Jingchao
Sui
ab,
Sarah
Pragnere
ab and
Nicholas A.
Kurniawan
*ab
aDepartment of Biomedical Engineering, Eindhoven University of Technology, PO Box 513, 5600 MB, Eindhoven, the Netherlands. E-mail: kurniawan@tue.nl
bInstitute for Complex Molecular Systems, Eindhoven University of Technology, PO Box 513, 5600 MB, Eindhoven, the Netherlands
First published on 22nd August 2025
The extracellular matrix (ECM) is critical in regulating cell behavior and tissue function. This recognition has driven the development of ECM surrogates to better understand cell–ECM interactions and advance biomedical applications. Hydrogels are promising candidates for this purpose due to their biocompatibility, tunability, and ability to embed cells in 3D environments. While early efforts in the design of ECM-mimicking hydrogels focused on macromolecular type, it is now clear that their biophysical parameters, such as polymer molecular weight, fibrous versus non-fibrous structures, pore size, and mechanical properties, significantly influence cell behavior. Understanding the interplay of these factors is crucial for the rational design of biomaterials, but remains challenging given the complexity of hydrogel systems and the rapid pace of new findings. This review critically evaluates hydrogels as ECM mimics for 3D cell cultures. We revisit key ECM properties to replicate, examine how hydrogel design can meet these needs, and summarize the impact of biochemical, structural, and mechanical features on cell behavior. We also explore how structural and mechanical properties—what cells “see” and “feel”—are interrelated and jointly affect cell function. Our analysis concludes that strategic combinations of polymeric materials will play an important role for next-generation hydrogels to replicate physiological conditions and independently enable precise control over key parameters. These advancements will enhance our understanding of cell–ECM interactions and support the development of innovative biomaterials for tissue engineering and regenerative medicine.
Historically, the first system of hydrogels approaching a 3D cell culture appeared in the 1970s and was based on floating collagen gels,5–8 followed soon after by systems based on reconstituted basement membranes (currently known as Matrigel).8,9 Around the 1990s, hydrogels evolved into a more complex system incorporating non-animal-origin biopolymers such as alginate10,11 for biocompatibility, ease of crosslinking, mild gelation conditions suitable for embedded cells, and their relatively low cost. Synthetic polymers such as dimethylaminoethyl methacrylate (DMAEMA) and methacrylic acid (MAA)12 were then incorporated for better controllability and superior mechanical properties compared to collagen. Hybrid hydrogels13 surfaced as attempts to combine the controllability and advantages of both natural proteins and synthetic components.
From a broader perspective, the use of hydrogels as a biomaterial to embed cells has ushered in the era of three-dimensional (3D) cell culture. Indeed, unlike two-dimensional (2D) culture systems or scaffolds, embedding cells in hydrogels provide a 3D mechanical environment to the cells, which better mimics the in vivo 3D environment.14 Pioneering studies and observations highlighted the impact of dimensionality on cell behavior. Firstly, 2D culture places cells in a planar environment, which imposes restrictions on cell migration along the direction perpendicular to the plane. Besides, cells cultured on 2D plans lack direct exposure to physical confinement as well as the spatial gradients of soluble factors from the surrounding microenvironment, which is present in the 3D environment.15 Secondly, (ventral/dorsal) polarization of cells such as mesenchymal cells in 2D cultures does not appear when these cells are embedded in 3D hydrogels.16 All these aspects influence cellular processes such as cell phenotypic fate. For example, mouse and human mammary epithelial cells embedded in hydrogels maintained a normal phenotype in vitro, while these cells display tumor-like characteristics when cultured in a traditional 2D environment.6,17,18 Later studies have shown that dimensionality also impacted the in vitro maintenance of the phenotype of fully differentiated cells, such as chondrocytes19 or osteocytes.20 These findings thus indicate the importance of 3D culture of cells embedded in physiologically relevant ECM-mimicking hydrogels for better recapitulating in vivo cellular environments and preserving cell native phenotypes.
Initially, hydrogel development for cell culture focused on selecting the macromolecular type. However, the increasing complexity of hydrogel-based systems has revealed that cell responses are also strongly influenced by other, more biophysically oriented parameters, such as polymer molecular weight, fibrous versus non-fibrous structures, pore dimensions, and mechanical properties. The role of each of these parameters, especially in combinatorial permutation with each other, in directing cell behavior and the eventual tissue function is far from understood, even though this knowledge is crucial for the rational design of biomaterials. Moreover, the rapid pace at which various observations of cell response in different types of hydrogel environments are reported has often made it difficult not only to identify which parameters should be considered but also to distill the key determining parameters.
In this review, we will therefore take a step back and critically examine the use of hydrogels as an ECM mimic for 3D cell cultures. We will start by revisiting the properties of the ECM to be reproduced and how this need can be met with the appropriate hydrogel design. Then we will discuss how the structural and mechanical properties of the hydrogels (i.e., what cells see and what cells feel, respectively) are interrelated and strongly impact cell function.
![]() | ||
Fig. 1 Key properties of the ECM. (A) Tissues (left) are composed of cells embedded in ECM (middle). Cells can bind to ECM components such as fibrous collagen I, fibronectin, and hyaluronic acid chain via specific receptors such as integrins (right). (B) Key biochemical (molecular composition), structural (biphasic material, bundle thickness, pore size, and (an)isotropy), and dynamic (degradability and degree of crosslinks) parameters of the ECM that influence the mechanical properties and cell behavior. (C) Example of the unique mechanical behaviors of biological tissues. Left: viscoelastic behavior can be measured by stress-relaxation or dynamic mechanical analysis experiments. In native tissues, viscoelasticity (represented by the values of tan![]() ![]() |
ECM components such as fibronectin and GAGs can bind to various other molecules, such as growth factors, and are involved in the activation of several biological processes.27 For example, heparan sulfate, a type of GAG, can bind to vascular endothelial growth factor and fibroblast growth factor-2 (FGF-2), and its binding to FGF-2 is required for the corresponding signaling and cellular response.28 ECM can directly bind and release growth factors, thereby concentrating them in the vicinity of cells; it can also sequester growth factors to protect them from degradation.28,29 ECM can regulate (enhance30 or inhibit31) the activity and signaling of growth factors, and reciprocally, growth factors regulate the ECM by stimulating cells synthesis or degrading ECM components.28 Extensive overview of ECM–growth factor interactions can be found in review articles by Taipale and Keski-Oja32 and by Schultz and Wysocki;28 specifically for fibronectin–growth factor interactions, see review articles by Clark and colleagues.30,33
As mentioned above, specific components of the ECM enable cell adhesion. Cells bind to ECM components via specific amino acid sequences known as cell-adhesion peptides. Different adhesion peptides activate different receptors on the cell membrane, which in turn trigger different cellular responses (Table 1). Notably, integrins, key transmembrane receptors found only in animal cells, connect the ECM to the inside of the cell, and integrin activation facilitates different bidirectional signal transmission pathways: from the outside toward the inside of cells (outside-in), and from the inside of cells toward the outside (inside-out).23 The binding of cells to ECM via integrin also enables cells to respond to growth factors, and to increase the synthesis of growth factors.28
Matrix component | Binding sequence | Associated cell membrane receptors | Examples of effects of receptor activation |
---|---|---|---|
Collagen I | GFOGER | α1β1, α2β1, α3β1, α10β1, α11β1, discoidin domain receptors (DDR)2 | β1 subunits are essential for tissue repair.34 α11β1 is a major collagen receptor on fibroblastic cells2 while α10β1 binds only to collagen II |
DDRs regulate migration, proliferation, differentiation, survival, and MMP expression | |||
Fibronectin | GRGDS | ανβ3, ανβ5 | ανβ3 activates cell contraction35 and fibroblast activation36 |
Fibrin | RGD | ανβ3, α3β226,37 | |
Laminin | IKVAV | α3β1, α6β1, α6β4, α6β7, α7β1, α1β1, α2β1, α10β12 | Integrins α3β1 and α6β4 can either suppress or promote tumor development and progression38 |
GAGs | — | CD44, RHAMM/CD168, ICAM-12 | The binding of hyaluronic acid (HA) to CD44 and RHAMM guides muscle development,2 and chondroitin sulfate modulates axon growth27 |
Although the main macromolecular components of the ECM are the same throughout the human body, the ECM is highly tissue-specific. For example, the bone matrix is dense and stiff, whereas the ECM in soft connective tissue forms a loose mesh. This diversity and tissue specificity of ECM arise from variations in its specific composition ratios and structural morphology. Therefore, in the next section, we will briefly discuss the main morphological features and parameters of interest in the ECM.
Collagen fibers exhibit a highly hierarchical structure. Three polypeptide chains wrap together into tropocollagen, a triple-helix line approximately 300 nm in length and 1.2–1.5 nm in diameter. These tropocollagens assemble to form fibrils, which in turn assemble into fibers, with a diameter of about 10 μm and a length of millimeters. This hierarchical fibrillar structure allows for mechanical adaptability at different levels, thus providing a variety of mechanical properties and efficient distribution of loading. The network of interconnected fibers that they form also endows the ECM with a strong yet tunable mechanical resilience at very low solid volume fractions, thereby providing the necessary space for other ECM components and cells.39
Elastic fibers, composed of microfibrillar glycoproteins embedded in elastin, also exhibit a hierarchical structure.40 The formation of the elastic fibers starts with the synthesis of the microfibrillar glycoproteins, which act as a scaffold for the deposition of tropoelastin. Crosslinking of the elastin stabilizes the structure.40 In the media of arteries, elastic fibers form concentric layers of elastic lamellae, between which the vascular smooth muscle cells reside. The number of layers depends on the species, ranging from a few layers in mice to more than 50 layers in human arteries. Arteries are subjected to blood pressure, causing cyclic circumferential stretching, which induces a larger strain on the inner side of the arteries. The simultaneous extension and unfolding of the elastic lamellae compensate for this gradient of strain and maintain an even distribution of the circumferential stress throughout all the layers. Thus, the structural heterogeneity contributes to the maintenance of tissue homeostasis.41
The assembly and morphology of ECM fibrils strongly vary depending on the protein type as well as the tissues, organs, the mechanical functions they perform, and the species in which they are found.42 For example, collagen I fibrils are thicker (150–300 nm) than collagen III fibrils (25–100 nm). This difference corresponds to collagen type I predominating in organs bearing high tensile stress, such as bone and tendon, while collagen type III is a major structural component in internal hollow organs.23 Diameters of ECM fibrils in the dermis range from 60 to 120 nm and increase with depth, as mechanical stresses like stretch and compression are more prevalent deeper beneath the skin surface.42,43 The diameters of collagen fibers and elastin fibers in human pulmonary alveolar walls are both close to 1 μm and collagen fiber diameters slightly increase with age, as the alveolar wall size increases while its integrity has to be maintained against the mechanical stresses of respiration.44
The fibers of the ECM are not only important for tissue function (and mechanical properties, as will be discussed further below), but also define pores and spaces in the ECM through which cells can migrate and communicate. The pore size of ECM is typically in the micrometer range and varies across different tissues, organs, and species. Examples of the extremes include human corneal epithelial basement membrane and Descemet's membrane, with pore diameters of ∼92 nm and ∼38 nm,45 respectively, while human lateral meniscus has a pore diameter of 37–48 μm.46Table 2 summarizes the pore sizes of decellularized ECM in different species and tissues.
Decellularized ECM originate from | Pore diameter | Fiber diameter | E | Measurement method for pore and fiber diameter | Measurement method for E | Ref. | |
---|---|---|---|---|---|---|---|
Macroscale | Nanoscale | ||||||
Mice lung | ∼17.5 μm | ∼1.13 μm | ∼17 kPa | ∼2.9 kPa | Scanning electron microscope (SEM) | Macroscale E is measured by macroindentation in a universal micromechanical system; nanoscale E is measured by AFM | 47 |
Mice liver | ∼13.0 μm | ∼1.6 μm | ∼20 kPa | ∼5.6 kPa | |||
Mice kidney | ∼10.7 μm | ∼1.31 μm | ∼22 kPa | ∼6.1 kPa | |||
Mice spleen | ∼10.0 μm | ∼1.0 μm | ∼40 kPa | ∼3.6 kPa | |||
Mice ovary | ∼1.4 μm | ∼0.73 μm | ∼19 kPa | ∼6.5 kPa | |||
Porcine liver | ∼10 μm | — | 1.25–1.31 kPa | Silver staining of sections observed under inverted microscope | Compression test | 48 | |
Rat liver | — | 0.58 ± 0.12 μm | 97 ± 21.22 kPa | SEM | Tensile test | 49 | |
Mice salivary glands | 10–30 μm | — | 120 Pa | SEM | Micro-indentation tester | 50 |
Another crucial structural characteristic of ECM fibers is their orientation, which is directly linked to the physiological function of the tissues. For example, blood vessels are composed of three layers (intima, media, and adventitia) with different fiber organizations in each layer, optimized to sustain shear stress induced by fluid flow as well as intramural pressure induced by the heartbeat.51 In articular cartilage, collagen fibers are oriented parallel to the surface in the superficial layer before transitioning to a perpendicular orientation in the deeper zone to resist wear and transmit the load.52 Tendon fibers are aligned parallel to the main axis to resist the tension.53
The measured tissue stiffness depends on tissue hydration state, scale (from molecule to whole tissues), and test type (traction, bending, indentation compression, shear).56 Single collagen molecules and collagen fibrils (∅ 50–200 nm) have a Young's modulus of a few GPa.57–59 Computational studies estimated a decrease of Young's modulus from the single molecule level (∼6 GPa) to the fibril level (∼0.9 GPa), attributed to the staggered arrangement of collagen molecules in fibrils.60 At a larger scale, the measured elastic modulus of a single collagen fiber (∅ 325 ± 40 nm) was 100–360 MPa.61 Note that in some works describing the stiffness of ECM fibers such as collagen or fibrin, the terms “fibers” and “fibrils” are often used interchangeably to describe the fibril-bundle or fibril structures. Collagen fiber is not extensible, with a breaking strain of 10% for crosslinked fibril and Young's modulus in the range of 200–500 MPa. In contrast, a single fibrin fiber has an elastic modulus on the order of 1–10 MPa with a breaking strain of over 300%.62,63 For elastin, the elastic modulus is approximately 1 MPa with a breaking strain of 150%.64 The rather stiff and not extensible collagen fibers and soft and extensible fibrin and elastin correspond to their functions mentioned above: preventing excessive stretch and supporting elastic deformation, respectively.
At the tissue level, non-mineralized tissues exhibit a wide range of elastic moduli, from hundreds of Pa for the brain65 to several MPa for articular cartilage,66 while mineralized tissues such as bone and dentin can reach several GPa.67,68 The elastic and shear moduli of ex vivo tissues can be measured using mechanical testing machines through compression, tension, bending, or shear experiments. In vivo measurements of tissue require non-invasive methods: Boyer et al.69 measured Young's modulus of skin in vivo by their non-contact airflow system, while Pailler-Mattei et al.70 employed indentation experiments to exclude the elastic response of subcutaneous muscle, accurately estimating the skin's Young's modulus in vivo. Table 3 summarizes the influence of different measuring methods and different anatomical sites on the measured elastic modulus of bone (a hard, mineralized tissue) and skin (a soft, non-mineralized tissue), exemplifying the variations of measurement methods and the measured mechanical properties of human collagenous tissues.
Tissue | Elastic modulus | Method | Ref. |
---|---|---|---|
Bone cortical | 17.4 GPa | Longitudinal compression | 71 |
9.6 GPa | Transversal compression | ||
20 GPa | Nanoindentation | 72 | |
8 GPa | AFM | 73 | |
Bone trabecular | 18 MPa | Longitudinal compression | 72 |
6–10 MPa | Transversal compression | ||
Bone osteoid | 27 kPa | AFM | 74 |
Skin in vivo | 20–1120 kPa | Torsion | 75 and 76 |
4–89 kPa | Indentation | 70 and 77–79 | |
25–260 kPa | Suction | 80,81 | |
0.3–20 MPa | Traction | 82 | |
Skin ex vivo | 70–160 MPa | Traction | 83 and 84 |
0.1–322 kPa | AFM | 85 and 86 |
The viscoelastic response of ECM is critical in understanding ECM. For native soft biological tissues, the ratio of loss modulus to storage modulus (tanδ) ranges from 0.1 to 0.2 when measured at 1 Hz.88 At higher shear frequencies (50–500 Hz), soft tissues such as the brain, liver, and muscles behave more like liquids, with tan
δ values ranging from 0.3 to 0.721 (Fig. 1C). The energy dissipation during loading occurs through three main mechanisms:88 the unbinding or breaking of non-covalent weak bonds among fibers such as collagen,94 the release of entanglements among polymer fibers, and the unfolding of protein fibers such as fibrin.95
Strain-stiffening refers to a mechanical property characterized by a material becoming stiffer or exhibiting increased resistance to deformation as it is increasingly strained. Besides the stress–strain plot, the differential modulus K′ as a function of applied stress σ is also used to describe strain-stiffening (Fig. 1C – right). K′ can be calculated by δσ/δγ. The material exhibits a linear response at low stress levels, where K′ is almost constant and equals the plateau modulus G0. As the stress σ increases beyond a critical threshold σc, the material enters a non-linear regime where K′ can be described as K′ = σm, with m being the stiffening index. This critical stress σc and the corresponding critical strain γc mark the end of the linear regime and the onset of strain-stiffening. Collagen-based tissues are representative examples of strain-stiffening, such as vessels, tendons, and ligaments, which are primarily composed of collagen fibers51,53 (Fig. 1C). On the contrary, strain softening describes a phenomenon wherein a material exhibits decreased resistance to deformation with increasing strain. For instance, during shear loading, tissues such as the brain97 and liver98 exhibit shear-softening.
Some pathologies can alter ECM composition, structure, and mechanical properties, which are inextricably linked.87 Fibrosis, a disease accounting for 45% of deaths in industrialized nations,109 is marked by excessive ECM deposition, which in turn stimulates fibroblasts to increase ECM production, thus activating a positive feedback loop.110 Fibrosis is associated with an overexpression of transglutaminases and LOXs, as well as advanced glycation end products, which increase the extent of collagen and elastin fiber crosslinking in the tissue, thereby increasing the stiffness and reducing the viscosity of the ECM.35 Additionally, fibrosis is also associated with an increased HA content,111 which enhances the swelling pressure, induces isotropic stretching of the collagen fibers, and diminishes anisotropic alignment under tensile stresses, further leading to enhanced isotropic stiffening and reduced anisotropic strain-stiffening of collagen fibers. This leads to the tissue generally becoming more mechanically linear36 and diminishes the effectiveness of long-distance cell force transmission.112 The ECM structure of the fibrotic tissue also changes: in conditions such as idiopathic pulmonary fibrosis, the collagen in ECM is characterized as more highly anisotropic compared to nonfibrotic lung.113 Cancer impacts the ECM structural properties and mechanical properties such as stiffness and viscoelasticity.114 While some cancers such as breast scirrhous carcinoma, prostate cancer, and thyroid cancer increase ECM stiffness, others like intraductal and papillary carcinoma soften it.42 ECM stiffness affects cancer cell morphology, proliferation, invasion, and therapeutic efficacy. However, these effects depend on the cancer type and are not universally applicable, indicating that while ECM stiffness is a promising target for cancer treatment, it necessitates further investigation.115 The alignment of ECM fibers perpendicular to the tumor boundary not only increases tissue stiffness but also forms structures that act as routes that facilitate cancer cell migration.116,117 Cancer-associated fibroblasts are one of the sources that produce and direct the assembly of an anisotropic network of collagen I, which promotes tumor cell spread from the primary tumor site.118 Keloid fibroblasts also induce ECM anisotropy.119
Polymers not derived from natural ECM, such as non-mammalian polysaccharides124 (e.g., alginate, dextran) or synthetic polymers125 (e.g., polyethylene glycol (PEG),126 polyvinyl alcohol (PVA),127 ureido-pyrimidinone (UPy)128), require functionalization to enable cell binding. While various peptides can be used for functionalization, RGD is the most common. Additionally, polymers not derived from the ECM cannot be naturally degraded by cells. This limitation can be overcome by incorporating MMP-degradable crosslinkers. For example, PEG gels can be crosslinked with MMP-sensitive bonds to facilitate cell-mediated degradation.129–132
The use of synthetic polymers has improved our knowledge of the effect of the biochemical environment on cell behavior. Notably, the functionalization of polymers enables researchers to vary ligand density independently of polymer concentrations133,134 and to vary ligand type independently of mechanical properties.135 This revealed that both ligand type135 and ligand density134 influence cell phenotype and traction forces. For example, valvular interstitial cells embedded in PEG remained round in the presence of IKVAV peptides but were able to spread in the presence of RGD.135 Increasing the concentration of RGD peptide in PEG hydrogels from 0.5 mM to 2 mM enabled the formation of vinculin complexes at the periphery of mouse embryonic fibroblasts, which resulted in larger strains and stresses on the surrounding matrix.134
Synthetic hydrogels also enable studies on the importance of matrix degradation for cell spreading and differentiation. For example, the differentiation of osteoblasts to mature osteocytes requires dendrite extension, which is enhanced in PEG gel containing MMP-sensitive crosslinkers.130 Furthermore, degradability is of major importance for regenerative medicine as it enhances cellular invasion from neighboring tissues136 and neo-tissue formation.131,137
Pore size can be imaged using scanning electron microscopy (SEM), though this technique introduces a bias as it requires drying the hydrogels, which can distort the microstructure.121 Other imaging techniques such as confocal microscopy associated with image post-processing144 or fluorescence recovery after photobleaching (FRAP) can be used to estimate pore size in hydrated states.121 Mean pore size or mean pore diameter can be represented by permeability and can be determined by diffusion-driven transport of FITC-labeled dextran (Table 4). Maximum pore size can be estimated by DNA electrophoresis.145 The polymer network of hydrogels creates molecular-size openings in the nanometer scale for most non-fibrous synthetic hydrogels, and micrometer scale in fibrous hydrogels such as collagen and fibrin.54,146 The range of pore sizes depends on the experimental conditions, for example, concentrations of polymers and crosslinkers, and temperature.
Polymer | Pore diameter | Fibril diameter | E | G′ | G” | Measurement method for pore and fibril diameter | Measurement method for mechanical properties (E, G′, G′′) | Ref. |
---|---|---|---|---|---|---|---|---|
Collagen-I 1 mg mL−1, gelation temperature 22 °C | 11 μm | — | — | ∼0.7 Pa | — | Confocal reflectance microscopy images | Rheology | 148 |
Collagen-I 1 mg mL−1, gelation temperaturen 37 °C | ∼5 μm | ∼2 Pa | ||||||
Collagen 1 mg mL−1 | 5.27 μm | 126.83 nm | 1.40 kPa | — | — | Pore size: confocal laser scanning microscope of fluorescence image of collagen matrix; collagen fibrils: scanning electron microscopy | Effective elastic modulus was measured by a custom-made indentation apparatus | 149 |
Collagen 1 mg mL−1 + Na2SO4 0.025 mol L−1 | 4.73 μm | 142.00 nm | 1.65 kPa | |||||
Collagen 1 mg mL−1 + Na2SO4 0.10 mol L−1 | 6.57 μm | 161.66 nm | 1.73 kPa | |||||
Collagen 1 mg mL−1 + Na2SO4 0.25 mol L−1 | 6.84 μm | 201.85 nm | 2.01 kPa | |||||
Collagen-I 1 mg mL−1 | — | 40 ± 3.1 nm | — | 2.63 ± 0.86 kPa | — | Aomic force microscope (AFM) image | Rheology | 150 |
Collagen-I 1.2 mg mL−1 (from rat tail) | — | — | ∼0.36 kPa | 119 ± 63 Pa | 13 ± 6 Pa | — | Macroindentation by universal micromechanical system; rheology | 47 |
Collagen 1.5 mg mL−1, pH = 7 | 4.5 μm | — | — | 6.1 Pa | — | Bubble analysis from confocal reflectance images | Oscillatory shear rheology | 151 |
Collagen-I 1.5 mg mL−1; (from pure rat tail monomers) | 7.7 ± 1.4 μm | — | 63.0 ± 48.5 Pa | — | — | Confocal laser scanning microscopy | AFM | 152 |
Collagen-I 1.5 mg mL−1; (rat tail monomers![]() ![]() ![]() |
7.3 ± 0.7 μm | 101.2 ± 68.5 Pa | ||||||
Collagen-I 1.5 mg mL−1; (from pure bovine skin monomers) | 6.9 ± 1.5 μm | 76.1 Pa | ||||||
Collagen-I 2 mg mL−1 | 1.051 μm | 1.004 μm | 0.208 kPa | — | — | Confocal reflectance microscopy & ImageJ | Indenter (pre-loaded compression followed by retraction until separation, result calculated from slope of experimentally measured force-indentation adjusted area (as AFM)) | 153 |
Collagen 2 mg mL−1 | 17.48 μm | ∼850 nm | 49.59 Pa | — | — | Confocal microscope | AFM | 154 |
Collagen 2.5 mg mL−1 pH = 9 | 10.9 μm | — | — | 61.0 Pa | — | Bubble analysis from confocal reflectance images | Oscillatory shear rheology | 151 |
Collagen 2.5 mg mL−1 pH = 7 | 11.4 μm | 20.1 Pa | ||||||
Collagen 2.5 mg mL−1 pH = 6 | 13.1 μm | 5.7 Pa | ||||||
Collagen 2.5 mg mL −1 | 3.06 μm | 0.73 μm | 90.1 ± 3.0 Pa | — | — | Confocal laser scanning microscope | Colloidal probe force spectroscopy | 155 |
Collagen 2.5 mg mL−1, with covalently intramolecular crosslinking of fibrils by 20 mM EDC (1-ethyl-3-(3-dimethyl-aminopropyl)-carbodiimide) | 3.75 μm | 0.72 μm | 163.6 ± 6.6 Pa | |||||
Collagen-I 3 mg mL−1 (from rat tail) | 1.13 μm | 0.184 μm | 2.16 kPa | — | — | Confocal reflectance microscopy | Indentation | 156 |
Collagen-I 3 mg mL−1 | 1.70 μm | — | 3.15 kPa | — | — | Confocal reflectance microscopy | Indentation | 157 |
Collagen 3 mg mL−1 | 13.0 ± 13.8 μm | — | — | 0.05 kPa | — | Scanning electron microscope (SEM) | Rheology | 129 |
Collagen 3 mg mL−1, in buffer solution of pH 6.5 | — | 86.04 nm | — | 18.94 Pa | 3.84 Pa | Cryo-SEM imaging | Rheology | 158 |
Collagen 3 mg mL−1, in buffer solution of pH 7 | 138.21 nm | 31.82 Pa | 6.22 Pa | |||||
Collagen 3 mg mL−1, in buffer solution of pH 7.4 | 71.67 nm | 23.56 Pa | 4.35 Pa | |||||
Collagen 3 mg mL−1, in buffer solution of pH 7.8 | 71.62 nm | 28.32 Pa | 5.25 Pa | |||||
Collagen-I 3 mg mL−1 (from pure rat tail monomers) | 5.8 ± 0.4 μm | — | 292.9 ± 321.9 Pa | — | — | Confocal laser scanning microscopy | AFM | 152 |
Collagen-I 3 mg mL−1 (rat tail monomers![]() ![]() ![]() ![]() |
5.9 ± 1.3 μm | 326.2 ± 260.1 Pa | ||||||
Collagen-I 3 mg mL−1 (from pure bovine skin monomers) | 5.2 ± 1.6 μm | 141.5 Pa | ||||||
Collagen-I 4 mg mL−1, gelation temperature 22 °C | ∼7 μm | — | — | ∼70 Pa | — | Confocal reflectance microscopy | Rheology | 148 |
Collagen-I 4 mg mL−1, gelation temperature 37 °C | ∼2.5 μm | ∼110 Pa | ||||||
Collagen 4 mg mL−1 (bovine collagen-I (97%) and collagen-III (3%)), pH = 7 | 1.3 μm | — | — | 74.4 Pa | — | Bubble analysis from confocal reflectance images | Oscillatory shear rheology | 151 |
Collagen 4 mg mL−1 | ∼1.4 μm | ∼0.9 μm | — | 101 Pa | — | FEI NOVA nanoscanning electron microscopy | Rheology | 159 |
Collagen 4 mg mL−1 | 10.82 μm | ∼850 nm | 64.39Pa | — | — | Confocal microscope | Elastic modulus by AFM | 154 |
Collagen 4 mg mL−1, gelation temperature 37 °C | 1.17 μm | — | — | 240.26 Pa | — | Confocal laser scanning microscope | Rheology | 160 |
Collagen 4 mg mL−1, gelation temperature 15 °C | 4.37 μm | 332.56 Pa | ||||||
Collagen 6 mg mL−1, gelation temperature 37 °C | 148 nm | 124 nm | 0.783 kPa (DMTA), 0.458 kPa (AFM) | 252 Pa (shear moduli G) | SEM | Dynamic mechanical thermal analysis (DMTA) & AFM for static compression (omit poroelastic effect); rheology for shear moduli | 161 | |
Collagen 6 mg mL−1, gelation temperature starts from 4 °C and heat to 37 °C | 182 nm | 163 nm | 0.718 kPa (DMTA), 0.464 kPa (AFM) | 2064 Pa (shear moduli G) | ||||
Collagen 6 mg mL−1 | 7.22 μm | ∼850 nm | 100.15 Pa | — | — | Confocal microscope | Elastic modulus by AFM | 154 |
Collagen 6 mg mL−1, gelation temperature 37 °C | 1.24 μm | — | — | 397.33 Pa | — | Confocal laser scanning microscope | Rheology | 160 |
Collagen 6 mg mL−1, gelation temperature 15 °C | 4.12 μm | 694.47 Pa | ||||||
Gelatin 6.0 wt%, altering intrafibrillar crosslinking by various methacryloyl groups and/or photoirradiation time | 93.7 nm | 28.0–30.2 nm | 0.5–11.2 kPa | 1.0 kPa–2.1 kPa | 0.006–0.02 kPa | Diffusion-driven transport of 70 kDa FITC-dextran molecules | Cyclic compression experiments using a Mach-1 Mechanical tester; rheology | 162 |
HA 30 mg mL−1, crosslinked | ∼ 6 μm | — | — | 537 Pa | — | FEI NOVA nanoscanning electron microscopy | Rheology | 159 |
Collagen-I 2 mg mL−1 + HA 0.2 mg mL−1, with molecular weight 100–150 kDa (LMW) | 1.134 μm | 1.156 μm | 0.187 kPa | — | — | Confocal reflectance microscopy & ImageJ | Indenter (pre-loaded compression followed by retraction until separation, result calculated from slope of experimentally measured force-indentation adjusted area (as AFM)) | 153 |
Collagen-I 2 mg mL−1 + HA 0.2 mg mL−1, with molecular weight 1.01–1.8 MDa (HMW) | 1.159 μm | 1.187 μm | 0.215 kPa | |||||
Collagen-I 2 mg mL−1 + HA 2 mg mL−1 LMV | 1.158 μm | 1.419 μm | 0.404 kPa | |||||
Collagen-I 2 mg mL−1 + HA 2 mg mL−1 HMV | 1.207 μm | 1.631 μm | 0.628 kPa | |||||
Collagen-I 4 mg mL−1 + HA 4 mg mL−1 LMV | 1.195 μm | 1.685 μm | 0.627 kPa | |||||
Collagen-I 4 mg mL−1 + HA 4 mg mL−1 HMV | 1.165 μm | 2.049 μm | 0.727 kPa | |||||
Collagen 3 mg mL−1 + HA 1 mg mL−1 | 1.92 μm | — | 6.44 kPa | — | — | Confocal reflectance microscopy | Indentation | 157 |
Collagen 4 mg mL−1 + HA 30 mg mL−1 | ∼1.7 μm | ∼0.9 μm | — | 1676 Pa | — | FEI NOVA nanoscanning electron microscopy | Rheology | 159 |
Methacrylated hyaluronic acid (HA-MA) 1 wt% + metacrylate gelatin (GelMA) 0.5% (w/w); with high degree of methacrylation of HA (specific ratios not stated) | 12.90 μm | — | 18.54 kPa | 1187 Pa | 246.64 Pa | Field-emission SEM | Universal testing machine for compressive moduli; rheology for G′ and G′′ | 163 |
Methacrylated hyaluronic acid (HA-MA) 1 wt% + metacrylate gelatin (GelMA) 0.5% (w/w); with middle degree of methacrylation of HA | 14.43 μm | 13.43 kPa | 308 Pa | 100.17 Pa | ||||
Methacrylated hyaluronic acid (HA-MA) 1 wt% + metacrylate gelatin (GelMA) 0.5% (w/w); with low degree of methacrylation of HA | 23.17 μm | 6.05 kPa | 118 Pa | 47.37 Pa | ||||
GelMA 5% (w/v) | Measured after swollen of 1/7/14 days: 310/265/301 μm | — | 5.2 kPa | — | Stress relaxation percentage 8.2% | SEM | Uniaxial compression for elastic modulus and stress relaxation for G′′ | 164 |
GelMA 10% (w/v) | 192/124/199 μm | 55.6 kPa | 5.5% | |||||
GelMA 15% (w/v) | 163/182/229 μm | 161.1 kPa | 6.3% | |||||
GelMA 5 wt% | 7.4 μm | — | 0.034 kPa | — | — | SEM | Rheology for measurement of G′ and G′′ (data not displayed for both) and calculation of E![]() |
165 |
GelMA 10 wt% | 7.3 μm | 1.82 kPa | ||||||
GelMA 15 wt% | 2.6 μm | 2.68 kPa | ||||||
Poly (ethylene glycol) diacrylate (PEGDA)-GelMA, 7.5![]() ![]() |
9.1 μm | 6.18 kPa | ||||||
Poly (ethylene glycol) diacrylate (PEGDA)-GelMA, 1![]() ![]() |
3.6 μm | 2.01 kPa | ||||||
Imine-cross-linked cellulose nanocrystals (CNCs) and gelatin, CCNC-to-Cgel = 0.5, Ctotal = 1.0 wt% | 1.2 μm | 107 ± 35 nm | — | 4 Pa | 0.4 Pa | Pore size by Darcy permeability of hydrogels; diameter of CNC by transmission electron microscopy (TEM) | Rheology | 166 |
Imine-cross-linked cellulose nanocrystals (CNCs) and gelatin, CCNC-to-Cgel = 0.5, Ctotal = 6.0 wt% | 57 nm | 31 ± 14 nm | 254 Pa | 10.5 Pa | ||||
Gelatin 1 wt% + alginate 3 wt% | 0.69 μm | — | 83.16 kPa | — | — | SEM | Compression by electronic universal material testing machine | 167 |
Gelatin 1 wt% + alginate 3 wt% + 2 wt% bioactive nanoparticles (BNPs) | 0.60 μm | 67.89 kPa | ||||||
Gelatin 1 wt% + alginate 3 wt% + 5 wt% BNPs | 0.57 μm | 63.99 kPa | ||||||
Gelatin 2 wt% + alginate 5 wt% | 0.57 μm | 136.59 kPa | ||||||
Gelatin 2 wt% + alginate 5 wt% + 2 wt% BNPs | 0.55 μm | 146.767 kPa | ||||||
Gelatin 2 wt% + alginate 5 wt% + 5 wt% BNPs | 043 μm | 152.906 kPa | ||||||
Alginate 1% w/v + 100 mM Ca2+; sank into 0.3% chitosan solution after gelation | 101.4 ± 48.1 μm | — | 42. 8 ± 3.1 kPa | — | — | SEM | Compression tests | 168 |
Alginate 3% w/v + 10 mM CaCl2 | 23.3 ± 1.2 nm | — | — | 628 ± 101 Pa | — | Mesh size was calculated based on rheological measurement: ![]() |
Rheology | 169 |
Alginate 3% w/v + 40 mM CaCl2 | 13.7 ± 1.0 nm | 3144 ± 686 Pa | ||||||
Alginate 1% w/v + collagen 0.125% w/v + 100 mM Ca2+; sank into 0.3% chitosan solution after gelation | 93.7 ± 46.7 μm | — | 55.3 ± 6.6 kPa | — | — | SEM | Compression tests | 168 |
Phenolated alginate 1.2% w/v + nano-hydroxyapatite 1% w/w | 486 ± 16.72 μm | — | 12.8 kPa | — | — | SEM | Tensile machine | 170 |
Phenolated alginate 1.2% w/v + collagen-I 0.5% w/v + nano-hydroxyapatite 1% w/w | 470.5 ± 17.84 μm | 15.6 kPa | ||||||
Alginate 5 mg mL−1 + collagen 1.33 mg mL−1 + CaCl2 5.0 mM | 45.9 nm | 54.9 ± 10.8 nm— | 0.14 ± 0.01 kPa— | 46.9 Pa | — | Pore diameter is calculated by diffusive transport of FITC labeled dextran; SEM for fiber diameter | Rheology for measurement of G′ and G′′ (data not displayed for G′′) and calculation of E![]() |
171 |
Alginate 5 mg mL−1 + collagen 1.33 mg mL−1 + CaCl2 7.5 mM | 55.1 nm | 108 Pa | ||||||
Alginate 5 mg mL−1 + collagen 1.33 mg mL−1 + CaCl2 10 mM | 40.4 nm | 374 Pa | ||||||
Alginate 5 mg mL−1 + collagen 1.33 mg mL−1 + CaCl2 12.5 mM | 42.5 nm | 665 Pa | ||||||
Alginate 5 mg mL−1 + collagen 1.33 mg mL−1 + CaCl2 15 mM | 50.1 nm | 59.7 ± 12.3 nm | 2.7 ± 0.39 kPa | 902 Pa | ||||
Fibrin, 30 mg mL−1 | 32.2 ± 3.8 nm | — | — | 250 ± 74Pa | — | Mesh size was calculated based on rheological measurement: ![]() |
Rheology | 169 |
IPN (fibrin 30 mg mL−1 + alginate 3% w/v) + 10 mM CaCl2 | 21.3 ± 1.7 nm | 831 ± 205 Pa | ||||||
IPN (fibrin 30 mg mL−1 + alginate 3% w/v) + 40 mM CaCl2 | 15.1 ± 2.3 nm | 2588 ± 1266 Pa | ||||||
Fibrin 10 mg mL−1 + collagen 2.5 mg mL−1 + thrombin 2.5 U ml−1 | 0.4085 μm (median) | — | 2886 Pa | — | — | SEM | AFM | 172 |
Matrigel 50% v/v in DMEM | 3.234 μm | — | — | 8.92 Pa | — | Quick-freeze, deep-Etch electron microscopy | Rheology | 173 |
Matrigel 67% v/v in DMEM | 3.115 μm | 17.76 Pa | ||||||
Matrigel 75% v/v in DMEM | 2.986 μm | 32.33 Pa | ||||||
Matrigel 25% v/v + alginate 16 mg mL−1 + CaCO3 174 mM | 51.2 ± 2.8 μm | — | 2.66 ± 0.84 kPa | 0.85 kPa | 0.18 kPa | SEM | Rheology for measurement of G′ and G′′ and calculation of E![]() |
174 |
Matrigel 25% v/v + alginate 28 mg mL−1 + CaCO3 453 mM | 38.4 ± 15.9 μm | 8.98 ± 1.29 kPa | 2.90 kPa | 0.58 kPa | ||||
Matrigel 25% v/v + alginate 40 mg mL−1 + CaCO3 500 mM | 32.3 ± 9.1 μm | 18.27 ± 3.17 kPa | 6.02 kPa | 0.67 kPa | ||||
Decellularized porcine arterial ECM-derived hydrogel 30 mg mL−1 (after lyophilized, liquefied, and enzymatic digestion) | Distribution: <5 μm (∼50%); 5–50 μm (∼50%) | Distribution: 1–2.5 μm (∼67%); 2.5–5 μm (∼30%); <1 μm (∼1.5%); 5–10 μm (∼1.5%) | 3.47 ± 1.3 kPa (spatial elastic modulus) | ∼8 kPa | 1.4–1.8 kPa | SEM | Spatial elastic modulus by AFM; rheology | 175 |
Chitosan 4% w/w | 4.96 nm | — | 110.51 ± 19.33 kPa | — | — | SEM | Tensile stress–strain tests | 176 |
The tetramer peptide IIZK (Ac-Ile-Ile-Cha-Lys-NH2) 1 mg mL−1 in 1× PBS | — | 9 nm–16 nm (no difference for the two concentrations) | — | 7.3 kPa | 1.1 kPa | SEM and Cryo-TEM | Rheology | 177 |
The tetramer peptide IIZK 10 mg mL−1 in 1× PBS | 139.7 kPa | 12.2 kPa | ||||||
IKVAV appended with 9-Fluorenylmethoxycarbonyl modification (Fmoc IKVAV) 20 mM; (YIGSR and IKVAV are laminin mimetic pentapeptides) | — | 30 ± 5 nm (AFM); 15 ± 2.2 nm (TEM) | — | ∼267 ± 47 Pa | — | Fiber diameter by AFM or TEM | Rheology | 178 |
Fmoc YIGSR 20 mM | 20 ± 6 nm (AFM); 11.5 ± 1.2 nm (TEM) | ∼674 ± 100 Pa | ||||||
Fmoc IKVAV 10 mM + Fmoc YIGSR 10 mM | 85 ± 14 nm mix with 32 ± 6.5 nm (AFM); 28.3 ± 2.5 nm mix with 9.5 ± 1.9 nm (TEM) | ∼937 ± 124 Pa | ||||||
4-arm PEG maleimide 5 mM (10 kDA) + MMP-sensitive crosslinker 10 mM | Median pore diameter 13.3 ± 14.1 μm | — | — | 1.4 ± 0.3 kPa | — | SEM | Rheology | 129 |
4-arm PEG maleimide 5 mM (10 kDA) + 4-arm PEG thiol 3.33 mM (10 kDA) + MMP-sensitive crosslinker 3.33 mM | 10.7 ± 11.0 μm | 7.9 ± 0.7 kPa |
Collagen mesh size can be predicted by the equation from Jansen et al.:179ξ = (1/ρl)0.5, with ξ the average mesh size, ρl the length of collagen fibers per unit volume defined as ρl = cp/μ, with cp the concentration of collagen, and μ the average mass–length ratio of the fibers. Fiber diameter d and μ both decrease with increasing cp or increasing temperature. These parameters can be calculated using the turbidity–wavelength relationship of collagen fibrous gels, as measured by light scattering. However, it is important to note that collagen sourced from different species and different extraction and purification processes have different polymerization times, mechanical properties, and structures, even under the same polymerization conditions. The most prevalent collagen I sources are bovine dermis,180 murine tendon,180 and porcine skin.181 Collagen can be extracted through acid or enzymatic (pepsin) treatment. Pepsin-treated collagen self-assembles into multimeric fibrils more slowly, resulting in a larger mesh size (3–5 μm) and longer fibrils. In contrast, acid-extracted collagen, even with the same protein content, forms smaller pores (1–2 μm) and shorter fibrils.182
Similarly to collagen, fibrin pore size depends on fibrinogen and thrombin concentrations. Wufsus et al.183 estimated the fibrin network mesh size ξ from the square root of hydraulic permeability Ks (ξ = Ks1/2) or from measurements of shear modulus , and the results of the two methods had good agreement. The mesh size decreased from 300–400 nm to 20–30 nm when the fibrinogen concentration increased from 3 mg mL−1 to 100 mg mL−1, and the most obvious change occurred over the range of 3–30 mg mL−1. These findings were confirmed by other studies.184–186
Pore size has a direct effect on cells by inducing mechanical confinement, but also an indirect effect by influencing nutrient diffusion (Fig. 2A).121 The intrinsic pore size of hydrogels has been shown to significantly influence the migratory behavior of cells and alter cellular fate decisions. The nanometer scale meshes of synthetic polymer networks confine cells, making cell migration or morphological change nearly impossible without deforming or degrading the hydrogel.54 Rigid matrices with mesh sizes less than 10% of nuclear cross section (≤7 μm2) were reported to be able to block the squeezing through of cell nuclei.187 Frequent squeezing through small intrinsic pores comparable in size to a cell or its nucleus and characterized by high stiffness may result in DNA damage.188
![]() | ||
Fig. 2 Relation between hydrogel structure and cell behavior. (A) When the pore size is smaller than the nucleus size, cells are confined within the pore and can't migrate or spread. When the pore size is larger than the nucleus, cells can squeeze in and elongate. Human mesenchymal stromal cells were encapsulated in either a hyaluronic gel (HA-gel Hydrogel) with a nanopore size (a) or a cell encapsulatable cryogel (HA-gel CECG) with macropores (b). Cells in the HA-gel expressed a low level of actin, while cells in the HA-gel CECG showed a high level of actin and were able to spread. The increased pore size also enabled cell growth, as indicated by the presto blue assay (c) (reproduced from ref. 189 with permission from IOP Publishing, copyright 2019). (B) Fiber orientation can guide cell orientation. Endothelial cells were embedded in collagen gel with preferential or aligned collagen fibers. When fibers were randomly oriented, cells were also randomly oriented (a). When fibers were aligned in a preferential direction (anisotropy), cells aligned in that direction (b) and formed longer vessels (c) (reproduced from ref. 190 with permission from American Chemical Society (ACS), copyright 2018). |
Non-fibrous anisotropic hydrogels can be created by including rod-shaped microgels in a surrounding hydrogel. Microgels containing magnetic particles embedded in a PEG, fibrin, or collagen matrix have been successfully oriented using a magnetic field. The effect on cell growth and orientation depends on the differences in mechanical properties between the microgel and the embedding gel, the density of microgels, their dimensions, and their interactions with the embedding gels.197–199 These approaches illustrate how microscale anisotropy within hydrogels can influence cell behavior.
Similarly, hydrogels with pre-aligned fibers have deepened our understanding of anisotropy on cell behavior (Fig. 2B). They have notably provided insights into cell migration in anisotropic environments, which is a hallmark of the tumor environment.116,191,192 Aligned collagen matrices promote cell protrusions along the fibers, accelerating and directing cell spreading and migration in the direction of the fibers. This behavior depends on focal adhesion kinase localizations and Rac1 activity, as their inhibition eliminated protrusion anisotropy in the aligned matrices.191
Anisotropy within hydrogels not only influences cell migration but also mediates the organization of the surrounding matrix. Cell contractile forces cause a denser matrix immediately surrounding the cells151 and stiffen the local collagen fiber network at the leading edge.200 Cells can also align the surrounding fibers by exerting traction forces along their long axis.194,201 This tension-driven collagen fiber alignment enhances the effective transmission of cellular forces, increases cellular stiffness, and drives further matrix stiffening.202,203 Importantly, cell elongation is determined more by the organization of surrounding collagen fibrils than by the overall stiffness of the matrix,194 highlighting the intricate relationship between matrix structure and cellular mechanics.
![]() | ||
Fig. 3 Influence of mechanics on cell behavior. (A) When the polymer network embedding the cells cannot be deformed by the cells, cells are confined and cannot spread or migrate. Fibroblasts in elastic alginate remain round while those in viscoelastic alginate can spread (reproduced from ref. 133 with permission from Nature Publishing Group, copyright 2016). (B) Viscoelasticity can influence cell differentiation. Shorter relaxation times increase stem cell differentiation into osteoblasts, as visualized by the increased ALP staining (reproduced from ref. 133 with permission from Nature Publishing Group, copyright 2016). (C) Paratensile signaling is a communication mode in which matrix fibers transmit mechanical strain between cells. This results in collagen alignment between neighbor cells (reproduced from ref. 205 with permission from Elsevier, copyright 2022). Nonlinear elasticity increases this effect as strain stiffening extends the strain field generated by cell contractility.203 |
As introduced in section 3.2.1, increasing the gelation temperature of collagen hydrogel leads to decreased pore size and fiber diameter.179,215 Control of both temperature and polymer concentration can modify collagen hydrogel stiffness without altering the microstructure. For instance, Yang et al.148 maintained similar pore sizes (5–7 μm) while increasing collagen I concentration from 1 to 4 mg mL−1, leading to a rise in G′ from 2 Pa to 110 Pa by adjusting gelation temperature (22 °C to 37 °C). Similarly, Seo et al.161 showed that gradually increasing gelation temperature from 4 °C to 37 °C for 6 mg mL−1 collagen increased pore size and fibril diameter by 20–30% while it boosted stiffness sevenfold. Changes in pH also affect collagen hydrogels: higher pH increases fibril length, decreases fibril diameter, and raises stiffness.216 For example, Sun et al.151 demonstrated that G′ rose from 5.7 Pa to 60.0 Pa as pore size decreased from 4.9 to 1.7 μm at constant collagen concentration when pH increased from 6.0 to 9.0.
Most methods to tune the viscoelasticity focus on the polymer phase of the material by changing either the molecular weight of the polymer or the nature of the crosslinks. Decreasing the molecular weight decreases the half-stress relaxation time (τ1/2), meaning that the viscous behavior of the material increases.88,218 This method has been notably applied with alginate133,219 and HA.220 For example, when the average molecular weight of alginate decreased from 280 kDa to 35 kDa, the τ1/2 decreased fourteen-fold.219 In a study on HA-collagen interpenetrating networks, reducing the average molecular weight of HA from 120 kDa to 20 kDa resulted in a thirty-seven-fold decrease in τ1/2, whereas the molecular weight of pure HA hydrogels showed minimal impact on τ1/2 under the same conditions.220 Another option is to modify the type of crosslinks. Ideal static covalently crosslinked polymer networks result in an elastic material that does not exhibit viscoelasticity.221 In contrast, non-ideal incomplete covalent crosslinking polymers and loose ends exhibit energy dissipation, thus viscoelasticity.88,222 Alternatively, weak crosslinks (noncovalent physical crosslinks or dynamic covalent crosslinks) can stabilize the polymer network, giving rise to viscoelastic materials.21,218 For instance, a covalently crosslinked alginate network exhibits an elastic behavior, while ionically crosslinked alginate is viscoelastic.222,223 The introduction of reversible (dynamic) covalent crosslinking can also generate mechanically stable hydrogels with viscoelastic behavior.224 This type of crosslinking has been used successfully with chitosan,225 HA,220 PEG,226–229 and alginate.230 Additionally, it is possible to covalently couple different materials, such as using PEG spacers with alginate chains to decrease the τ1/2 of the alginate hydrogel, independently of the initial elastic modulus.133,219
Although most methods to tune viscoelasticity focus on the solid phase polymers, it is possible to tune the viscoelasticity by modifying the viscosity of the aqueous phase. Adding dextran, a homo-polysaccharide of glucose, to the aqueous phase of agarose or polyacrylamide reduces the relaxation time (τ) and the instantaneous elastic modulus, which represents the initial elastic response. However, it maintains similar equilibrium moduli, which represent the static response after viscoelastic relaxation during compression tests.231,232 This is caused by the interaction between dextran and the hydrogen bonding between water and the polymers. As less water is available to bind to the polymer chains, more water can easily flow between the polymer chains, thereby reducing the relaxation time and instantaneous modulus but maintaining the equilibrium modulus.231
Tuning the relaxation behavior of the hydrogels independently from the initial stiffness has unraveled the effect of each parameter on cell behavior. Decreasing the relaxation time (faster relaxation hydrogels) while maintaining the initial elastic modulus enhanced cell spreading (Fig. 3A)133,233,234 and osteogenic differentiation (Fig. 3B) of mesenchymal stromal cells.133,234 This effect was mediated through integrins and actomyosin contractility, but not through traditional focal adhesions.133
Viscoelasticity impacts individual cells, but also organoid development and differentiation. Hydrogels with different viscoelastic properties influence the growth, fusion, and matrix secretion of organoids from various tissues such as cartilage,235 kidney,236 intestine, and breast epithelium.237 Faster relaxation rates seem to improve cartilage and intestine organoid growth and differentiation, but more studies are required to draw general conclusions on the effect of viscoelasticity on organoids.
Strain-stiffening properties are seldom observed in synthetic hydrogels.96 One of the few strain-stiffening synthetic polymers is PIC, which shows a low critical stress σc of tens of Pa and a stiffening index m of about 3/2, similar to biopolymer hydrogels.240,241 Keeping PIC concentration constant, increasing polymer chain contour length L will increase the plateau modulus G0 and critical stress σc with G0∝L2 and σc∝L,242 while decreasing m,241 resulting in stiffer and less stress-sensitive gels. Addition of an increased amount of peptide sequence GRGDS to PIC polymers will decrease G0 and σc exponentially, thus the strain-stiffening responsiveness is higher.243
Nonlinear hydrogels have improved our understanding of cell–cell mechanical communication. Nonlinear elasticity increases cell–cell mechanical communication via paratensile signaling, a communication mode in which mechanical strain is transmitted between cells via matrix fibers35 (Fig. 3C). This effect is enhanced by strain stiffening as it extends the strain field generated by cell contractility244 and thus promotes long-range mechanical signaling.202,245 The importance of strain-stiffening for mechanical communication depends on the cell type. For neurons, the strain-stiffening is not obvious considering the lower level of stress that neurons can exert on the surrounding matrix. For cells that can induce rather large stresses through contractility such as fibroblasts,246 hMSCs,245,247 or breast cancer cells,202,248 the effects of strain-stiffening become obvious. This phenomenon is critical in the context of fibrosis. Myofibroblasts embedded in a collagen gel can generate force, which is then transmitted through fibrous collagen, and activate quiescent fibroblasts embedded in the same gel, thus propagating fibrosis.249
Photopolymerization is a common technique for manufacturing photoresponsive hydrogels.252 Photopolymerization can induce a second crosslinking step to increase stiffness135,253 or degrade the gel to reduce stiffness.254,255 Recently, PEG–alginate hydrogels have been used to tune the relaxation rate of the gels over time via photopolymerization of the PEG network.256 Incorporation of PEG molecules in the alginate network decreased the τ1/2. Though the τ1/2 was still higher than the characteristic τ1/2 of most biological tissues (∼10 s), the authors produced gels with relaxation times varying from 80 to 800 seconds. As this method is compatible with cell encapsulation, the relaxation time of the gels can be tuned during cell culture. Similarly, norbornene–HA hydrogels allow spatiotemporal control over stiffness and relaxation rate upon photocrosslinking.257 Despite these promising results, it is important to note that photocrosslinking can be toxic for cells because of the photoinitiator used and the type of light required.258
Temperature- and pH-responsive hydrogels have been widely explored for regenerative medicine applications. For example, injectable hydrogels designed to gelate at physiological temperatures (20–37 °C)259,260 or specific pH conditions261,262 have been developed to encapsulate cells for in vivo treatments, such as heart tissue regeneration or stem cell delivery. While these applications are promising, they fall outside the scope of this review, which focuses on hydrogels as ECM mimics for studying cell behavior in vitro.
Other environmental stimuli include magnetic fields and hydration. Magnetic fields can tune the stiffness and induce anisotropy, guiding cell alignment,263,264 while 4D bioprinting makes use of the degree of hydration to induce bulk shape.265
Hydrogels can also respond to cellular stimuli, such as MMP secretion266 or traction forces. For example, predetermined patterns of mesenchymal cells can dynamically fold collagen substrates into certain desired patterns, such as hollow tube structures. The folded structure generates vessels when the substrate is seeded with HUVEC.267
Recently, Major et al. have implemented a hydrogel showing a progressively stiffening of compressive modulus for about 10 to 15 kPa without requiring external stimuli during the 3 week culturing time. The hydrogel is composed of adipose-derived ECM and silk fibroin, which can be initially photo-crosslinked by visible light and added photoinitiators. The stiffening is due to the spontaneous formation of β-sheet secondary structures through hydrogen bonding in silk fibroin after crosslinking and during the culture time.268
IPNs can also combine ECM-derived polymers with polymers lacking cell adhesion sites, enabling tuning the mechanical properties by modifying nonfibrous components without changing protein concentration or the structure of fibrous protein components. However, the combination of different polymers changes the bulk behavior of the material. A well-known type of IPN is alginate–collagen IPN, in which collagen fibrils network intercalate with alginate mesh. Ionic crosslinking of alginate keeps the IPN fibrous architecture and mesh size unchanged. Low concentration (<2 mM) or no Ca2+ for crosslinking alginate leads to IPN storage modulus close to pure collagen hydrogel of the same concentration, and increasing Ca2+ concentration from 2.5 mM to 10–15 mM leads to stiffness increase from tens of Pa to about 1 kPa.171,272 Another example is IPN gels made of collagen and GAG, notably hyaluronic acid. Although HA has multiple cell receptors, a study showed that its effects are similar to that of inert alginate when combined with collagen, proving that the observed effects were mechanical rather than biochemical.273 Compared with a pure collagen gel of the same concentration, adding HA to the collagen network showed a two- to four-fold increase in E153,156,157 or a fifteen-fold increase in G′,159 while maintaining approximately the same pore diameter and fiber diameter. However, adding other GAG (eg, chondroitin sulfate, dermatan sulfate) to collagen to form hydrogel did not influence stiffness.156 Incorporating HA in collagen gels reduced cell-traction forces and cardiomyocytes alignment.273 Additionally, HA in collagen gels also reduces plastic deformation.274
The combination of two distinct polymer networks, where the fibrous component is not derived from ECM proteins, can also create IPNs. For instance, Chen et al.162 developed a fibrous hydrogel using aldehyde-modified cellulose nanocrystals (a-CNCs) as the structural component and incorporated gelatin methacryloyl (GelMA) for intrafibrillar photocrosslinking without affecting interfibrillar crosslinking among a-CNCs. By increasing the degree of GelMA crosslinking—either through higher methacryloyl group content or prolonged photoirradiation—G′ was enhanced from several kilopascals to over 10 kPa, while maintaining consistent a-CNC and GelMA concentrations, as well as pore and fibril diameters. This demonstrates how tuning intrafibrillar crosslinking can significantly modulate hydrogel stiffness without altering other structural parameters.
While IPNs generally result in homogeneous hydrogels at the cell scale, heterogeneous hydrogels can be created using large inclusions of sacrificial materials to form large void structures. These templating methods involve using a pre-existing structure or template to guide the creation of desired porosity in the final hydrogel material.142 One common technique involves incorporating solid particles (porogens) into the hydrogel precursor, which are then removed after crosslinking to create larger pore sizes, ranging from tens to several hundreds of micrometers.143,275,276 Gelatin is a good porogen as it is biocompatible, dissolves at 37 °C, and is soluble in water. Incorporating gelatin beads allows for tuning pore size independently of porosity. This method has been successfully used with PEG,277 alginate,278 and various photocrosslinkable bioinks.276 This principle can be extended to any cytocompatible sacrificial material. For example, lyase-digestible alginate and alginate resistant to lyase digestion have been patterned to create microchannels after digestion.279
The inclusion of larger pores by using sacrificial material allows us to uncouple the effects of porosity from pore size on cell behavior. Porosity (26% vs. 65%) influenced cell growth in PEG gels with pores created by leaching gelatin, while pore size (122–233 μm) had no effect. Similarly, porosity, but not pore size, influenced cell spreading.277 Osteoblast-like cells cultured in bioprinted constructs also showed enhanced proliferation with 40% porosity and pore sizes of the cell scale (40 μm). Furthermore, matrix mineralization appeared more uniform in the porous constructs than in the bulk hydrogels after 14 days.276 Larger porosities forming channels inside the hydrogel enable the connection to a perfusion system, which supports cell proliferation and tissue formation.279 Thus, porogens can be used to enhance nutrient diffusion and waste removal to compensate for the reduced diffusion in the core of hydrogels.
![]() | ||
Fig. 4 Interplay between structural and mechanical properties. (A) Structural properties and mechanical behavior of hydrogels are interdependent and interrelated with no simple one-to-one correspondence. For example, increasing the crosslinkers will increase stiffness but also reduce the intrinsic porosity of the polymer network. The presence of fibers confers a nonlinear behavior, while porous hydrogels have a more linear behavior. Nonlinearities and crosslinking types both influence the relaxation behavior. Fiber anisotropy leads to mechanical anisotropy with higher stiffness in certain directions. (B) Collagen and fibrin exhibit strain-stiffening, but not agarose or polyacrylamide. (C) Strain-stiffening and viscoelasticity in natural hydrogels influence each other. Crosslinking partially cancels this effect. (B) and (C) are reproduced from ref. 94 with permission from National Academy of Sciences, copyright 2016. |
Both intrinsic pore size (or mesh size) and mechanical properties influence cell migration. If the mesh size is larger than the cell nucleus, cells can move freely. If the mesh size is smaller, cells must either degrade or deform the matrix. Cells can migrate through a matrix that is plastic enough and of nanometer-scale intrinsic pore size without using proteases, by using protrusions to open up micrometer-scale channels for migration.281 If the gels are elastic and too stiff for the cells to deform, this creates a mechanical confinement for the cells, which negatively impacts various biological processes.230,282,283 Viscoelasticity also enables cell spreading at constant stiffness and pore size,133,284 as previously mentioned in section 3.3.2. However, increasing mesh size decreases stiffness,54,207–214 making it difficult to separate stiffness from mesh size. Thus, while mesh size influences overall stiffness, factors such as stiffness, viscoelasticity, and plasticity also affect how intrinsic pore size impacts cells. It is worth noting that while this is true for the intrinsic porosity (mesh size), the impact of larger pores created by including porogens is different: in this context, porosity but not pore size influences the bulk stiffness. Besides, the local stiffness (cell scale) remains unchanged.277
In addition to pore size, structural anisotropy also impacts mechanical properties as structural anisotropy leads to mechanical anisotropy. Consequently, the effects of these two properties are difficult to decouple. For collagen hydrogel with the same concentration, anisotropically oriented collagen fibers enhance the hydrogel elastic modulus along the fiber orientation compared to hydrogels with isotropically oriented fibers.285 The anisotropic distribution and stiffness of matrix fibers stabilize cell protrusions in the direction of the alignment and promote cell migration,280 as well as fibroblast to myofibroblasts transition.286 However, increased elastic modulus alone through crosslinkers does not promote cell elongation.280
Besides the reciprocal influence between mechanical and structural properties, mechanical properties can also interact and influence each other. For instance, Nam et al.94 demonstrated the interplay between relaxation and strain-stiffening in fibrous hydrogels. Fibrin and collagen exhibit both strain-stiffening (Fig. 4B) and viscoelasticity (Fig. 4C). The higher the initial strain, the faster the stress relaxation, leading to a rapid decrease in the strain-stiffening effect over time. Numerical simulations suggested that force-dependent unbinding followed by rebinding of fibers can explain stress-enhanced stress relaxation in collagen networks. Covalent crosslinking reduces this effect. Thus, increasing collagen stiffness through transglutaminase or glutaraldehyde crosslinking also alters the material's viscoelastic and nonlinear response. With the aid of magnetic microrheometry and a probabilistic modeling approach to enhance the analysis of the sparse probe-generated data, Arasalo et al.246 recently reported an enhanced measurement and mapping of the spatial heterogeneity of viscoelasticity and stiffness in a 3D collagen hydrogel embedded with cancer-related fibroblasts. Their study observed collagen stiffening and bundle formation in the collagen matrix in the vicinity of cells, and a decrease in the collagen phase angle, indicating a more elastic behavior, which relates to the stiffened collagen bundle formation by cell force.
It is important to note that while stiffness is commonly reported, other mechanical properties (such as viscoelasticity and nonlinearity) and structural properties (such as pore size and the presence of fibers) are rarely mentioned or discussed. This may explain why some studies report conflicting results regarding the effect of stiffness on cell behavior. For example, Hadjipanayi et al.287 observed 3.5 times more proliferation of fibroblasts over 2 days in collagen gels at 143 kPa compared to gels at 42 kPa. Conversely, Shie et al.288 found twice as much proliferation of fibroblasts in 50 kPa GelMA gels compared to 139 kPa gels. The differences in the viscoelastic behavior of these materials may explain such opposite observations; collagen exhibits viscoelastic behavior, whereas GelMA is more elastic. Chaudhuri et al.133 proposed a mechanism to explain enhanced cell spreading and proliferation in viscoelastic matrices. A cell embedded within a three-dimensional matrix initially imparts strains on the matrix, leading to forces and stresses that resist this strain. In purely elastic matrices, these forces remain constant, preventing any remodeling of the matrix microenvironment. In viscoelastic matrices, however, forces within the matrix can gradually dissipate over time. The extent of mechanical remodeling depends on the rate of stress relaxation. In matrices with rapid relaxation, this process facilitates adhesion-ligand clustering, changes in cell shape, and proliferation.
Differences in ligand densities can also explain discrepancies between studies. Increasing collagen density to raise the elastic modulus also increases ligand density, while increasing the crosslinking of GelMA does not affect it. Moreover, gelatin, unlike collagen, has lost its ability to form fibers, which also impacts cell behavior.289 Thus, several factors can explain the contrasting cell growth rates observed in these studies.
Besides, the next generation of hydrogels should involve strategic combinations of multiple polymers to better replicate physiological conditions and achieve more precise control over these variables. Because of the aforementioned intertwining of structural and mechanical factors, designing such hydrogels with desired properties typically requires laborious experiments with trial and error. The recent rapid development of artificial intelligence (AI) and its subfield machine learning (ML) has already allowed more efficient optimization of parameters for 3D and 4D printing of hydrogel.290–292 Such approaches can also be further developed as promising tools to advance the design of functional hydrogels,293–295 possibly via the strategy of simultaneously combining hydrogel compositions and structural properties to predict their influence on the mechanical properties of multi-component hydrogels, in combination with cellular mechanobiology to address their complex impact on “what cells see” and “what cells feel”.
This journal is © The Royal Society of Chemistry 2025 |