanically active and biochemically multifunctional hydrogel wound dressing to accelerate wound closure and wound healing. Chem Mater. 2020; 32(23):9937-9953.

212.
Mau JR, Hawkins KM, Woo SLY, Kim KE, McCullough MBA. Design of a new magnesium-based anterior cruciate ligament interference screw using finite element analysis. Journal of Orthopaedic Translation. 2020; 20:25-30.

213.
Chang RASH, Shanley JF, Kersh ME, Harley BAC. Tough and tunable scaffoldhydrogel composite biomaterial for soft-to-hard musculoskeletal tissue interfaces. Sci Adv. 2020; 6(34):eabb6763.

214.
Liang H, Chao L, Xie D, et al. Trabecular-like Ti-6Al-4V scaffold for bone repair: a diversified mechanical stimulation environment for bone regeneration. Compos B Eng. 2022; 241:110057.

215.
Chen CX, Chen JH, Wu W, et al. In vivo and in vitro evaluation of a biodegradable magnesium vascular stent designed by shape optimization strategy. Biomaterials. 2019; 221:119414.

216.
Liu JP, Yao XH, Wang ZW, et al. A flexible porous chiral auxetic tracheal stent with ciliated epithelium. Acta Biomater. 2021; 124:153-165.

217.
Theocharidis G, Yuk H, Roh H, et al. A strain-programmed patch for the healing of diabetic wounds. Nat Biomed Eng. 2022; 6:1118-1133.

218.
Legant WR, Miller JS, Blakely BL, Cohen DM, Genin GM, Chen CS. Measurement of mechanical tractions exerted by cells in three-dimensional matrices. Nat Methods. 2010; 7(12):969-971.

219.
Yang BG, Wei KC, Loebel C, et al. Enhanced mechanosensing of cells in synthetic 3D matrix with controlled biophysical dynamics. Nat Commun. 2021; 12(1):3514.

220.
Shen ZQ, Ye HL, Yi X, Li Y. Membrane wrapping efficiency of elastic nanoparticles during endocytosis: size and shape matter. ACS Nano. 2019; 13(1):215-228.

221.
Zhang K, Feng Q, Fang Z, Gu L, Bian L. Structurally dynamic hydrogels for biomedical applications: pursuing a fine balance between macroscopic stability and microscopic dynamics. Chem Rev. 2021; 121(18):11149-11193.

222.
de Melo BAG, Jodat YA, Mehrotra S, et al. 3D printed cartilage-like tissue constructs with spatially controlled mechanical properties. Adv Funct Mater. 2019; 29(51): 1906330.

223.
Yang T, Tamaddon M, Jiang L, et al. Bilayered scaffold with 3D printed stiff subchondral bony compartment to provide constant mechanical support for longterm cartilage regeneration. Journal of Orthopaedic Translation. 2021; 30:112-121.

224.
Song G, Qian Z, Liu X, et al. Bioinspired intervertebral disc with multidirectional stiffness prepared via multimaterial additive manufacturing. Adv Funct Mater. 2023; a(n/a):2300298.

225.
Li J, Zhang Q, Huang R, Li X, Gao H. Towards understanding the structure-property relationships of heterogeneous-structured materials. Scripta Mater. 2020; 186: 304-311.

226.
Yang H, Abali BE, Timofeev D, Müller WH. Determination of metamaterial parameters by means of a homogenization approach based on asymptotic analysis. Continuum Mech Thermodyn. 2020; 32:1251-1270.

227.
Kolken H, Lietaert K, van der Sloten T, et al. Mechanical performance of auxetic meta-biomaterials. J Mech Behav Biomed Mater. 2020; 104:103658.

228.
Feng Y, C€olfen H, Xiong R. Organized mineralized cellulose nanostructures for biomedical applications. J Mater Chem B. 2023; 11:5321-5349.

229.
Zhang L, Hu Z, Wang MY, Feih S. Hierarchical sheet triply periodic minimal surface lattices: design, geometric and mechanical performance. Mater Des. 2021; 209: 109931.

230.
Mardling P, Alderson A, Jordan-Mahy N, Le Maitre-Mahy CL. The use of auxetic materials in tissue engineering. Biomater Sci-Uk. 2020; 8(8):2074-2083.

231.
Shirzad M, Zolfagharian A, Bodaghi M, Nam SY. Auxetic metamaterials for boneimplanted medical devices: recent advances and new perspectives. Eur J Mech Solid. 2023; 98:104905.

232.
Xin X, Liu L, Liu Y, Leng J. 4D printing auxetic metamaterials with tunable, programmable, and reconfigurable mechanical properties. Adv Funct Mater. 2020; 30(43):2004226.

233.
Jiang Y, Shi K, Zhou L, et al. 3D-printed auxetic-structured intervertebral disc implant for potential treatment of lumbar herniated disc. Bioact Mater. 2023; 20: 528-538.

234.
Chen Z, Wu X, Xie YM, Wang Z, Zhou S. Re-entrant auxetic lattices with enhanced stiffness: a numerical study. Int J Mech Sci. 2020; 178:105619.

235.
Dou X, Mehwish N, Zhao C, Liu J, Xing C, Feng C. Supramolecular hydrogels with tunable chirality for promising biomedical applications. Accounts Chem Res. 2020; 53(4):852-862.

236.
Yao Y, Yuan H, Huang H, Liu J, Wang L, Fan Y. Biomechanical design and analysis of auxetic pedicle screw to resist loosening. Comput Biol Med. 2021; 133:104386.

237.
Kolken HM, Janbaz S, Leeflang SM, Lietaert K, Weinans HH, Zadpoor AA. Rationally designed meta-implants: a combination of auxetic and conventional metabiomaterials. Mater Horiz. 2018; 5(1):28-35.

238.
Amin F, Ali MN, Ansari U, Mir M, Minhas MA, Shahid W. Auxetic coronary stent endoprosthesis: fabrication and structural analysis. J Appl Biomater Funct Mater. 2015; 13(2):127-135.

239.
Kapnisi M, Mansfield C, Marijon C, et al. Auxetic cardiac patches with tunable mechanical and conductive properties toward treating myocardial infarction. Adv Funct Mater. 2018; 28(21):1800618.

240.
Panico M, Langella C, Santulli C. Development of a biomedical neckbrace through tailored auxetic shapes. Emerging Science Journal. 2017; 1(3):105-117.

241.
Pattinson SW, Huber ME, Kim S, et al. Additive manufacturing of biomechanically tailored meshes for compliant wearable and implantable devices. Adv Funct Mater. 2019; 29(32):1901815.

242.
Yan Y, Li Y, Song L, Zeng C. Pluripotent stem cell expansion and neural differentiation in 3-D scaffolds of tunable Poisson's ratio. Acta Biomater. 2017; 49: 192-203.

243.
Song L, Ahmed MF, Li Y, Zeng C, Li Y. Vascular differentiation from pluripotent stem cells in 3-D auxetic scaffolds. Journal of tissue engineering and regenerative medicine. 2018; 12(7):1679-1689.

244.
Chen Y-W, Wang K, Ho C-C, Kao C-T, Ng HY, Shie M-Y. Cyclic tensile stimulation enrichment of Schwann cell-laden auxetic hydrogel scaffolds towards peripheral nerve tissue engineering. Mater Des. 2020; 195:108982.

245.
Lutzweiler G, Ndreu Halili A, Engin Vrana N. The overview of porous, bioactive scaffolds as instructive biomaterials for tissue regeneration and their clinical translation. Pharmaceutics. 2020; 12(7):602.

246.
Bobbert F, Janbaz S, Van Manen T, Li Y, Zadpoor A. Russian doll deployable metaimplants: fusion of kirigami, origami, and multi-stability. Mater Des. 2020; 191: 108624.

247.
Kim HW, Kim TY, Park HK, et al. Hygroscopic auxetic on-skin sensors for easyto- handle repeated daily use. ACS Appl Mater Interfaces. 2018; 10(46): 40141-40148.

248.
Jin Y, Xie C, Gao Q, et al. Fabrication of multi-scale and tunable auxetic scaffolds for tissue engineering. Mater Des. 2021; 197:109277.

249.
Lin C, Zhang L, Liu Y, Liu L, Leng J. 4D printing of personalized shape memory polymer vascular stents with negative Poisson's ratio structure: a preliminary study. Sci China Technol Sci. 2020; 63(4):578-588.

250.
Peng B, Wei Y, Qin Y, et al. Machine learning-enabled constrained multi-objective design of architected materials. Nat Commun. 2023; 14(1):6630.

251.
Zhou M, Hou J, Zhang G, et al. Tuning the mechanics of 3D-printed scaffolds by crystal lattice-like structural design for breast tissue engineering. Biofabrication. 2019; 12(1):015023.

252.
Park YJ, Kim JK. The effect of negative Poisson’s ratio polyurethane scaffolds for articular cartilage tissue engineering applications. Adv Mater Sci Eng. 2013; 2013: 853289.

253.
Warner JJ, Gillies AR, Hwang HH, Zhang H, Lieber RL, Chen S. 3D-printed biomaterials with regional auxetic properties. J Mech Behav Biomed Mater. 2017; 76: 145-152.

254.
Kirillova A, Maxson R, Stoychev G, Gomillion CT, Ionov L. 4D biofabrication using shape-morphing hydrogels. Adv Mater. 2017; 29(46):1703443.

255.
Lantada AD, Muslija A, García-Ruíz JP. Auxetic tissue engineering scaffolds with nanometric features and resonances in the megahertz range. Smart Mater Struct. 2015;24(5):055013.

Outlines

/