Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

REvIEWS

The stiffness of living tissues and its


implications for tissue engineering
Carlos F. Guimarães   1,2,3, Luca Gasperini   1,2, Alexandra P. Marques   1,2,3 and
Rui L. Reis   1,2,3 ✉
Abstract | The past 20 years have witnessed ever-​growing evidence that the mechanical
properties of biological tissues, from nanoscale to macroscale dimensions, are fundamental
for cellular behaviour and consequent tissue functionality. This knowledge, combined with
previously known biochemical cues, has greatly advanced the field of biomaterial development,
tissue engineering and regenerative medicine. It is now established that approaches to engineer
biological tissues must integrate and approximate the mechanics, both static and dynamic,
of native tissues. Nevertheless, the literature on the mechanical properties of biological tissues
differs greatly in methodology, and the available data are widely dispersed. This Review gathers
together the most important data on the stiffness of living tissues and discusses the intricacies
of tissue stiffness from a materials perspective, highlighting the main challenges associated
with engineering lifelike tissues and proposing a unified view of this as yet unreported topic.
Emerging advances that might pave the way for the next decade’s take on bioengineered tissue
stiffness are also presented, and differences and similarities between tissues in health and disease
are discussed, along with various techniques for characterizing tissue stiffness at various
dimensions from individual cells to organs.

Materials engineering involves the design or tuning of questioned. Currently, when considering the stiffness of
1
3B’s Research Group, the structure of a material to produce a set of desired engineered constructs comprising cells on a 3D support,
Research Institute on properties1. Living tissues are essentially materials engi­ one must think not only of its biomechanical integration
Biomaterials, Biodegradables neered by nature itself to have a specific structure that with host tissue but also of cell mechanostimulation.
and Biomimetics, University
affects cell properties and drives all consequent biolo­g­ From adhesion to differentiation, the most relevant
of Minho, Headquarters of
the European Institute
ical events. Tissue engineering tries to replicate this feat cellular events from early embryogenesis onwards4 are
of Excellence on Tissue by designing the structure of a material to recapitulate a all, in some way, affected by tissue stiffness. As such,
Engineering and Regenerative predetermined cell response. any modern tissue engineering strategy has to consider
Medicine, Guimarães, When tissue engineering began, more than 25 years the stiffness of the native biological tissue, as well as the
Portugal.
ago, ‘stiffness’ was already a term mentioned in the liter­ implications of this stiffness for the behaviour of resi­
2
Life and Health Sciences
ature2. In the beginning, the premise was simple: a tissue dent cells. These factors require not only bulk material
Research Institute,
3B’s Research Group, substitute must be biomechanically able to fulfil the analysis but also an understanding of the microscale and
Research Institute on functions of the tissue it replaces and, as such, should nanoscale domains within which the native extracellular
Biomaterials on Biomaterials, have mechanics similar to those of the native tissue. matrix (ECM) components function in vivo. Moreover,
Biodegradables and Hard scaffolds began to be used to engineer bone-​like cells themselves have a characteristic stiffness, which
Biomimetics, Portuguese
Government Associate
structures and hydrogels for soft tissue structures, which is a consequence not only of their interaction with the
Laboratory, University were combined with specific biochemical cocktails that surrounding microenvironment but also of their bio­
of Minho, Braga and could direct the behaviour of cells on these structures, logical and/or genetic status. For example, cancer cells
Guimarães, Portugal. which simply served the purpose of support. However, frequently have altered (either increased or decreased)
3
The Discoveries Centre for this paradigm shifted with the understanding that the stiffness5. If tissue stiffness was initially considered a
Regenerative and Precision stiffness of surfaces, independently of their shape, could simple concept, it is now undeniably worthy of thorough
Medicine, University of Minho,
Guimarães, Portugal.
be detected and responded to by cells3. and integrative study.
✉e-​mail: The ripples of this discovery are still reverberating The Review brings together these topics using an
rgreis@i3bs.uminho.pt throughout diverse fields of research, but it truly revo­ approach that aims to be as interdisciplinary as the field
https://doi.org/10.1038/ lutionized tissue engineering as the supremacy of bio­ of tissue engineering itself. After a brief overview of what
s41578-019-0169-1 chemical cues in driving cellular behaviour began to be stiffness is and how the mechanical properties of living

Nature Reviews | Materials


Reviews

tissues are measured, we consider each component in moduli in the 0.2–5.0-GPa range, higher than for all
turn, starting with the smallest — ECM molecules and biological tissues except bone. By contrast, the stiff­
single cells — and continuing up to bulk living tissues ness of some foodstuffs is comparable to that of tissues:
and organs, which are compared and discussed from a the elastic moduli of panna cotta range from 100 Pa to
biomaterials perspective. Finally, we outline the main 1 kPa (refs7,8), the elastic moduli of gummy bears and
avenues to follow when attempting to engineer life­ bananas are between 50 and 100 kPa, an apple has an
like tissues, emphasizing the most recent advances in elastic modulus of about 1 MPa and a carrot has an e­ lastic
time-​changing mechanics and identifying the main modulus of around 7 MPa (refs9,10). Similarly, from a bio­
challenges for the future — the questions that must be logical perspective, fat is clearly far softer than ­muscle
answered to allow tissue stiffness to be fully decoded and tissue, which itself is far softer than bone. Naturally,
manipulated. the stiffness of a tissue can be quantified and precisely
analysed but, despite the term being widely mentioned
The concept of stiffness in the biomedical field, analyses of tissue stiffness are
The mechanical properties of a material — most nota­ sometimes insufficiently detailed.
bly its stiffness — relate to loads and deformations; that Stiffness is a general structural property that depends
is, the forces exerted on the material and the resulting not only on the material itself but also on its amount
changes in its shape. To fully understand the stiffness of and distribution (shape). For example, the hollow
living tissues, one has first to explore these underlying nature of bones confers an increased stiffness-​to-weight
concepts. The stiffness of a structure derives from the ratio. During compression or extension (stretching) of
following two premises: first, that when a structure is a material, the whole cross-​sectional area of the mate­
exposed to a certain load, it will deform; and second, that rial equally sustains the stress, whereas during bending
the ratio between this load and the consequent deforma­ or torsion, the material furthest from the midpoint or
tion yields the stiffness of the structure, meaning how centre line sustains most of the stress. For this reason,
much load is necessary to achieve a certain deformation6. structural construction elements often have T-​shaped or
Although there is no absolute definition of what consti­ L-​shaped cross-​sectional shapes, which maximize their
tutes a ‘stiff ’ or ‘soft’ material, rubber and foodstuffs are stiffness while minimizing the weight and amount of
generally considered soft materials, wood and plastics material used. Moduli and stiffness are intimately related
are of intermediate stiffness, and steel is among the stiff­ concepts and therefore these terms are frequently used as
est commonly encountered materials. Clearly, the load synonyms. However, stiffness is a property of a structure,
necessary to deform a stiff material, such as steel, will be whereas moduli describe the properties of the material
far higher than that required to similarly deform a softer composing that structure. As such, various moduli relat­
material, such as wood or rubber. ing to the intrinsic elastic properties of materials6 are
The stiffness of materials commonly used in manu­ reflected in the stiffness of the final structure. These
facturing is far higher than that of human tissues or moduli are derived from mathematical conversions of
organs, including materials used specifically because of load versus deformation relationships obtained from
their deformability, such as the plastics used in ketchup standardized tests on samples of a standardized size and
bottles. The vast majority of industrial plastics (such shape (Box 1).
as polypropylene, nylon or polyethylene) have elastic For example, Young’s modulus (E) can be calculated
by subjecting a material to uniaxial stress resulting from
compression or extension and measuring elastic (that is,
Box 1 | standards in mechanical testing reversible) deformation (strain) in the linear region of
the stress–strain curve:
standardized protocols must be followed in mechanical analyses if the data so obtained
are to be interpreted with a good level of confidence and are to be comparable between σ (ε ) F ∕A FL 0
different investigational studies. accordingly, in the past decade, there has been a E≡ = = , (1)
great effort to introduce standardized methods for the analysis of tissue-​engineered ε ΔL ∕L 0 AΔL
products, such as the astM standards developed by astM Committee F04 and related
subcommittees, which aim to improve the safety, quality and consistency of these where σ is uniaxial stress (force per unit surface), ε is
products. although these standards were not specifically developed for the purpose strain, F is the force exerted on an object (uniaxial stress),
of performing mechanical tests on human tissues, they can provide guidance for A is the cross-​sectional area perpendicular to the applied
appropriate testing. For example, astM F561-19 covers recommendations for the force, ΔL is the amount by which the length of the object
handling and analysis of post-​mortem tissue and living tissue samples surgically changes (ΔL has a positive value for a stretched material
removed from humans and animals, whereas astM F2150-13 provides guidance on
and a negative value for a compressed material) and L0 is
the selection of appropriate test methods for analysing the bulk physical, chemical,
mechanical and surface properties of such samples. astM D638-14, a standard for
the original length of the object. Thus, the axial stiffness
testing polymeric materials under uniaxial extension, states that it might not be (k) of a longitudinal structure such as a beam can be
meaningful to compare the results of tests performed with widely different loads or calculated as
over widely different timescales. although biological systems differ substantially from
the plastics referred to in astM D638-14, when taking a cautious approach, we might EA
k= . (2)
reasonably extend these concepts to human tissues. Overall, however, the tissue L0
engineering field seems to be reluctant to adopt similar standards, or even to attempt
to follow existing standards. in most published reports, mechanical tests were either Equations 1 and 2 assume a linear relationship between
not performed according to any consensus standard, or if they were, the standard strain and stress (that is, Hooke’s law). In real-​life sce­
followed is rarely referenced.
narios this assumption might not hold for all levels of

www.nature.com/natrevmats
Reviews

Fig. 1 | Main mechanical deformations and representative curves. The main types of mechanical analysis include
tensile (part a), compressive (part b), shear (part c) and torsion (part d) deformations. e | Typically , static deformations
(such as tensile or strain deformations) yield a complex curve within which distinct regions can be identified, each of
which has important mechanical correlates. Young’s modulus (E) is calculated as the slope of the stress-​versus-strain
curve in the linear (elastic) region. With higher levels of strain, the material enters the plastic domain of the curve, where
deformations are no longer reversible and the material deforms permanently until fracture occurs. Alternatively , a
dynamic analysis can be performed, in which a strain within the linear (viscoelastic) region is applied repeatedly over
time, in cycles often with changes in frequency or temperature, to yield storage (E′) and loss (E″) moduli. The storage
modulus is related to elastic deformation of the material, whereas the loss modulus represents the energy dissipated by
internal structural rearrangements.

strain, particularly for polymers. The proportional limit related to each other through Poisson’s ratio (ν), given
represents the maximum stress at which stress and by the following equation:
strain are proportional, and varies for different materi­
als. Below this limit, the chemical bonds between atoms E = 2G (1 + υ). (3)
in the material stretch when under load but will recover
completely when the load is released. Above this limit, For many common materials, Poisson’s ratio is sim­
the bonds will break and slip past each other, leading to ilar to that of incompressible rubber (v = 0.5); thus,
non-​proportionality. Therefore, the elastic modulus is E is frequently approximated to 3G. Together, the shear
typically measured at low strain values (0.2%)1,11. and elastic moduli represent the properties of materials
Young’s modulus is one of the most common meas­ under two different types of load, which might suffice
ures of intrinsic material stiffness. Because it is inde­ to give a general idea of the rigidity of various biological
pendent of structure, Young’s modulus is widely used to tissues. However, different loads can be simultaneously
characterize the stiffness of both manufactured materials applied to the same tissue: for example, cartilaginous
and tissues (Fig. 1a,b). However, in addition to uniaxial tissues are routinely subjected to both compressive and
stress, biological tissues might also be subjected to defor­ shear deformations12. Furthermore, such complex loads
mations resulting from shear forces (Fig. 1c). The shear can cause materials to deform in multiple ways, such as
modulus (G) is calculated similarly to Young’s modulus torsion and bending (Fig. 1d). In the case of bending, one
in that stress (force per unit area) is divided by strain. side of the sample is subjected to compressive stress and
However, whereas for Young’s modulus stress and the opposite side is subjected to tensile stress. The min­
strain are both normal to the cross-​sectional area, for eralized constituent of bone, for example, as for other
the shear modulus they are parallel and associated with brittle materials, is highly resistant to compressive stress
an angular change. In isotropic materials, E and G are but not so much to tensile forces; consequently, bending

Nature Reviews | Materials


Reviews

can result in fracture. As such, it is important to consider referred to as ‘viscoelastic’. This type of material can be
different types of stress even in the same tissue. The most fully characterized only by time-​dependent tests, which
relevant types of stress in tissue engineering are those apply strain and measure the required force as a function
that are similar to physiological loads. of time (stress relaxation) or apply stress and measure
Furthermore, it is important to outline that some of strain as induced changes in shape (creep) as a func­
these mechanical relationships do not apply directly to tion of time, or through dynamic mechanical analysis13.
most biological tissues. Equation 3, for example, applies In dynamic mechanical analysis, loads are applied and
only to isotropic materials but is frequently erroneously released cyclically, often at differing frequencies or tem­
used for anisotropic materials — those with mechanical peratures, which facilitates measurement of the visco­
characteristics that differ according to the direction of elastic response of a material during faster deformations
measurement. For example, biological tissues have very than those derived by creep and stress relaxation tests.
complex, anisotropic structures in which many types The viscoelastic response of a material is used to
of matter are present and unevenly distributed; they derive the dynamic or complex modulus, which is usu­
differ greatly from the typically isotropic, continuum ally represented by storage and loss moduli. For uniaxial
structures used in manufacturing (in which few types forces, the storage modulus (E′) represents the elastic,
of matter are present and are continuously distributed in instantaneous and reversible response of the material:
the containing space). As such, even though it is useful deformation or stretching of chemical bonds while under
to discuss the relationships that exist between mechan­ load stores energy that is released by unloading. The loss
ical concepts at this simple level, these will, in practice, modulus (E″) represents the viscous time-​dependent
be affected by length scale, anisotropy, spatial varia­ response of the material and is related to irreversible
tions, non-​linear behaviour and other characteristics of rearrangements and remodelling of their internal struc­
biological tissues, discussed further in this Review. ture, such as the slippage of polymer chains past each
Viscoelasticity is another important point to discuss, other. Similarly, for deformations resulting from shear
as the natural response of all solid materials to stress is not forces, the shear storage modulus (G′) and the shear loss
purely elastic but also has a viscous component, which is modulus (G″)14 are frequently evaluated by rheology and
also the case with living tissues. When elastic and viscous oscillatory experiments (Table 1). As biological tissues
components are both prominent in defining the mechan­ generally have viscoelastic responses, these tests are
ical behaviour of a material, the material is generally extremely relevant in the biomechanical field.

Table 1 | Commonly used techniques for the mechanical characterization of living tissues
technique Concept Modulus sample Dimension astM standard refs
test methodsa
Tensile deformation Classic stress–strain analysis. Uniaxial stress is E (elastic) Mostly ex Macroscale D412, D638, D882, 220

applied to stretch the material and a relationship vivo tissue D1623, D1708,
is established with the resulting strain D3039, D3039M
Compressive Classic stress–strain analysis. Uniaxial stress E (elastic) Ex vivo tissue Macroscale D695, D1621 221

deformation is applied to compress the material and a


relationship is established with the resulting
strain; the compressor is as the same size as
or larger than the sample
Dynamic Cycles of tensile and compressive deformation E′, E″ Ex vivo tissue Macroscale D5024, D5026 222

mechanical analysis (viscoelastic)


Shear rheometry Application of small-​amplitude oscillatory shear G′, G″ (shear, Ex vivo tissue Macroscale D5279 223,224

stress and quantification of the resulting strain viscoelastic)


Pipette or Establishment of the relationship between the E (elastic) Ex vivo tissue Macroscale NA 225,226

micropipette pressure of aspiration and the aspirated volume or microscale


aspiration of the sample
Indentation Indentation of the tissue with a probe of defined E (elastic) Mostly ex Nanoscale to E2546-15 105,227

geometry and calculation of the relationship vivo tissue macroscale


between indentation depth and probe load
(the probe must be smaller than the sample)
Atomic force Atomic-​level indentation (nanoindentation) or E Dry or wet Microscale, NA 5,141,228

microscopy shear rheology (atomic force microscopy-​based (indentation), ex vivo tissue nanoscale
rheology) G′, G″ (shear)
Magnetic resonance Magnetic resonance visualization of tissue G′, G″ (shear, In vivo tissue Macroscale NA 229,230

elastography deformation resulting from the introduction viscoelastic) (millimetre)


of shear waves into the tissue derived from
external vibrations
Ultrasonic shear Ultrasonic pulses produce shear waves through E (elastic) In vivo tissue Macroscale NA 231

wave elastography the tissue; the velocity of these waves is measured (millimetre)
and used to derive the tissue’s Young’s modulus
NA, not available. aAdditional test methods are suggested by ASTM F2150-13 where applicable; sample preparation and conditioning guidelines are provided by
ASTM F1634 and ASTM F163.

www.nature.com/natrevmats
Reviews

Timescale Approaches such as computational homogenization are


Some materials exhibit different responses when based, in brief, on defining a representative volume ele­
deformed at different speeds. For example, the toy ment of smaller (usually microscale) dimensions than
Silly Putty is made of a silicone blend that exhibits a the complete structure, which can then be used to model
strong elastic response at high rates of deformation and its governing macroscale behaviour24,25. Within the con­
a strong viscous or liquid-​like response at low rates of text of tissue mechanical properties, such an approach
deformation15. Mechanical tests need to take into con­ involves first defining the microscale components that
sideration the potential for time dependency. A fre­ govern the mechanical properties of the tissue, such as the
quency sweep, for example, is commonly performed in elastin and collagen fibres of the arterial wall. Second,
dynamic tests to evaluate the response of the sample to the properties of these components can be assessed by a
different speeds of deformation. During these tests, the technique that probes their mechanics at the same scale,
storage modulus typically increases with rising defor­ such as atomic force microscopy (AFM) indentation.
mation frequency; that is, the elastic response of these Finally, the characteristics of the small-​scale elements
materials increases with the speed of deformation. Thus, can be used to mathematically model their macroscale
deformation speed can play a crucial part in defining the mechanics, which in this example are those of the whole
response of a material under load16,17. Particularly slow arterial wall26. Similarly, complex scaffolds used in tissue
speeds of deformation can approximate a static system. engineering can also be designed by modelling the indi­
Some mechanical tests can be performed under either vidual components of their architecture, which are then
such quasi-​static conditions or dynamic conditions18. combined to find best-​fit shapes that closely approach
Three important consequences can be derived from the mechanical properties of biological tissues27.
these considerations: first, that under quasi-​static con­ In the next few sections we critically explore the
ditions, the time-​dependent behaviour (viscoelasticity) properties of different tissues in terms of their static and
of materials is simplified because the behaviour of the dynamic mechanics, beginning with the microscale
material corresponds to its ‘associated’ elasticity19,20; and nanoscale domains of biological structures and
second, that quasi-​static testing cannot provide infor­ continuing in size order up to the macroscale level.
mation about the viscous and/or elastic responses that
occur under dynamic loads — namely, those resembling Tissue mechanics at increasing scales
physiological deformations; and third, that substantial To understand the mechanical complexity of an organ
differences can be expected to occur at distinct load or a complete functional tissue, one must first consider
timescales, and as such, direct comparisons between its smallest constituents. A biological tissue is funda­
scales are inadvisable (Box 1). mentally composed of ECM and cells, which coexist in
a delicate synergy: the ECM is deposited by cells, which
Spatial scale strongly interact with it. Beyond the fact that ECM struc­
Differences between the macroscopic properties and tures themselves have inherent mechanical properties,
bulk properties of a material and the effects of surface they are also involved in cellular signalling (derived
microscale or nanoscale features should also be con­ from both biochemical cues and mechanotransduction
sidered. Surfaces can behave very differently to bulk events), which has consequences for the mechanical
material: according to the Nobel laureate Wolfgang properties of cells28,29. ECM mechanics and cell mechan­
Pauli, ‘God made the bulk; surfaces were invented by ics are intimately intertwined and together contribute to
the devil’21. Atoms on the surface of a solid are in an overall tissue behaviour. As such, one must not only con­
anisotropic environment because they can interact with sider the individual properties of ECM and cells but also
other atoms on the surface, in the bulk below, and in understand the development of functional biological
liquid or gas phases. The macroscopic definition of tissues. Multicellular structures require analyses of the
Young’s modulus as an intrinsic material property that locations where cell–cell connections are established and
is independent of size sometimes fails at nanoscale or how these cellular structures bind to the ECM they pro­
microscale dimensions, where the moduli of materials duce and thereby create the basal mechanical framework
might not only differ but also can become size depend­ of biological tissues. Here we use a sequential approach,
ent22. As such, assessments of the mechanical properties starting with an analysis of ECM and cells and gradu­
of biological materials should examine the bulk tissue ally integrating their mechanical behaviour, informed by
as well as its constituents at nanoscale or microscale remarkable insights from developmental biology studies.
dimensions because not only are the moduli themselves
likely to be different at different scales but also because ECM components
at these smaller dimensions cells sense the material and The type, concentration and organization of ECM com­
individually respond to its mechanical characteristics23. ponents (which are all fabricated by cells) deeply affect
These intricacies, together with the wide range of avail­ the mechanical properties of both healthy and diseased
able investigative techniques, make the mechanical char­ tissues (discussed further in the section Abnormal tissue
acterization of biological tissues highly complex, and states). The literature on mechanical properties of individ­
suggest that standardization could lead to improvement ual ECM components is scarce, and most reported analy­
(Box 1; Table 1). ses were done on bulk tissue samples or on structures
Considerable research effort has focused on ways (such as hydrogels) engineered from ECM components30.
to homogenize the microscale properties of a material Nonetheless, several well-​known ECM components occur
and thereby to derive those ruling at the macroscale. in many different tissues and are critical for homeostasis,

Nature Reviews | Materials


Reviews

Table 2 | Mechanical moduli of extracellular matrix components stress–strain curve should be mimicked by materials
with similar J-​shaped curves40.
Component technique E refs
Collagen molecule X-​ray diffraction 3–9 GPa 32
Elastin. Elastin, so named for its elasticity-​enhancing
Collagen fibril AFM indentation or 2–7 GPa (tensile deformation), 33 capabilities, is another ECM component that is inti­
tensile deformation 5.0–11.5 GPa (indentation) mately related to collagen in terms of its contribution
Collagen fibre Microbending 100–360 MPa 232 to tissue mechanics. Both collagen and elastin form
fibrous networks that are intimately interwoven but
Fibronectin fibre Microelectromechanical ≤3.5 MPa (strain dependent) 49

system tensile deformation remain structurally independent of each other. Elastin


is frequently present in organs such as skin, lungs and
Fibrin fibre AFM tensile deformation 1–28 MPa 233,234
bladder and has an important functional role in blood
or microbending
vessels, namely arteries, where specific combinations of
Elastin fibre Tensile deformation 1.1–1.2 MPa 235,236
collagen and elastin determine the mechanical proper­
AFM, atomic force microscopy ; E, uniaxial elastic Young’s modulus. ties of the arterial wall and consequently the regulation
of blood pressure and blood flow41. Importantly, elastin
healing and regeneration, namely proteins (such as col­ shows a notably high linear elastic extension. Thus, the
lagen, elastin, fibronectin, laminin, tenascin and vit­ addition of elastin reduces the elastic modulus of a tis­
ronectin) and glycosaminoglycans (GAGs)31. Each of sue, and as such leads to a more compliant (easier to
these components has distinct functions and, naturally, deform) structure. This effect is accounted for by the
different mechanical properties (Table 2). two orders of magnitude difference between the elastic
moduli of collagen and elastin fibres (Table 2).
Collagen. Collagen is the most abundant ECM com­ The main function of hydrated elastin is to elastically
ponent and is consequently the one that has been most stretch and contract, thereby acting as the driving force
thoroughly explored in terms of its mechanics. From of elastic tissue recoil on release of the force that caused
single molecules to fibrils, collagen structures have the deformation42,43. The interplay of elastin and colla­
elastic moduli in the gigapascal range32,33. The final gen leads to J-​shaped stress–strain curves that show a
contribution of these structures to the bulk mechan­ high initial linear elastic extension, owing to the align­
ics of living tissues depends mainly on collagen abun­ ment of elastin’s disordered structure (entropic), with
dance and organization. Collagen exists in 28 currently linear behaviour at increased deformations owing to
known types, which act not only as structural proteins the molecular stretching of collagen fibres (enthalpic).
but also have other functions, such as cellular adhesion, In practical terms, increasing the elastin-​to-collagen
that are similarly important for overall ECM assem­ ratio leads to a decrease in overall tissue stiffness and
bly and its consequent tissue mechanics (reviewed an increase in its deformability and recoil, which can
elsewhere34,35). Collagen structures and specific domains be used in the development of elastic tissue engineering
within the molecule can also be mineralized, leading to scaffolds44. By contrast, a decreased elastin to collagen
the formation of truly stiff ECMs such as those present ratio has the opposite effects, as observed in several
in bone36. arterial conditions45.
Some tissues are soft at low strains owing to a random
collagen fibre orientation. On stretching, the collagen Other ECM molecules. GAGs are also highly represented
molecules align and become an extremely stiff network, across distinct tissues, but their mechanics are poorly
which shifts the Young’s modulus of the tissue towards understood. In part, this lack of investigation could be
that of the collagen fibres themselves37–39. Thus, tensile due to their apparently minor contribution to the tensile-​
deformation leads to a J-​shaped stress–strain curve in like moduli of tissues compared with that of ECM com­
which deformation can initially be obtained by relatively ponents such as collagen46. Nevertheless, GAGs are
low stress, whereas much higher stress is needed to reach fundamental for water retention, shock absorption,
higher deformations. For example, it does not take much lubrication and tissue viscoelasticity, properties that
force to pinch and pull a few millimetres of skin, but make them extremely important for the biomechanics of
further deformation requires a much higher (and more tissues such as cartilage, specifically its bulk compressive
painful) force. Once the force is released, skin returns stiffness47.
to its original shape. Thus, biological tissues have the The mechanical contribution of GAGs in articu­
capacity to limit the deformation resulting from high lar cartilage is highly dynamic. Aggrecan, one of the
levels of strain owing to reorganization of their structural main cartilage GAGs, can have distinct mechanical
fibres (mainly collagen), which gradually increases tis­ contributions depending on the loading conditions. On
sue stiffness and prevents damage from excessive strain. quasi-​static loading, both electrostatic and steric inter­
As such, it is important to characterize the moduli of actions are responsible for the mechanical function of
tissues at both high and low strain levels. An attempt to GAGs in cartilage, whereas under dynamic load (such
engineer a lifelike tissue based on its low-​strain stiffness as during running), solid–fluid interactions (that is,
would probably fail to fulfil its function of preventing between GAGs and the interstitial fluid within cartilage)
high-​strain damage. Conversely, a tissue engineered become the main contributor. These interactions, also
to reflect only the high-​strain modulus would be too known as poroelasticity, lead to cartilage interstitial
stiff to have the compliant mechanics required for limb fluid pressurization and self-​stiffening of cartilage with
movement. Thus, biological tissues with a J-​shaped increasing loading frequencies48.

www.nature.com/natrevmats
Reviews

Fibronectin, another common component of the potential4. In 2007, a pivotal review entitled ‘The cell as
ECM in various tissues, also has an interesting role a material’57 for the first time considered cells as a mate­
in tissue mechanics. Softer than most fibrillar ECM rial in their own right with characteristic mechanical
structures, fibronectin fibres can be stretched by cell-​ properties that could be investigated by distinct tech­
generated tension and are able to extend more than niques, one of the most important of which was AFM
eightfold before 50% of the fibres will experience rup­ indentation. Overall, cells are soft entities; most have
ture49. Furthermore, fibronectin fibre stiffening leads to bulk elastic moduli in the 0.1–10.0-kPa range, which
an increase in E of up to 3.5 MPa (Table 2), a process hardly varies with cell type (Table 3). However, some
accompanied by a force-​dependent exposure of cryp­ variability-generating details require further discussion.
tic binding sites, which leads to changes in cell behav­ Firstly, cells (like tissues) are heterogeneous. Despite
iour49. This evidence indicates a highly dynamic synergy the softness of whole cells, certain cellular constituents,
between cell and fibronectin mechanics, which has also namely cytoskeleton fibres, have elastic moduli in the
been shown to contribute prominently to the mechani­ gigapascal range58. The contribution of these fibres to
cal driving of tissue development50. Collagen, fibronectin the mechanics of cells varies with the type of load applied.
and many other ECM components interact directly with Intermediate filaments, actin filaments and microtubules
cells, which not only synthesize but also remodel and (in order of increasing stiffness) respond differently to
adhere to them. The interaction is mutual: cells both cre­ distinct types of deformation. Microtubules and actin
ate and influence the ECM components that drive tis­ filaments show substantial resistance to compressive
sue mechanics, which in turn regulate the behaviour of forces, whereas intermediate filaments are best able to
cells. Thus, it is easy to understand how small defects in endure tensional loads59. Furthermore, some cytoskel­
ECM molecules might lead to devastating diseases such eton fibres (such as microtubules and actin filaments)
as osteogenesis imperfecta (brittle bone disease)51, junc­ have a polarized molecular structure that enables them
tional52 and dystrophic53 epidermolysis bullosa (severe to act as molecular motors, which leads to a further level
blistering of the skin) and Ehlers–Danlos syndrome of mechanical complexity. Actin filaments, which dictate
collagenopathies (widespread connective tissue fragility, cell shape and consequently direct events ranging from
including life-​threatening vascular instability)54,55. motility to growth and differentiation, are frequently
complexed with myosin to form a contractile struc­
Cellular mechanics ture termed ‘actomyosin’ (also known as stress fibres)60.
The mechanical properties of cells (as well as their In most animal cells, cell mechanics are dictated by a
shape)56 are closely intertwined with their behaviour. mesh of actomyosin known as the cortex, a mechanically
Only a few hours after fertilization, the mechanical rigid and highly plastic structure immediately below the
pro­p erties of the oocyte already dictate its viability cell membrane61. Within this actomyosin mesh, myosin

Table 3 | Mechanical moduli of distinct cell types


Cell type technique Modulus Modulus values (kPa) (condition) refs
Differentiating MSCs Micropipette E 0.89 (osteogenic), 0.22 (adipogenic) 77

aspiration
MSCs AFM indentation E 2.5 (spherical)a, 3.2 (spread)b 99

Osteoblasts AFM indentation E 2.6 (spherical)a, 6.5 (spread)b 99

Pancreatic cells AFM indentation E 0.54 (healthy cell), 0.54 (carcinoma) 5

Cervical epithelial cells AFM indentation E 2.05 (healthy cell), 2.80 (carcinoma) 159

Fibroblasts AFM indentation E 0.5–30.0 (cell body) 70,237

Pleural effusion cells AFM indentation E 0.38 (metastatic cell), 2.53 (healthy cell) 160

Lung cells AFM rheology G′ 0.5 (alveolar cell), 0.7 (bronchial cell) 238

Embryonic stem cells Magnetic twisting G 0.55 (naive), ≤2 (differentiated) 79

cytometry
Adipose stem cells AFM indentation E 0.6–1.6 76

Smooth muscle cells AFM indentation E 5.9–7.7 239

Cardiomyocytes AFM indentation E 41 97

Keratinocytes AFM indentation E 120–340 240

Kidney proximal tubule cells AFM indentation E 0.35 241

Hepatocytes AFM indentation E 0.75-1 242

Thyroid cells AFM indentation E 1.2 (healthy cell), 1.3 (cancer) 72

Eye lens cells AFM indentation E 4.83 (nuclear cells), 0.22 (cortical cells) 243

AFM, atomic force microscopy ; E, uniaxial elastic Young’s modulus; G, shear modulus; G′, shear storage modulus (dynamic);
MSC, mesenchymal stem cell. aSpherical cells have a round morphology owing to a short culture duration. bOn continued culture,
cells lose their initial spherical morphology and exhibit spread over the surface.

Nature Reviews | Materials


Reviews

Box 2 | towards a unified description of cellular mechanics components can facilitate understanding of the mechan­
ical properties of whole cells. It is useful to picture a cell
the past three decades have seen substantial efforts to define the ruling principle as composed of a membrane encapsulating a very com­
behind cellular mechanics and its consequential effects on mechanotransduction plex elastic network comprising the cytoskeleton and
responses and all consequent cellular events (including cell membrane protrusion, organelles within a fluid (cytosol), collectively termed
adhesion, migration and mitosis). the cellular tensegrity model proposes that the
the cytoplasm66. Unlike most engineered gels, the cyto­
cytoskeleton is a tensioned or prestressed network responsible for maintaining cellular
stability, in the same way that tensegrity architecture allows buildings and bridges to plasm is continuously out of equilibrium owing to the
stay unchanged and robust for countless years213. subsequently, other researchers active role of molecular motors such as myosin. These
hypothesized that cells behave as a soft glassy material; that is, having timescale-​free motors expend energy, and the resulting fibre polymeri­
and frequency-​scale-free rheological behaviour that obeys a weak power law214. zation and network contraction allow intricate mechan­
this notion became known as the soft glassy rheology (sGr) model, which was later ical events to occur at the cellular cortex. Accordingly,
complemented by evidence that at extreme timescales or frequencies, cellular the cortex is characterized as an active gel67. Overall, it
behaviour deviated from a single power law and showed timescale dependency215. is important to understand that living cells integrate a
Nevertheless, among its fundamental characteristics, the sGr model implies that cells set of mechanical properties that are among the most
(in common with other soft glasses) are metastable structures that are naturally out of remarkable and complex found in nature, and which can
equilibrium. although this characteristic is in agreement with the recent description
currently be explained only by resorting to several differ­
of cortical cell networks as out-​of-balance, active gels67, some difference exists
between the sGr model and a stable, equilibrated tensegrity model. Both models are ent theories (Box 2). This heterogeneity will undoubtedly
still used to describe the distinct phenomena governing cell function: both models need to be considered when one is trying to assess the
suggest that actomyosin prestress is the link between cytoskeletal stabilization and bulk stiffness of single cells68.
power-​law modulation of cell behaviour216. Nevertheless, a general unification of these Secondly, when addressing the mechanical properties
models is yet to be established. of cells in adhesive scenarios, it is important to under­
stand that, as a consequence of adhesion and spread­
ing, cells might become extremely thin69. As such, AFM
is able to pull on actin fibres, which generates consider­ indentation depths have to be taken into consideration
able tension, or prestress60. This actomyosin-​generated to avoid ‘bottom effects’ (in which measured stiffness
tension in the cell cortex is responsible for cell shape values are too high owing to an influence of the sur­
and integrity in a fashion that parallels architectural face beneath the cells rather than the mechanics of cells
tensegrity structures62. Moreover, this tension propagates themselves)70. Several methods for correcting for bot­
through cell–cell and cell–ECM junctions, and as such tom effects can be applied in multilayer cell mechanics71.
is fundamental to the architecture and function of liv­ This effect might partially explain why some reported
ing tissues63. Actomyosin-​generated tension is typically mechanical properties of cell monolayers grown on tis­
increased in the intracellular stress fibres that act as an sue culture polystyrene might not recapitulate in vivo
essential link between the cytoskeleton and focal adhe­ ones. Simultaneously, the surface to which the cells
sions, which connect cells to the ECM. This role explains adhere also affects cytoskeletal tension and consequent
why regions of the cell near focal adhesions are stiffer bulk mechanics, and these factors should be prefer­
than the surrounding areas60. For all these reasons, when entially analysed on surfaces that are mechanically as
one is assessing cell mechanics by AFM indentation, the native-​like as possible. For example, the elastic modulus
moduli derived will be highly dependent on the exact of thyroid cells is more than twofold higher when they
membrane zone indented64, a clear source of variability. are cultured on tissue culture polystyrene rather than on
Furthermore, it is important to understand that a softer, thyroid-​like surface72. Nevertheless, even when
cytoskeleton fibres are not the only cell component to the moduli mismatch of cells grown on a native-​like
have an effect on cell stiffness or shape. Interactions of surface is not as large as that of cells cultured on tissue
cytoskeletal fibres with the intracellular and extracellular culture polystyrene, bottom effects might still have a
environment are also involved. For example, cellular considerable effect on measurements of relevant moduli,
activity can take advantage of cytoplasmic changes indicating the need for adequate correction73.
and osmotic forces. Early theories hypothesized that Cell spreading and surface stiffness are not the only
local cell membrane protrusions (also termed ‘blebs’) sources of variability. When cells are constrained in a
were initiated by severing of cytoskeletal fibres, which fixed area, the shape they assume can have a greater effect
was thought to locally reduce the elastic modulus on their mechanics than the surface stiffness itself, which
and thereby lead to an increased cell volume owing is especially important to consider when cells are cultured
to osmotic effects. The consequent rise in hydrostatic on surfaces with non-​flat topographies (such as grooves
pressure was suggested to force the membrane into a or micropillars)74,75. This knowledge might also prove
protrusion, which would eventually be stabilized by important in future initiatives to standardize the meas­
actin diffusing into it and then polymerizing65. In 2005, urement of cell mechanics and thereby obtain more uni­
however, a more likely mechanism was suggested: that form, comparable results. The health status of cells can
local actomyosin contraction at the cortex compresses also give rise to large differences in stiffness (discussed
nearby cytoplasm, which increases local hydrostatic in the section entitled ‘Abnormal tissue states’).
pressure and leads to the formation of blebs even when Altogether, the mechanical properties of cells reflect
cell membrane protrusion is opposed by membrane interesting biological events of relevance to tissue engi­
tension and osmotic pressure66. neering. For example, with regard to cell differentia­
Together, these complex phenomena indicate how an tion, adult stem cells from different niches were shown
appreciation of the mechanical heterogeneity of cellular to have distinct elastic moduli depending on their

www.nature.com/natrevmats
Reviews

commitment: osteogenic stem cells were stiffer than of magnitude to reach the 20 GPa of cortical bone91. This
adipogenic stem cells76,77. Similarly, the differentiation of requirement leads to two important principles: first that,
embryonic stem cells is intimately connected to changes fundamentally, tissue development always proceeds in
in stiffness78,79. Although the mechanical properties of the direction of increasing stiffness, and second that soft
cells themselves cannot yet be artificially engineered environments are always present before stiffer ones arise
or directly manipulated, the mechanical properties of — even after a bone fracture, the regeneration of bone
their surroundings and consequent cell mechanics can is preceded by the accumulation of considerably softer
be80,81. Nevertheless, understanding the mechanics of materials: haematoma, granulation tissue and fracture
bulk tissues requires not only an overview of single-​cell callus (in order of increasing stiffness), within which
and ECM structural mechanics but also an understand­ ECM fibres accumulate and eventually mineralize to form
ing of how single-​cell mechanics integrate to generate stiff bone92,93. To understand how early embryonic struc­
the mechanics of multicellular structures, and of how tures stiffen and eventually develop the mechanical char­
single-​cell and multicellular mechanics combine with acteristics of mature tissues, we must further consider the
ECM mechanics to generate overall tissue mechanics. interplay of ECM and cellular mechanics.
The early embryonic ECM acts mostly as a base­
Cell–cell and cell–ECM interactions ment membrane for embryonic epithelia. Nevertheless,
Many cell types are able to self-​assemble into multi­ the mechanical effects of cell–ECM interactions can
cellular monolayers resembling the epithelial sheets first be observed here. Epithelial cells, as part of their
formed during early development. Where cells contact normal activity, secrete and deposit ECM fibrils, which
each other, they form cell–cell adhesions through the form the basement membrane to which the cells adhere,
binding and assembling of specific proteins, such as thereby gradually increasing its stiffness94. The stiffness
cadherins and catenins82. The resulting epithelial sheet of this early extracellular structure then begins to over­
still behaves actively in mechanical terms, just as sin­ take that of the attached cells, increasing from 20 kPa
gle cells do, being able to sustain extreme strain under at early embryonic stages up to 800 kPa at later times.
constant tension83. As the number of constituent cells Additionally, the ECM fibrils have a preferential orien­
increases, the active tension of the resulting monolayer tation that leads to heterogeneous directional stiffness94,
also increases in a linear fashion84. In this scenario, which represents the earliest appearance of ECM-​related
where the predominant interactions are between cells tissue anisotropy. In the embryonic stages of nematode
rather than with the minimal ECM, stresses can propa­ development, the apical ECM was shown to be essen­
gate over long ranges85. Monolayers of up to thousands tial for cell anchoring and transmission of actin-​derived
of cells were shown to undergo stress relaxation very stresses through the formation of an actin–ECM com­
similarly to single cells, in that the cell–cell junctions posite with distinct mechanical properties95. It is this
serve as stable connections enabling these monolayers very early ECM configuration that causes the shift in
to behave in mechanical terms as single cells ruled by mechanics from single cell-​like to tissue-​like, although
actomyosin dynamics86. Briefly, the application of suf­ the embryonic tissue still goes through considerable
ficient stress leads to the rearrangement of intracellular further dynamic changes related to cell migration and
actomyosin, which in turn promotes extension of the morphogenesis96.
monolayer, thereby dissipating stress and resulting in As the development of these embryonic structures
a new mechanical steady state86. However, once a stiff continues, so does the deposition and accumulation of
ECM is present, stress no longer propagates across the ECM, thereby generating ever more complex structures.
monolayer in the same way85. Eventually, this process leads to the formation of stiffer
As cells begin to aggregate and organize into multi­ tissues. The mechanics of some of these stiffer tissues
layered 3D structures, such as organs and tissues, the are mostly dependent on the ECM, whereas the stiffness
number of cell–cell interactions increases substantially87. of softer, highly cellular tissues is closer to that of their
Research on the mechanics underlying early develop­ constituent cells. Cardiomyocytes have elastic moduli
ment has shown that the viscoelastic (stress relaxation) close to those reported for cardiac tissue97 because the
behaviour of cells still occurs in embryonic tissue and mechanics of muscle tissue are principally dependent
that the overall stiffness of this tissue is also close to that on those of its contractile, actomyosin-​rich cells98. By
of single cells (below 1 kPa)88. Nevertheless, the elas­ contrast, both mesenchymal stem cells (also known as
tic modulus of embryonic tissues shows considerable bone marrow stromal cells) and the osteoblasts they
variation and is also responsible for driving morpho­ differentiate into during osteogenesis are quite soft in
genesis88. At this stage, tissue responses are guided not comparison with bone77,99 because bone mechanics are
only by quick-​dissipating cellular stresses but also by governed mostly by the abundant, highly mineralized
supracellular, persistent ones89. ECM. Naturally, in highly cellular tissues at late stages
The initial phases of tissue formation occur in ex­ of development, the heterogeneity and mechanics of the
tremely dynamic, active and soft environments, where constituent cell types play an important part in the final
three essential entities govern mechanical outcomes: volu­ stiffness of the tissue100. Behind this heterogeneity are
metric growth (driven mainly by cellular proliferation), differences in cell types and their respective structures,
active forces and the material properties of local tissues90. spanning from the cell membrane101 to the previously
The precursors of cartilage, bone and other stiff and soft discussed intricacies of intracellular architecture.
tissues are all present within this soft structure, some of Benchmarking of tissue engineering approaches
which must progressively stiffen by up to seven orders against knowledge obtained from developmental biology

Nature Reviews | Materials


Reviews

has already proved important in the recapitulation of compiled information (Table 4; Supplementary Table 1),
adequate stimuli prompting the in vitro generation some important points are worth discussing further.
of desired cell phenotypes102. Even though much remains
to be learned within this field, understanding of how the Dimension scale and direction
mechanics of different biological components are inte­ Differences between tissue mechanics measured at
grated into a tissue is already possible on a basic level, nanoscale, microscale or macroscale dimensions are
and is expected to facilitate similar analyses of fully extremely relevant to tissue engineering, and there­
developed biological structures. fore dimension should always be considered (Table 4;
Supplementary Table 1). Such differences of scale are
The stiffness range of living tissues mostly a consequence of the heterogeneity of biological
Not surprisingly, a system as complex as a human organ­ tissues, as other authors have already reported105. In gen­
ism comprises tissues that span a remarkable spectrum eral terms, if a tissue is deformed in a tensile manner or
of stiffness; elastic moduli range from the 11 Pa of intes­ the whole tissue is compressed, the behaviour of the bulk
tinal mucus103 to the 20 GPa of cortical bone91. Between material contributes to the overall measured mechani­
these extremes, almost all orders of magnitude are cal moduli. By contrast, if a tissue is indented, only the
represented by a distinct tissue (Fig. 2). region where the indentation occurs (and eventually its
Neural tissues are among the softest of the human nearby surroundings) will contribute to the measured
body104, which should be expected considering their ana­ moduli. The smaller the indenting tool (for example,
tomical protection and how easily they can be damaged. microscale versus nanoscale), the less likely it is that the
These are followed by most abdominal organs (such as measured moduli will be representative of the whole
the pancreas, spleen and liver) and muscles, and finally tissue105. For example, the response to compression of a
by supportive structures such as cartilage, tendons, liga­ whole bone sample integrates both trabecular and cor­
ments and eventually bone. From the analysis of the tical bone mechanics, whereas microindentation of the
same sample is likely to reflect deformation of only one
of these two types of bone106 and nanoindentation will
E Stiff target only a single cell (such as an osteoblast) or ECM
100
component (such as a collagen fibre)107.
10 Similarly, microindentation studies of spinal cord
1 GPa
reported values in the 0.5–1.5-kPa range108, whereas
tensile deformation studies reported values three orders
100 of magnitude higher (1.23 MPa)109. However, micro­
10 indentation involves not only a difference of scale
when compared with tensile deformation but also a
1 MPa
difference in the direction of the applied force; micro­
100 indentation is usually transverse to the tissue, whereas
tensile deformation is mostly longitudinal. Other cor­
10
roborative evidence is provided in reports showing that
1 kPa ­dermis macroindentation results in elastic moduli of
100
~35 kPa (ref.110), whereas tensile deformation results in
elastic moduli in the 50–150-kPa range111. Even when, as
10 in this case, analyses were conducted at the same scale
1 Pa (which might be expected to lead to a fairly small dis­
crepancy), substantial differences can result from altered
Br s
n
Th Fat
nc s
Ki as
ey
ns

la L is
id er

Sp ng
Th een
M oid
e
Bl kin
Co der
ea

Ne ut
po art ve
st ge
ga e
Te ent

Ti Bo n
ta ne
m
Lu l
u

Pa mu

ge

cl

Li ren
ai

o
Ir
m iv
re

iu
uc

Le

re C r
dn

rn

ly ila

nd
us
S
yr

m
ad
l

directionality. These factors explain why certain tech­


y

n
M

niques are not appropriate for direct comparison. Even


ry
ac

if the same type of modulus is derived, the dimension at


ltu
8%

cu

which it is measured as well as its direction must always


ue
ss

be considered.
Ti

Fig. 2 | the stiffness of living tissues spans a full pascal-​to-gigapascal range. The The mechanics of engineered tissues at all dimen­
elastic moduli (E) of different tissues as described in the literature are reported on sions are fundamental for their function: the bulk
the left (logarithmic) scale. Tissues are organized by increasing crescent moduli. mechanics are important to ensure in vivo stability,
Central nervous system tissues, as well as most abdominal organs and skin, have whereas microscale and/or nanoscale properties are
moduli on the submegapascal level and as such are generally characterized as essential for regulating cellular behaviour112.
soft tissues (in biological terms). Cartilage, ligament, tendon and bone are the
stiffest tissues of the human body. For comparison purposes, the moduli of some Tissue anisotropy or composite material?
common tissue engineering materials are included in the graph: 8% acrylamide Composite materials are made of distinct components,
gel217, tissue culture polystyrene218 and titanium (used for dental and bone implants)219.
each with different properties, that when combined
Tissue culture polystyrene, the standard cell-​culture substrate, is clearly stiffer than
almost all biological tissues, and this substrate stiffness is an important source of error form a material with properties that are different from
in mechanical analyses derived from in vitro studies. The reported moduli are derived those of the individual constituents. These materials are
from studies across different animal models and types of deformation, although we used in high-​end applications owing to their optimal
have prioritized studies using human tissues and physiological-​like deformations performance and as such, natural evolution might have
(mostly at macroscale dimensions). followed a similar path113. Many biological tissues easily

www.nature.com/natrevmats
Reviews

fit within the above definition of composite materials114. materials consisting of fibres with specific unidirectional
Some even seem to be composite materials made of mechanical properties bundled together by a resin,
other composites, tendon being an obvious example115. which results in a highly anisotropic material. Similarly,
A tendon is a hierarchical structure made of vari­ tendons are capable of sustaining a high level of stress
ous types of collagen fibres bundled together in sub­ through tensile deformation of the collagen fibres and
fascicles, which in turn are bundled into fascicles of fascicles. The resulting tissue anisotropy is the reason
increasing hierarchy to form the tendon itself. This why some mechanical relationships (such as equation 3)
organization resembles that of synthetic composite do not hold true and should not be used in the analysis
of biological tissue mechanics, and also explains the dif­
ferences in elastic moduli derived by mechanical tests
Table 4 | Mechanical moduli of human tissues
using different directions of deformation.
tissue Dimension Modulus Modulus value refs Parallels could also be drawn between laminated
(condition) composite materials and organs such as the skin, which
Nervous system is composed of three layers of distinct but interacting tis­
Brain Macroscale G 1–3 kPa 169,244,245 sue (epidermis, dermis and hypodermis). The skin has
bulk elastic moduli of up to several hundred kilopas­
Spinal cord Macroscale E 1.23 MPa (tensile) 109
cals116, whereas the fibre-​reinforced dermis117 has elastic
Nerve Macroscale E 5 MPa 246
moduli of around 35–150 kPa and the hypodermis has
Connective tissue elastic moduli of around 2 kPa, similar to the values for
Bone Nanoscale E 1.28–1.97 GPa 127 adipose tissue110,111,118,119. The lens of the eye comprises a
stiff capsule with elastic moduli of 2–3 MPa and a very
Bone Macroscale E 10.4–20.7 GPa 91
soft core (0.8–11.8 kPa)120–123. In the lung, the pleurae
Cartilage Macroscale G 5.7–6.2 MPa 247
are also considerably stiffer than the parenchyma124
Adipose tissue Macroscale E 1.6–5.5 kPa 130,138,248 (Supplementary Table 1). In such tissues, the mechan­
Ligament Macroscale E 25–93 MPa 249–251 ics of individual components might be more relevant
for their engineering than the bulk tissue properties.
Muscle tissue
Finally, some organs also present variable mechanics
Cardiac muscle Macroscale E 8 kPa 252
along their length; for example, the elastic moduli of gut
Cardiac muscle Macroscale G 5–50 kPa 170,253 tissue range from 789 kPa at the ileum to 323 kPa at the
Skeletal muscle Macroscale E, G 5–170 kPa 254–257 descending colon125.
Endothelial or epithelial tissues Tissue condition
Skin Macroscale E 60–850 kPa (35 kPa 110,116,119,258
Tissue condition can also have important mechan­
dermis, 2 kPa hypodermis) ical consequences. Bone mechanical properties can
Skin Nanoscale E 4.5 MPa (epidermis), 140
differ depending on the collection site, the processing
0.1 MPa (dermis) method and storage126 as well as host-​related factors
Lung Macroscale G 0.84–1.50 kPa 259,260 such as age and health status127. Several reports have also
Lung Nanoscale E 1.96 kPa 261 considered the mechanical properties of decellularized
tissues, which not only lack the cellular component but
Other organs
are frequently stiffer than native tissue as a result of the
Kidney Macroscale G 4 kPa (cortex), 5 kPa 262,263
treatments needed to remove cells. This increased tis­
(medulla), 8 kPa (sinus) sue stiffness has been reported for decellularized lung128,
Spleen Macroscale G 15–20 kPa 143–146
pancreas129 and fat130. Furthermore, even with cellular
Liver Macroscale E 4.0–6.5 kPa 264–266 tissues, increased preservation duration can also lead to
stiffening, as shown with liver131.
Liver Macroscale G 2 kPa 267,268

Thymusa Macroscale E 2.11 kPa 130


The applied deformation
Thyroid Macroscale E 9–50 kPa 168,269–271
Native-​like deformations are generally the most use­
Thyroid Macroscale G 1.3–1.9 kPa 272 ful to derive mechanical properties relevant to tissue
engineering. For example, the mechanics of bone and
Pancreas Macroscale E 2.9 kPa 273
cartilage might seem fairly straightforward to analyse.
Pancreas Macroscale G 1.1–2.1 kPa 229,274,275
However, bone can be considered a composite mate­
Bladder Macroscale G 50 kPa (empty) up to 276
rial consisting of mineral, organic and water phases.
100 kPa (full) A compressive test under static load would be useful
Cornea Macroscale E 0.2 MPa 277 to characterize the contribution of the mineral com­
Cornea Nanoscale E 7.5–109.8 kPa b 278 ponent (which has mainly an elastic response) to
bone mechanics, but only dynamic analysis would
Lens Nanoscale E 0.4 MPa (capsule) 279
be able to integrate the viscous contribution of the
Lens Macroscale E 2.3–3.3 MPa (capsule) 121,122
collagen-​rich organic phase, which confers damping
Lens Macroscale G 0.2–10.3 kPa 123 and shock-​absorption properties on bone. Similarly,
E, elastic modulus; G, shear modulus. Decellularized. Distinct layers of the cornea have
a b testing of dry bone will not assess the contribution
different moduli, as detailed in Supplementary Table 1. of hydration132. Articular cartilage, moreover, must

Nature Reviews | Materials


Reviews

Healthy tissue collagen fibrils and cells become more aligned, result­
ing in an extremely stiff structure that can easily reach
Epithelial cells elastic moduli of 100–200 MPa (ref.139). This particular
Proliferating cells mechanical response of tendons is also responsible for
Basement membrane their negative Poisson’s ratio33. Therefore, when one is
Fibroblasts studying tissues with non-​linear stress–strain relation­
ECM ships, emphasis should be given to strain levels, and
low-​strain deformations should not be compared with
Fibrosis Cancer high-​strain ones.

Gaps in the literature


Many examples exist of tissues that have been far less
thoroughly explored with regard to their mechanical
properties. Peripheral nerves, ligaments, gut tissue and
spleen have been analysed in depth at bulk, macro­
scale dimensions but lack further study at nanoscale or
microscale levels. In other tissues, little to no exploration
↑ Tissue stiffness ↑↑ Tissue stiffness of mechanics has been done. For example, the mechan­
• ECM deposition ↑↑ Mass effects (solid stress) ical properties of the thymus have been reported only
• Conversion to • ECM deposition for decellularized tissue130. Other tissues, such as skin,
myofibroblast phenotype • Myofibroblast phenotype
• Neoplastic proliferation require further exploration at smaller scales to clarify
some discrepancies; the mechanical stiffness of epider­
Fig. 3 | tissue mechanics are altered in disease states. Fibrosis and cancer are the two mis is higher than that of dermis in some140 but not all141
main disease states in which the mechanical properties of tissues are prominently affected. nanoindentation studies. Further analysis is necessary
In both conditions, the desmoplastic deposition and crosslinking of extracellular matrix to fully understand and to engineer the multilayered
(ECM) fibres, together with fibroblast proliferation and differentiation into myofibroblasts, mechanics of skin and skin-​like constructs.
lead to increased tissue stiffness. These changes are a response to prolonged healing,
tissue damage (fibrosis) or neoplastic lesions (cancer). Moreover, in tumours, aberrant cell Abnormal tissue states
proliferation compresses the surrounding healthy tissue compartments, leading to mass Scientists have identified several different causes of
effects (such as nerve and blood vessel compression) and an increase in solid stress, which altered tissue stiffness, including cancer, advanced age,
further stiffens the local tissue. Additionally , in cancer, malignant cells can further subvert
diabetes mellitus, cardiovascular disease and fibrosis142.
stromal entities in the tumour microenvironment towards a profibrotic phenotype166,
Nevertheless, some changes in bulk mechanics are not
resulting in increased local fibrosis.
truly caused by changes in the tissues themselves but
rather are attributable to changes in blood pressure and
support the interfaces between bones at the joints and their circulatory consequences. For example, spleen
allow the relative movement of limbs. Therefore, car­ stiffness can increase from 15–20 kPa in healthy indi­
tilage is natively subjected to repetitive, cyclic stress, viduals143–146 to 50 kPa in patients with liver fibrosis147
involving compressive as well as shear forces, which — an abnormal spleen state that is not a consequence of
is why engineering of this tissue presents such a ther­ having abnormal spleen tissue. In the next sections, we
apeutic challenge 133. As such, emphasis should be discuss the main causes of abnormal tissue mechanics:
given to dynamic mechanical analysis and tests such fibrosis and cancer (Fig. 3). Of note, most ECM changes
as shear rheology. Cartilage can have higher moduli in individuals with cancer are fibrotic responses to the
under dynamic testing than under static deforma­ neoplastic event148.
tion134–136 as a consequence of the previously discussed
electrostatic and solid–fluid interactions. Furthermore, Fibrosis
the long-​term responses of complete tissues to applied Fibrosis can be briefly described as the aberrant deposi­
stress or deformation (such as creep and strain relax­ tion and accumulation of connective-​tissue-like ECM in
ation) are thought to be crucial for their physiologi­ a tissue or organ resulting from an imbalance between
cal behaviour. As an example, blood vessels dilate its production and its degradation. This imbalance
in response to increased blood flow137. Nevertheless, leads to the formation of a disorganized and excessively
these responses are frequently not considered in studies crosslinked fibrous network, along with the release of
of tissue mechanics. inflammatory signals, myofibroblast differentiation and
other cellular and/or molecular events149,150, resulting in
Non-​linear behaviour a tissue with increased overall stiffness. Although these
Several biological tissues do not have a linear stress– adverse effects of fibrosis and/or scarring can be disre­
strain relationship, and this non-​linearity can hinder garded in small wounds, large cutaneous scars and wide­
comparisons of different studies on the mechanics of spread fibrosis in tissues or organs such as the lung151,
living tissues. Examples of non-​linearity show up in early kidney152 and liver153 are serious life-​threatening condi­
epithelial monolayers85,86 and can also be seen in adipose tions. Increased liver stiffness is itself a major diagnostic
tissue, in skin and even in tendons, which all deform criterion for liver disease154: a fibrotic liver is stiffer than
easily at low levels of strain but become much stiffer as a healthy one at the both macroscopic level155,156 and the
strain increases37,138. In tendons, as the tissue stretches, microscopic level157.

www.nature.com/natrevmats
Reviews

Cancer non-​healing, fibrotic wound are shared by neoplastic


Unhealthy cells, including cancer cells, frequently have lesions. In both injury and neoplasia, abnormal cell
altered mechanical properties, although the direction proliferation leads to a desmoplastic response (that is,
of these alterations shows no consistent trend. Some dense overgrowth of basement-​membrane-like con­
authors report that cancerous cells are stiffer than their nective tissue with low cellularity and either a translu­
healthy counterparts158,159, whereas others report the cent, cartilage-​like appearance or a hardened texture
opposite5,160. Clearly, each type of tissue and disease can with disorganized blood vessel infiltration). ECM
lead to distinct cell mechanical outcomes161,162. Many of deposition and tissue stiffening similarly occur in the
the signalling pathways that are frequently altered in tumour microenvironment148, where these processes
cancer cells converge on changes in the expression of are further amplified by the ability of cancer cells to
cytoskeleton-​associated proteins, namely those respon­ drive stromal cells towards a profibrotic phenotype166.
sible for actin fibre polymerization and myosin con­ In cancer, moreover, the abnormal cell proliferation
traction163,164. These processes are fundamental for the also generates increased solid stress, caused not by a
mechanical properties of cells and can even have a direct change in the stiffness of the tissue itself but instead by
effect on the ECM mechanics of tumours. Thus, since the increased pressure resulting from tumour growth
cancer cell signalling is highly heterogeneous and varies in a constrained physical volume (mass effects), which
not only with the type of cancer but also within the same compresses nearby healthy tissue148. Together with the
disease type, it is natural that this heterogeneity might increased solid stress, the previously discussed increase
equally transpose to the resulting cancer cell stiffness. in cancer cell tension derived from actin-​related and
Cancer has been characterized as ‘a wound that does myosin-​related signalling might also increase the stiff­
not heal’165. As such, much of the characteristics of a ness of nearby ECM via tensing or realignment of ECM
fibres167. The combination of increased cell stiffness
and solid stress can drastically increase the stiffness of
Static mechanics some malignant lesions: the elastic moduli of healthy
Stiff
thyroid tissue (9.0–11.4 kPa) can increase by a full order
of magnitude to 44–110 kPa in patients with papillary
adenocarcinoma168.

Chronic diseases
Finally, the mechanical properties of tissues might also
be affected in other potentially life-​threatening chronic
diseases unrelated to fibrotic responses or neoplas­
Lifelike changes
In time In space tic alterations. For example, amyloid deposition dis­
eases such as Alzheimer disease169 and cardiac muscle
Intrinsic Extrinsic hypertrophy both increase tissue stiffness170, whereas
Stress stiffening Stiffness gradients osteoporosis reduces bone strength and stiffness171.
Altogether, we believe that pathological changes in
tissue stiffness can always be traced to altered amounts
or functions of its two fundamental constituents: cells
(number and/or phenotype) and ECM (deposition and/or
degradation). Knowledge of abnormal tissue states
and the mechanical changes that occur in disease are
Mechanical deformation important not only for diagnostic purposes but also for
Stress relaxation
under load tissue engineering and recapitulation of such diseases
in in vitro models. These approaches are expected to
generate novel therapeutic targets that allow restoration
of healthy tissue mechanics and mechanotransduction
responses149.

Engineering stiffness into tissues


Fig. 4 | engineering of lifelike tissue mechanics. The ultimate goal of considering Once the mechanics of a complete native tissue are
stiffness in tissue engineering strategies is the recapitulation of qualities of living tissue. understood, the question that remains is how can one
Static mechanical properties are initially considered by trying to mimic the stiffness engineer such mechanics when attempting to construct
and/or softness of the native structure. However, the dynamic nature of biological tissues a lifelike tissue? This question led tissue engineers to
leads to a need to incorporate changes in mechanics (either in space or in time). Time-​ search for novel materials, natural and synthetic, alone
varying mechanics can be obtained by use of materials with stress-​stiffening or stress-​ or blended together and processed through physically,
relaxation responses (whereby the material either stiffens or relaxes in response to a
chemically and biochemically distinct approaches in the
continued duration or increased level of stress, respectively). Alternatively , mechanical
deformation can be applied directly to constructs to mimic these properties when the attempt to recapitulate the properties of living tissues.
material itself lacks such dynamicity. Space-​changing mechanics are found in tissue Several main approaches to engineering the mechan­
interfaces such as those found in regions where cartilage is gradually becoming bone. ical properties of living tissues and ways to trigger
The characteristics of these gradually changing structures can be mimicked by gradients desired mechanotransduction responses by cells can be
of structure, composition and mechanical properties. identified in the literature (Fig. 4).

Nature Reviews | Materials


Reviews

The stiffness of structures (in its simplest form, their for the normal functioning of tissues that are naturally
elastic moduli or shear moduli) is the oldest target within subjected to dynamic loads, including the lung, cartilage
the field. If in the beginning the mechanical properties of and tendons. However, even stiff tissues such as bone
a bulk material were seen only as a requirement for scaf­ exhibit appreciable stress relaxation during active regen­
fold integration and integrity2, emerging discoveries and eration93. Moreover, stress relaxation of engineered tis­
the rise of mechanobiology112 have placed mechanical sues is, like stress stiffening, capable of affecting stem cell
properties at the same level as long-​known biochemical differentiation93. Remarkably, in non-​adhesive environ­
cues. Changing the elastic moduli of the cellular envi­ ments, stress relaxation alone was enough to completely
ronment, either approaching or moving away from those shift the behaviour of chondrocytes from round and
of native tissues, can lead to tremendous differences in static to lifelike cartilage-​forming entities188. This pheno­
cellular responses from simple adhesion172,173 and mor­ type change was due to the ability of the engineered
phology174–176 to the more complex differentiation177–179. material to relax faster than a sample that lacked stress
This variation can be of such a magnitude that high-​ relaxation but had similar initial stiffness, which enabled
throughput technologies are now being applied to derive the cells to become less confined, expand, proliferate and
the best-​fit conditions in terms of surface stiffness for behave as they would in their native environment.
distinct cell types173,180, and approaches to do the same in However, all these studies involved hydrogel struc­
3D environments are both needed and expected. tures with low moduli of a few kilopascals. Despite being
Nevertheless, the scientific community has started far from the stiffness of cartilage or bone, these hydrogel
to shift its attention to more complex stiffness sce­ environments strongly promoted osteogenic and chron­
narios involving dynamic and active tissue mechanics drogenic commitment80,188. As previously mentioned,
(reviewed elsewhere in the epithelial context181). Two the natural evolution of tissues during development
seminal studies brought to light the possibility of achiev­ progresses from soft to stiff environments (for instance,
ing and tuning lifelike, time-​changing forces in 3D bone and cartilage both develop from soft precursor
environ­ments by exploring the properties of biomi­ structures), and so it is only logical that soft environ­
metic material that have profound effects on cellular ments would promote tissue development. An engi­
behav­iour, possibly greater than the influence of static neered structure as soft as adipose tissue, which never
stiffness80,93. Stress stiffening and stress relaxation repre­ matures into a stiffer tissue, can never mechanically
sent the capability of specific materials to either stiffen replace missing mature cartilage or bone. Nonetheless,
or relax in response to changes in stress state or contin­ the evidence obtained in both living tissues and engi­
ued stimulus, respectively. Initially reported on 2D sur­ neered structures clearly indicates that a certain degree
faces182,183, these properties were soon translated to of dynamicity resulting from either stress relaxation or
3D environments and have proved to be very important. stress stiffening (both of which approximate the natu­
ral stimuli that cells receive) is of primal importance for
Stress stiffening recapitulating appropriate cellular behaviour.
Stress stiffening is a natural property observed in many
biological structures, such as gels of the cytoskeletal Mechanical stimulation
­proteins actin and vimentin. Other ECM molecules, Not all materials are capable of dynamic features, and
such as collagen and fibrin, also exhibit substantial specific applications might require the use of static struc­
increases in stiffness beyond a critical stress value184,185. tures. As such, tissue engineers have been attempting to
This behaviour is fundamental for the proper function approximate in vitro the dynamic fluctuations in force
of many biological structures, and its translation to syn­ and shape that happen in vivo by actively stimulating
thetic materials eluded scientists for quite some time. cell–material constructs with controlled mechanical
Once engineering of stress stiffening became poss­ deformations. This approach has been exploited for
ible, controlled changes in the onset of stiffening alone quite some time189 but is increasingly being explored in
(within the same final modulus) were shown to be single- tissue engineering for the maturation of constructs. Most
handedly capable of redirecting the commitment of of the time, the stimulus is an attempt to recapitulate
mesenchymal stem cells from adipogenic to osteogenic stresses that the native tissue endures in vivo, such as
lineages80. A report published in 2019 showed that this compression and shear forces in cartilage differentia­
kind of stress stiffening could be tuned to encompass a tion190 or tensile deformations in skin and hair regen­
50-fold increase, similar to the forces exerted by myosin eration191. In other circumstances, not-​so-native forces
molecular motors: a remarkably lifelike scenario186. can also prove effective, such as stretching in cartilage
formation192 or even nanovibrations of mesenchymal
Stress relaxation stem cells in osteogenesis193.
Stress relaxation represents the capability of a material
to relax over time after the application of strain. This Time-​dependent changes
stress relaxation dissipates energy through a viscoelastic One further option to achieve time-​dependent changes
response as opposed to a purely elastic one (for example is to use cells as the actuators. This quite lively field
slippage versus stretching of chemical bonds in a cova­ of study represents a new paradigm in modern tissue
lently crosslinked hydrogel)182. Viscoelasticity is also engineering. Traditional tissue engineering approaches
present in many biological tissues consisting of a fibrous have aimed to direct cell behaviour by actively stimu­
network (such as collagen) within a non-​f ibrillary lating specific cellular receptors and pathways using
ECM187. Such viscoelastic behaviour is fundamental bioactive molecules, ligands and scaffolds. Although

www.nature.com/natrevmats
Reviews

these methods have shown promising outcomes, we now interface204, while hoping that the interface region is
appreciate that providing cells with a material environ­ stable enough (and large enough) to house interme­
ment that they can remodel, both biochemically and diate cell phenotypes and behaviours. However, inter­
mechanically, might be equally or even more important. face tissues in vivo typically demonstrate a gradient in
Such materials could be defined as biolabile rather than which changes are gradual rather than binary. Similar
bioactive, meaning that they are prone to change and gradients have been produced for osteochondral tissue
remodelling. Similarly, a biolabile environment enables engineering205, an interesting yet challenging approach
cells to bind to, degrade or plastically open up gaps in to the creation of lifelike interfaces. Nevertheless, mate­
the material to move through194, to physically contract rials with 3D gradients in mechanical characteristics
and expand, or mechanically change their shape195, or are still scarcely reported in the literature, and it will be
to remodel the material and deposit their own proteins interesting to see what the future holds with regard to
onto it196. The results of these 2018–2019 studies194–196 this approach to interface tissue fabrication, as scien­
show that tissue engineering might benefit not only from tists begin to develop the toolbox for soft-​to-hard tissue
highly specific cell stimulation cues but also from the bioengineering206.
cells ‘mastering their own fates through the matrix’197,
in vitro. Eventually, both types of stimuli might be used, Future perspectives
as happens in vivo, where local entities remodel and Scientists are beginning to develop and use advanced
regenerate tissues but also respond to systemic cues. biofabrication technologies to mimic the natural proper­
In vivo degradability and remodelling further high­ ties of biological tissues across different scales. Examples
light time-​dependent changes as an important dimen­ can be seen in the fields of additive manufacturing, where
sion of engineered tissue mechanics. The development the precise control and recapitulation of complex biolog­
of materials that degrade in close conjunction with the ical shapes207 is being combined with advanced bioinks
formation of new tissue is the holy grail of tissue engi­ for 3D printing208. Meanwhile, metamaterial research is
neering — the desired outcome being total regeneration also delivering very exotic shapes with complex mechan­
of native tissue and the complete removal of synthetic ical functionality to distinct structures209,210, and it will be
material or the appropriate remodelling of a natural exciting to see how these properties might be combined
scaffold. This degradability implies that the mechanical with biological systems. Simultaneously, a clear trend is
properties of the engineered tissue substitute will vary evident towards the engineering of materials with time-​
over time. Although it is outside the scope of this Review varying mechanics, which have been achieved either
to discuss the complex concept of degradability after through material-​intrinsic stress relaxation and/or stress
implantation, it is important to consider that different stiffening or through external mechanical stimulation.
degradation mechanisms of the tissue substitute could Thus, modern approaches to mechanical tissue engi­
lead to different changes in its mechanical prop­erties neering, which focus on maximizing cell stimulation,
through time. Degradation of polymeric substitutes, for seem to be drifting away from the classic paradigm of
example, might occur on the surface and/or in the bulk, biomechanical tissue replacement.
with or without a change in molecular weight198, which Although these are very interesting approaches, they
raises the prospect of time-​varying changes in their come with one overall caveat: extremely soft environ­
mechanical properties after implantation199 combined ments (up to a few kilopascals) have been shown to
with the previously discussed matrix deposition and promote the best cellular responses even when used
remodelling by cells both from the construct itself for generation of a stiff tissue. How then can these soft
and resident in its vicinity. These changes, even if desir­ engineered structures reach the elastic moduli of their
able and necessary for true construct integration, make living counterparts? Can we integrate time-​dependent
postimplantation outcomes extremely complex. As such, changes in mechanics into stiffer structures and still
advanced translational strategies might benefit from obtain the desired cellular responses? Will cellular
applying predictive methods such as machine learning200 proliferation, ECM deposition and overall tissue mat­
to understand and improve postimplantation outcomes. uration lead to appropriate construct stiffening or can
we push the capacity for stress-​stiffening of currently
Interface tissues available materials to even higher orders of magnitude?
Additionally, the mechanical changes natively occurring The answers to these essential questions will inform
in biological tissues often also have a spatial element, how to efficiently combine the biological advantages
which can be challenging to recapitulate. However, some of soft environments with the mechanical resilience of
tissue structures naturally bridge two distinct environ­ stiff structures. Furthermore, an important need remains
ments with different mechanics (typically one soft mate­ for further manipulating very uniform materials to
rial and one stiff material). Such tissues are often found approach natively non-​uniform structures, such as the
in vivo in locations where transitional tissues are formed, interfacial gradients, anisotropic environments and
such as in cartilage–bone, meniscus–bone, tendon–bone, overall multiscale biological complexity discussed in this
ligament–bone and muscle–tendon interfaces201. Review. In this regard, interesting advances have been
Attempts to engineer similar structures have focused reported in nanocomposite hydrogels211 and combina­
mainly on combining two mechanically distinct com­ tions of hydrogel printing with electrospinning212 that
partments (which might also have other distinct char­ enable the introduction of shape and directionality into
acteristics) into a single structure mimicking either a engineered constructs. Finally, the effects of postimplan­
bone and tendon interface202,203 or a cartilage and bone tation events, particularly remodelling and degradability,

Nature Reviews | Materials


Reviews

on construct mechanics represent an additional level of the section entitled ‘Spatial scale’. Additionally, cellular
complexity that is as yet far from being decoded and mechanics reveal themselves to be extremely complex,
properly integrated within tissue engineering strategies. even in the cell types for which a good picture is already
established. At this level, two main challenges seem to
Conclusions remain: first, the unification of tensegrity and soft glassy
We have seen that biological tissues have very distinc­ rheology models (Box 2), both of which might be linked
tive mechanics due to their structural complexity and by the prestress role of the cytoskeleton, and second,
heterogeneity, which only increase with evaluations of translation of the knowledge derived from cells in mono­
the properties of their individual constituents (both layer cultures to those within 3D environments, where
ECM and cells). Regarding bulk tissue mechanics, a further complexity is likely to arise. Closing these and
substantial amount of literature is derived from different other gaps in knowledge will illuminate the interplay of
animal models and distinct techniques. However, high force and shape in biological tissues and will underlie
levels of disparity persist when some values are derived the road map for engineering lifelike structures in health
and wrongful comparisons are made by not considering and disease states.
the dimension and/or type of deformation studied (in It seems clear that there will never be a true standard
terms of its magnitude, directionality and dynamicity). for the mechanical analysis of biological tissue and that
Furthermore, some bulk tissues have been analysed only variability in the results of such analyses will always be
at a single scale (macroscale, microscale or nanoscale). high. Nevertheless, we believe that by properly integrat­
As a consequence, these disparities raise questions that ing the results derived from different analysis methods
demand further research. towards material-​translatable conclusions, the mechan­
At microscale and nanoscale dimensions, the ical intricacies of living tissues can be well understood.
mechanics of several ECM structures remain unex­ Similarly, the tools to recapitulate each individual tis­
plored, despite their well-​known biochemical roles. sue might soon be at our disposal. Understanding how
Future research into these structures might be important to exploit synergies between soft and stiff materials to
to improve our understanding of how they contribute to maximize cell stimulation while also approaching the
bulk tissue mechanics and also for improving homo­ mechanics of the mature biological tissue could be
genization — that is, the prediction of bulk material the final frontier for engineering truly lifelike tissues.
properties informed by knowledge of their microscale
components’ mechanical behaviour, as discussed in Published online xx xx xxxx

1. Callister, W. D. Jr & Rethwisch, D. G. Materials polyester composites. Compos. Sci. Technol. 63, of arterial tissue based on homogenization and
Science and Engineering: An Introduction 8th edn 283–293 (2003). optimization. J. Biomech. 41, 2673–2680 (2008).
(Wiley, 2007). 14. Meyers, M. A. & Chawla, K. K. Mechanical Behavior 27. Hollister, S. J. & Lin, C. Y. Computational design of
2. Langer, R. & Vacanti, J. P. Tissue engineering. Science of Materials 2nd edn (Cambridge Univ. Press, 2009). tissue engineering scaffolds. Comput. Methods Appl.
260, 920–926 (1993). 15. Cross, R. Elastic and viscous properties of Silly Putty. Mech. Eng. 196, 2991–2998 (2007).
3. Discher, D. E., Janmey, P. & Wang, Y. Tissue cells feel Am. J. Phys. 80, 870–875 (2012). 28. Kim, S. H., Turnbull, J. & Guimond, S. Extracellular
and respond to the stiffness of their substrate. Science 16. Omari, E. A., Varghese, T., Kliewer, M. A., Harter, J. & matrix and cell signalling: the dynamic cooperation
310, 1139–1143 (2005). Hartenbach, E. M. Dynamic and quasi-​static mechanical of integrin, proteoglycan and growth factor receptor.
A seminal work reporting for the first time that testing for characterization of the viscoelastic properties J. Endocrinol. 209, 139–151 (2011).
mechanics alone affect the behaviour of cells. of human uterine tissue. J. Biomech. 48, 1730–1736 29. Miller, C. J. & Davidson, L. A. The interplay between
4. Yanez, L. Z., Han, J., Behr, B. B., Pera, R. A. R. & (2015). cell signalling and mechanics in developmental
Camarillo, D. B. Human oocyte developmental 17. Karunaratne, A., Li, S. & Bull, A. M. J. Nano-​scale processes. Nat. Rev. Genet. 14, 733–744 (2013).
potential is predicted by mechanical properties mechanisms explain the stiffening and strengthening 30. Alcaraz, J. et al. Laminin and biomimetic extracellular
within hours after fertilization. Nat. Commun. 7, of ligament tissue with increasing strain rate. Sci. Rep. elasticity enhance functional differentiation in
10809 (2016). 8, 3707 (2018). mammary epithelia. EMBO J. 27, 2829–2838 (2008).
5. Cross, S. E., Jin, Y. S., Rao, J. & Gimzewski, J. K. 18. Wang, L. & Liu, X. Characterization of viscoelastic 31. Keane, T. J., Horejs, C. M. & Stevens, M. M. Scarring vs.
Nanomechanical analysis of cells from cancer materials by quasi-​static and dynamic indentation. functional healing: matrix-​based strategies to regulate
patients. Nat. Nanotechnol. 2, 780–783 (2007). Meas. Sci. Technol. 25, 064017 (2014). tissue repair. Adv. Drug Deliv. Rev. 129, 407–419
6. Baumgart, F. & Cordey, J. Stiffness — an unknown 19. Schapery, R. A. Two simple approximate methods (2018).
world of mechanical science? Injury 32, 14–23 (2001). of Laplace transform inversion for viscoelastic stress 32. Sasaki, N. & Odajima, S. Stress–strain curve and
7. Pang, Z., Deeth, H., Sopade, P., Sharma, R. & Bansal, N. analysis. Calif. Inst. Technol. https://resolver.caltech. Young’s modulus of a collagen molecule as determined
Rheology, texture and microstructure of gelatin gels edu/CaltechAUTHORS:20141114-114344034 by the X-​ray diffraction technique. J. Biomech. 29,
with and without milk proteins. Food Hydrocoll. 35, (1961). 655–658 (1996).
484–493 (2014). 20. Schapery, R. A. Stress analysis of viscoelastic 33. Wenger, M. P. E., Bozec, L., Horton, M. A. &
8. Koga, Y., Koga, T., Kinekawa, Y. & Kitabatake, N. composite materials. J. Compos. Mater. 1, 228–267 Mesquidaz, P. Mechanical properties of collagen
Properties of a thermostable emulsion prepared from (1967). fibrils. Biophys. J. 93, 1255–1263 (2007).
process whey protein and olive oil; use as a cream-​ 21. Yofe, A. D. Physics at surfaces. Contemp. Phys. 29, 34. Bornstein, P. & Sage, H. Structurally distinct collagen
substitute and its practical application to panna-​cotta. 411–414 (1988). types. Annu. Rev. Biochem. 49, 957–1003 (1980).
J. Cook. Sci. Jpn 34, 154–163 (2001). 22. Abazari, A. M., Safavi, S. M., Rezazadeh, G. & 35. Shoulders, M. D. & Raines, R. T. Collagen structure
9. Williams, S. H., Wright, B. W., Truong, V., Villanueva, L. G. Modelling the size effects on the and stability. Annu. Rev. Biochem. 78, 929–958
den, Daubert, C. R. & Vinyard, C. J. Mechanical mechanical properties of micro/nano structures. (2009).
properties of foods used in experimental studies of Sensors 15, 28543–28562 (2015). 36. Zhang, W., Huang, Z. L., Liao, S. S. & Cui, F. Z.
primate masticatory function. Am. J. Primatol. 67, 23. McNamara, L. E. et al. The role of microtopography Nucleation sites of calcium phosphate crystals during
329–346 (2005). in cellular mechanotransduction. Biomaterials 33, collagen mineralization. J. Am. Ceram. Soc. 86,
10. Perry, J. M. G., Bastian, M. L., St Clair, E. & 2835–2847 (2012). 1052–1054 (2003).
Hartstone-Rose, A. Maximum ingested food size in 24. Peric, D. et al. On micro-​to-macro transitions for 37. Herchenhan, A. et al. Tenocyte contraction induces
captive anthropoids. Am. J. Phys. Anthropol. 158, multi-scale analysis of non-​linear heterogeneous crimp formation in tendon-​like tissue. Biomech.
92–104 (2015). materials: unified variational basis and finite element Model. Mechanobiol. 11, 449–459 (2012).
11. Davis, J. R. (ed.) Tensile Testing 2nd edn (ASM implementation. Int. J. Numer. Methods Eng. 87, 38. Hornsby, J. et al. Quantitative multiphoton microscopy
International, 2004). 149–170 (2011). of murine urinary bladder morphology during in situ
12. Wong, B. L., Bae, W. C., Gratz, K. R. & Sah, R. L. Shear 25. Geers, M. G. D., Kouznetsova, V. G. & uniaxial loading. Acta Biomater. 64, 59–66 (2017).
deformation kinematics during cartilage articulation: Brekelmans, W. A. M. Multi-​scale computational 39. Wiesinger, H. P., Rieder, F., Kösters, A., Müller, E. &
effect of lubrication, degeneration, and stress homogenization: trends and challenges. J. Comput. Seynnes, O. R. Are sport-​specific profiles of tendon
relaxation. Mol. Cell. Biomech. 5, 197–206 (2008). Appl. Math. 234, 2175–2182 (2010). stiffness and cross-​sectional area determined by
13. Pothan, L. A., Oommen, Z. & Thomas, S. Dynamic 26. Speirs, D. C. D., de Souza Neto, E. A. & Perić, D. An structural or functional integrity? PLoS One 11,
mechanical analysis of banana fiber reinforced approach to the mechanical constitutive modelling e0158441 (2016).

www.nature.com/natrevmats
Reviews

40. Ma, Y., Feng, X., Rogers, J. A., Huang, Y. & Zhang, Y. 64. Mandriota, N. et al. Cellular nanoscale stiffness 90. Stooke-​Vaughan, G. A. & Campàs, O. Physical control
Design and application of ‘J-​shaped’ stress–strain patterns governed by intracellular forces. Nat. Mater. of tissue morphogenesis across scales. Curr. Opin.
behavior in stretchable electronics: a review. Lab. Chip 18, 1071–1077 (2019). Genet. Dev. 51, 111–119 (2018).
17, 1689–1704 (2017). The data in this article provide evidence of how 91. Rho, J. Y., Ashman, R. B. & Turner, C. H. Young’s
41. Wagenseil, J. E. & Mecham, R. P. Elastin in large much influence intracellular forces and states have modulus of trabecular and cortical bone material:
artery stiffness and hypertension. J. Cardiovasc. on local cellular stiffness. ultrasonic and microtensile measurements.
Transl. Res. 5, 264–273 (2012). 65. Condeelis, J. Life at the leading edge: the formation J. Biomech. 26, 111–119 (1993).
42. Muiznieks, L. D., Weiss, A. S. & Keeley, F. W. Structural of cell protrusions. Annu. Rev. Cell Biol. 9, 411–444 92. McDonald, S. J. et al. Early fracture callus displays
disorder and dynamics of elastin. Biochem. Cell Biol. (1993). smooth muscle-​like viscoelastic properties ex vivo:
88, 239–250 (2010). 66. Charras, G. T., Yarrow, J. C., Horton, M. A., implications for fracture healing. J. Orthop. Res. 27,
43. Muiznieks, L. D. & Keeley, F. W. Molecular assembly Mahadevan, L. & Mitchison, T. J. Non-​equilibration 1508–1513 (2009).
and mechanical properties of the extracellular matrix: of hydrostatic pressure in blebbing cells. Nature 435, 93. Chaudhuri, O. et al. Hydrogels with tunable stress
a fibrous protein perspective. Biochim. Biophys. Acta 365–369 (2005). relaxation regulate stem cell fate and activity.
1832, 866–875 (2013). 67. Prost, J., Jülicher, F. & Joanny, J. F. Active gel physics. Nat. Mater. 15, 326–334 (2015).
44. Ryan, A. J. & O’Brien, F. J. Insoluble elastin reduces Nat. Phys. 11, 111–117 (2015). A seminal work on the effect of controlled 3D
collagen scaffold stiffness, improves viscoelastic An article providing a complete description of the stress relaxation on the behaviour of stem cells.
properties, and induces a contractile phenotype in physics of active gels. 94. Chlasta, J. et al. Variations in basement membrane
smooth muscle cells. Biomaterials 73, 296–307 68. Guo, M. et al. Cell volume change through water mechanics are linked to epithelial morphogenesis.
(2015). efflux impacts cell stiffness and stem cell fate. Development 144, 4350–4362 (2017).
45. Tsamis, A., Krawiec, J. T. & Vorp, D. A. Elastin and Proc. Natl Acad. Sci. USA 114, E8618–E8627 (2017). 95. Vuong-​Brender, T. T. K., Suman, S. K.
collagen fibre microstructure of the human aorta in 69. Lekka, M. & Laidler, P. Applicability of AFM in cancer & Labouesse, M. The apical ECM preserves
ageing and disease: a review. J. R. Soc. Interface 10, detection. Nat. Nanotechnol. 4, 72–72 (2009). embryonic integrity and distributes mechanical
20121004 (2013). 70. Gavara, N. & Chadwick, R. S. Determination of the stress during morphogenesis. Development 144,
46. Ahmadzadeh, H., Connizzo, B. K., Freedman, B. R., elastic moduli of thin samples and adherent cells using 4336–4349 (2017).
Soslowsky, L. J. & Shenoy, V. B. Determining the conical atomic force microscope tips. Nat. Nanotechnol. 96. Nerurkar, N. L., Lee, C. H., Mahadevan, L. & Tabin, C. J.
contribution of glycosaminoglycans to tendon 7, 733–736 (2012). Molecular control of macroscopic forces drives
mechanical properties with a modified shear-​lag 71. Dimitriadis, E. K., Horkay, F., Maresca, J., Kachar, B. & formation of the vertebrate hindgut. Nature 565,
model. J. Biomech. 46, 2497–2503 (2013). Chadwick, R. S. Determination of elastic moduli of thin 480–484 (2019).
47. Quinn, T. M., Dierickx, P. & Grodzinsky, A. J. layers of soft material using the atomic force 97. Benech, J. C. et al. Diabetes increases stiffness of live
Glycosaminoglycan network geometry may contribute microscope. Biophys. J. 82, 2798–2810 (2002). cardiomyocytes measured by atomic force microscopy
to anisotropic hydraulic permeability in cartilage 72. Rianna, C. & Radmacher, M. Comparison of nanoindentation. Am. J. Physiol. Physiol. 307,
under compression. J. Biomech. 34, 1483–1490 viscoelastic properties of cancer and normal thyroid C910–C919 (2014).
(2001). cells on different stiffness substrates. Eur. Biophys. J. 98. Somlyo, A. P. et al. Ultrastructure, function and
48. Tavakoli Nia, H. et al. Aggrecan nanoscale solid–fluid 46, 309–324 (2017). composition of smooth muscle. Ann. Biomed. Eng.
interactions are a primary determinant of cartilage 73. Kaushik, G., Fuhrmann, A., Cammarato, A. & 11, 579–588 (1983).
dynamic mechanical properties. ACS Nano 9, Engler, A. J. In situ mechanical analysis of myofibrillar 99. Darling, E. M., Topel, M., Zauscher, S., Vail, T. P.
2614–2625 (2015). perturbation and aging on soft, bilayered Drosophila & Guilak, F. Viscoelastic properties of human
49. Klotzsch, E. et al. Fibronectin forms the most myocardium. Biophys. J. 101, 2629–2637 (2011). mesenchymally-​derived stem cells and primary
extensible biological fibers displaying switchable 74. Tan, J. L. et al. Cells lying on a bed of microneedles: osteoblasts, chondrocytes, and adipocytes.
force-exposed cryptic binding sites. Proc. Natl Acad. an approach to isolate mechanical force. Proc. Natl J. Biomech. 41, 454–464 (2008).
Sci. USA 106, 18267–18272 (2009). Acad. Sci. USA 100, 1484–1489 (2003). 100. Li, X., Das, A. & Bi, D. Mechanical heterogeneity
50. Dray, N. et al. Cell–fibronectin interactions propel 75. Tee, S. Y., Fu, J., Chen, C. S. & Janmey, P. A. Cell shape in tissues promotes rigidity and controls cellular
vertebrate trunk elongation via tissue mechanics. and substrate rigidity both regulate cell stiffness. invasion. Phys. Rev. Lett. 123, 058101 (2019).
Curr. Biol. 23, 1335–1341 (2013). Biophys. J. 100, L25–L27 (2011). 101. Roduit, C. et al. Elastic membrane heterogeneity
51. Gautieri, A., Uzel, S., Vesentini, S., Redaelli, A. & 76. Gonzalez-​Cruz, R. D., Fonseca, V. C. & Darling, E. M. of living cells revealed by stiff nanoscale membrane
Buehler, M. J. Molecular and mesoscale mechanisms Cellular mechanical properties reflect the domains. Biophys. J. 94, 1521–1532 (2008).
of osteogenesis imperfecta disease in collagen fibrils. differentiation potential of adipose-​derived 102. Marturano, J. E. et al. Embryonically inspired scaffolds
Biophys. J. 97, 857–865 (2009). mesenchymal stem cells. Proc. Natl Acad. Sci. USA regulate tenogenically differentiating cells. J. Biomech.
52. Mavilio, F. et al. Correction of junctional epidermolysis 109, E1523–E1529 (2012). 49, 3281–3288 (2016).
bullosa by transplantation of genetically modified 77. Yu, H. et al. Mechanical behavior of human 103. Sotres, J., Jankovskaja, S., Wannerberger, K. &
epidermal stem cells. Nat. Med. 12, 1397–1402 mesenchymal stem cells during adipogenic and Arnebrant, T. Ex-​vivo force spectroscopy of intestinal
(2006). osteogenic differentiation. Biochem. Biophys. Res. mucosa reveals the mechanical properties of mucus
53. Wagner, J. E. et al. Bone marrow transplantation for Commun. 393, 150–155 (2010). blankets. Sci. Rep. 7, 1–14 (2017).
recessive dystrophic epidermolysis bullosa. N. Engl. 78. Norcross, S., Horsley, V., Mertz, A. F., Rosowski, K. A. 104. Tyler, W. J. The mechanobiology of brain function.
J. Med. 363, 629–639 (2010). & Dufresne, E. R. Edges of human embryonic stem cell Nat. Rev. Neurosci. 13, 867–878 (2012).
54. Germain, D. P. Clinical and genetic features of vascular colonies display distinct mechanical properties and 105. McKee, C. T., Last, J. A., Russell, P. & Murphy, C. J.
Ehlers–Danlos syndrome. Ann. Vasc. Surg. 16, differentiation potential. Sci. Rep. 5, 14218 (2015). Indentation versus tensile measurements of Young’s
391–397 (2002). 79. Poh, Y.-C. et al. Material properties of the cell dictate modulus for soft biological tissues. Tissue Eng.
55. De Paepe, A. & Malfait, F. The Ehlers–Danlos stress-​induced spreading and differentiation in Part B Rev. 17, 155–164 (2011).
syndrome, a disorder with many faces. Clin. Genet. embryonic stem cells. Nat. Mater. 9, 82–88 (2009). 106. Does, M. D. et al. Insights into reference point
82, 1–11 (2012). 80. Das, R. K., Gocheva, V., Hammink, R., Zouani, O. F. indentation involving human cortical bone:
56. Von Erlach, T. C. et al. Cell-​geometry-dependent & Rowan, A. E. Stress-​stiffening-mediated stem-​cell Sensitivity to tissue anisotropy and mechanical
changes in plasma membrane order direct stem commitment switch in soft responsive hydrogels. behavior. J. Mech. Behav. Biomed. Mater. 37,
cell signalling and fate. Nat. Mater. 17, 237–242 Nat. Mater. 15, 318–325 (2015). 174–185 (2014).
(2018). 81. Leong, K. W., Yim, E. K. F., Kulangara, K., Darling, E. M. 107. Haase, K. & Pelling, A. Investigating cell mechanics
57. Kasza, K. E. et al. The cell as a material. Curr. Opin. & Guilak, F. Nanotopography-​induced changes in focal with atomic force microscopy. J. R. Soc. Interface 12,
Cell Biol. 19, 101–107 (2007). adhesions, cytoskeletal organization, and mechanical 20140970 (2015).
This is the first article to consider cells as a properties of human mesenchymal stem cells. 108. Saxena, T., Gilbert, J., Stelzner, D. & Hasenwinkel, J.
material. Biomaterials 31, 1299–1306 (2009). Mechanical characterization of the injured spinal
58. Suresh, S. Biomechanics and biophysics of cancer 82. Conte, V. et al. Control of cell–cell forces and cord after lateral spinal hemisection injury in the rat.
cells. Acta Mater. 55, 3989–4014 (2007). collective cell dynamics by the intercellular adhesome. J. Neurotrauma 29, 1747–1757 (2012).
59. Fletcher, D. A. & Mullins, R. D. Cell mechanics and Nat. Cell Biol. 17, 409–420 (2015). 109. Oakland, R. J., Hall, R. M., Wilcox, R. K. & Barton, D. C.
the cytoskeleton. Nature 463, 485–492 (2010). 83. Latorre, E. et al. Active superelasticity in three-​ The biomechanical response of spinal cord tissue
60. Kumar, S. et al. Viscoelastic retraction of single dimensional epithelia of controlled shape. Nature to uniaxial loading. Proc. Inst. Mech. Eng. Part H 220,
living stress fibers and its impact on cell shape, 563, 203–208 (2018). 489–492 (2006).
cytoskeletal organization, and extracellular 84. Vincent, R. et al. Active tensile modulus of an epithelial 110. Pailler-​Mattei, C., Bec, S. & Zahouani, H. In vivo
matrix mechanics. Biophys. J. 90, 3762–3773 monolayer. Phys. Rev. Lett. 115, 248103 (2015). measurements of the elastic mechanical properties of
(2006). 85. Charras, G. & Yap, A. S. Tensile forces and human skin by indentation tests. Med. Eng. Phys. 30,
61. Salbreux, G., Charras, G. & Paluch, E. Actin cortex mechanotransduction at cell–cell junctions. Curr. Biol. 599–606 (2008).
mechanics and cellular morphogenesis. Trends Cell 28, R445–R457 (2018). 111. Pissarenko, A. et al. Tensile behavior and structural
Biol. 22, 536–545 (2012). 86. Khalilgharibi, N. et al. Stress relaxation in epithelial characterization of pig dermis. Acta Biomater. 86,
This article offers an interesting perspective on monolayers is controlled by the actomyosin cortex. 77–95 (2019).
cell cortex dynamics. Nat. Phys. 15, 839–847 (2019). 112. Iskratsch, T., Wolfenson, H. & Sheetz, M. P.
62. Ingber, D. E. Tensegrity-​based mechanosensing from 87. Gonzalez-​Rodriguez, D., Guevorkian, K., Douezan, S. Appreciating force and shape — the rise of
macro to micro. Prog. Biophys. Mol. Biol. 97, & Brochard-​Wyart, F. Soft matter models of developing mechanotransduction in cell biology. Nat. Rev.
163–179 (2008). tissues and tumors. Science 338, 910–917 (2012). Mol. Cell Biol. 15, 825–833 (2014).
This article provides a complete overview of 88. Serwane, F. et al. In vivo quantification of spatially 113. Guo, K. & Buehler, M. J. Nature’s way: hierarchical
tensegrity in cell mechanics and its parallel with varying mechanical properties in developing tissues. strength weakness. Matter 1, 302–303 (2019).
tensegrity architecture. Nat. Methods 14, 181–186 (2017). 114. Ramakrishna, S., Mayer, J., Wintermantel, E. &
63. Ingber, D. E. From mechanobiology to 89. Mongera, A. et al. A fluid-​to-solid jamming transition Leong, K. W. Biomedical applications of polymer-​
developmentally inspired engineering. Phil. Trans. R. underlies vertebrate body axis elongation. Nature composite materials: a review. Compos. Sci. Technol.
Soc. Lond. B Biol. Sci. 373, 20170323 (2018). 561, 401–405 (2018). 61, 1189–1224 (2001).

Nature Reviews | Materials


Reviews

115. Zhang, G. et al. Development of tendon structure 139. Wood, L. K. & Brooks, S. V. Ten weeks of treadmill 165. Dvorak, H. F. Tumors: wounds that do not heal.
and function: regulation of collagen fibrillogenesis. running decreases stiffness and increases collagen Similarities between tumor stroma generation and
J. Musculoskelet. Neuronal Interact. 5, 5–21 (2005). turnover in tendons of old mice. J. Orthop. Res. 34, wound healing. N. Engl. J. Med. 315, 1650–1659
116. Agache, P. G., Monneur, C., Leveque, J. L. & De Rigal, J. 346–353 (2016). (1986).
Mechanical properties and Young’s modulus of human 140. Peñuela, L. et al. Atomic force microscopy for 166. Dias Carvalho, P. et al. KRAS oncogenic signaling
skin in vivo. Arch. Dermatol. Res. 269, 221–232 biomechanical and structural analysis of human extends beyond cancer cells to orchestrate the
(1980). dermis: a complementary tool for medical diagnosis microenvironment. Cancer Res. 78, 7–14 (2018).
117. Menon, G. K. New insights into skin structure: and therapy monitoring. Exp. Dermatol. 27, 150–155 167. Huang, S. & Ingber, D. E. Cell tension, matrix
scratching the surface. Adv. Drug Deliv. Rev. 54, (2018). mechanics, and cancer development. Cancer Cell 8,
S3–S17 (2002). 141. Crichton, M. L. et al. The viscoelastic, hyperelastic 175–176 (2005).
118. Skulborstad, A. J., Swartz, S. M. & Goulbourne, N. C. and scale dependent behaviour of freshly excised 168. Lyshchik, A. et al. Elastic moduli of thyroid tissues under
Biaxial mechanical characterization of bat wing skin. individual skin layers. Biomaterials 32, 4670–4681 compression. Ultrason. Imaging 110, 101–110 (2005).
Bioinspir. Biomim. 10, 36004 (2015). (2011). 169. Murphy, M. C. et al. Regional brain stiffness changes
119. Hamasaki, T., Yamaguchi, T. & Iwamoto, M. 142. Lampi, M. C. & Reinhart-​King, C. A. Targeting across the Alzheimer’s disease spectrum. Neuroimage
Estimating the influence of age-​related changes extracellular matrix stiffness to attenuate disease: Clin. 10, 283–290 (2016).
in skin stiffness on tactile perception for static from molecular mechanisms to clinical trials. 170. Chaturvedi, R. R. et al. Passive stiffness of myocardium
stimulations. J. Biomech. Sci. Eng. 13, 17–00575 Sci. Transl Med. 10, eaao0475 (2018). from congenital heart disease and implications for
(2018). 143. Stefanescu, H. et al. Spleen stiffness measurement diastole. Circulation 121, 979–988 (2010).
120. Cui, J., Lee, C. H., Delbos, A., McManus, J. J. & using fibroscan for the noninvasive assessment 171. Vardakastani, V. et al. Increased intra-​cortical porosity
Crosby, A. J. Cavitation rheology of the eye lens. of esophageal varices in liver cirrhosis patients. reduces bone stiffness and strength in pediatric
Soft Matter 7, 7827–7831 (2011). J. Gastroenterol. Hepatol. 26, 164–170 (2011). patients with osteogenesis imperfecta. Bone 69,
121. Krag, S. & Andreassen, T. T. Mechanical properties of 144. Hu, X. et al. Indirect prediction of liver fibrosis by 61–67 (2014).
the human posterior lens capsule. Invest. Opthalmol. quantitative measurement of spleen stiffness using 172. Ye, K. et al. Matrix stiffness and nanoscale spatial
Vis. Sci. 44, 691 (2003). the fibroscan system. J. Ultrasound Med. 33, 73–81 organization of cell-​adhesive ligands direct stem cell
122. Danielsen, C. C. Tensile mechanical and creep (2014). fate. Nano Lett. 15, 4720–4729 (2015).
properties of Descemet’s membrane and lens capsule. 145. Veiga, Z. S. T. et al. Transient elastography evaluation 173. Zhou, Q. et al. Development of a novel orthogonal
Exp. Eye Res. 79, 343–350 (2004). of hepatic and spleen stiffness in patients with double gradient for high-​throughput screening of
123. Besner, S., Scarcelli, G., Pineda, R. & Yun, S. H. hepatosplenic schistosomiasis. Eur. J. Gastroenterol. mesenchymal stem cells–materials interaction.
In vivo Brillouin analysis of the aging crystalline lens. Hepatol. 29, 730–735 (2017). Adv. Mater. Interfaces 5, 4–11 (2018).
Invest. Ophthalmol. Vis. Sci. 57, 5093–5100 (2016). 146. Pawluś, A. et al. Shear wave elastography of the 174. Garreta, E. et al. Fine tuning the extracellular
124. Tenorio, L. E. M., Devine, K. J., Lee, J., Kowalewski, T. M. spleen: evaluation of spleen stiffness in healthy environment accelerates the derivation of kidney
& Barocas, V. H. Biomechanics of human parietal volunteers. Abdom. Radiol. 41, 2169–2174 (2016). organoids from human pluripotent stem cells.
pleura in uniaxial extension. J. Mech. Behav. 147. Chien, C. H. et al. Transient elastography for spleen Nat. Mater. 18, 397–405 (2019).
Biomed. Mater. 75, 330–335 (2017). stiffness measurement in patients with cirrhosis role in 175. Uynuk-​Ool, T. et al. The geometrical shape of
125. Davis, N. F. et al. Urinary bladder vs gastrointestinal degree of thrombocytopenia. J. Ultrasound Med. 35, mesenchymal stromal cells measured by quantitative
tissue: a comparative study of their biomechanical 1849–1857 (2016). shape descriptors is determined by the stiffness of the
properties for urinary tract reconstruction. Urology 148. Kalli, M. & Stylianopoulos, T. Defining the role of solid biomaterial and by cyclic tensile forces. J. Tissue Eng.
113, 235–240 (2018). stress and matrix stiffness in cancer cell proliferation Regen. Med. 11, 3508–3522 (2017).
126. Faingold, A. et al. The effect of hydration on and metastasis. Front. Oncol. 8, 55 (2018). 176. Branco da Cunha, C. et al. Influence of the stiffness of
mechanical anisotropy, topography and fibril 149. Mancini, M. L. & Sonis, S. T. Mechanisms of cellular three-​dimensional alginate/collagen-​I interpenetrating
organization of the osteonal lamellae. J. Biomech. fibrosis associated with cancer regimen-​related networks on fibroblast biology. Biomaterials 35,
47, 367–372 (2014). toxicities. Front. Pharmacol. 5, 51 (2014). 8927–8936 (2014).
127. Milovanovic, P. et al. Age-​related deterioration in 150. Coelho, N. M. & McCulloch, C. A. Contribution 177. Xie, J. et al. Substrate elasticity regulates adipose-​
trabecular bone mechanical properties at material of collagen adhesion receptors to tissue fibrosis. derived stromal cell differentiation towards
level: nanoindentation study of the femoral neck in Cell Tissue Res. 365, 521–538 (2016). osteogenesis and adipogenesis through β-​catenin
women by using AFM. Exp. Gerontol. 47, 154–159 151. Martinez, F. J. et al. Idiopathic pulmonary fibrosis transduction. Acta Biomater. 79, 83–95 (2018).
(2012). review. Nat. Rev. Dis. Prim. 3, 17074 (2017). 178. Lv, H. et al. Biomaterial stiffness determines stem cell
128. Melo, E. et al. Effects of the decellularization method 152. Meng, X. M., Nikolic-​Paterson, D. J. & Lan, H. Y. fate. Life Sci. 178, 42–48 (2017).
on the local stiffness of acellular lungs. Tissue Eng. Inflammatory processes in renal fibrosis. 179. Sun, A. X. et al. Chondrogenesis of human bone
Part C 20, 412–422 (2014). Nat. Rev. Nephrol. 10, 493–503 (2014). marrow mesenchymal stem cells in 3-dimensional,
129. Peloso, A. et al. The human pancreas as a source of 153. Tsochatzis, E. A., Bosch, J. & Burroughs, A. K. Liver photocrosslinked hydrogel constructs: Effect of cell
protolerogenic extracellular matrix scaffold for a new-​ cirrhosis. Lancet 383, 1749–1761 (2014). seeding density and material stiffness. Acta Biomater.
generation bioartificial endocrine pancreas. Ann. Surg. 154. Li, Q., Chen, L. & Zhou, Y. Diagnostic accuracy of liver 58, 302–311 (2016).
264, 169–179 (2016). stiffness measurement in chronic hepatitis B patients 180. Hadden, W. J. et al. Stem cell migration and
130. Omidi, E. et al. Characterization and assessment of with normal or mildly elevated alanine transaminase mechanotransduction on linear stiffness gradient
hyperelastic and elastic properties of decellularized levels. Sci. Rep. 8, 5224 (2018). hydrogels. Proc. Natl Acad. Sci. USA 114,
human adipose tissues. J. Biomech. 47, 3657–3663 155. Ogawa, S. et al. Relationship between liver tissue 5647–5652 (2017).
(2014). stiffness and histopathological findings analyzed by 181. Xi, W., Saw, T. B., Delacour, D., Lim, C. T. & Ladoux, B.
131. Ocal, S., Ozcan, U. M., Basdogan, I. & Basdogan, C. shear wave elastography and compression testing Material approaches to active tissue mechanics.
Effect of preservation period on the viscoelastic material in rats with non-​alcoholic steatohepatitis. Nat. Rev. Mater. 4, 23–44 (2019).
properties of soft tissues with implications for liver J. Med. Ultrason. 43, 355–360 (2016). 182. Chaudhuri, O. et al. Substrate stress relaxation
transplantation. J. Biomech. Eng. 132, 101007 (2010). 156. Pang, J. X. Q. et al. Liver stiffness by transient regulates cell spreading. Nat. Commun. 19, 6364
132. Yamashita, J., Furman, B. R., Rawls, H. R., Wang, X. elastography predicts liver-​related complications (2015).
& Agrawal, C. M. The use of dynamic mechanical and mortality in patients with chronic liver disease. 183. Guvendiren, M. & Burdick, J. A. Stiffening hydrogels
analysis to assess the viscoelastic properties of human PLoS One 9, e95776 (2014). to probe short- and long-​term cellular responses to
cortical bone. J. Biomed. Mater. Res. 58, 47–53 157. Desai, S. S. et al. Physiological ranges of matrix dynamic mechanics. Nat. Commun. 3, 792 (2012).
(2001). rigidity modulate primary mouse hepatocyte function 184. Storm, C., Pastore, J. J., MacKintosh, F. C.,
133. Buckwalter, J. A. & Mankin, H. J. Articular cartilage: in part through hepatocyte nuclear factor 4 alpha. Lubensky, T. C. & Janmey, P. A. Nonlinear elasticity
degeneration and osteoarthritis, repair, regeneration, Hepatology 64, 261–275 (2016). in biological gels. Nature 435, 191–194 (2005).
and transplantation. Instrum. Course Lect. 47, 158. Li, Q. S., Lee, G. Y. H., Ong, C. N. & Lim, C. T. AFM 185. Kouwer, P. H. J. et al. Responsive biomimetic networks
487–504 (1998). indentation study of breast cancer cells. Biochem. from polyisocyanopeptide hydrogels. Nature 493,
134. Zhu, W., Mow, V. C., Koob, T. J. & Eyre, D. R. Biophys. Res. Commun. 374, 609–613 (2008). 651–655 (2013).
Viscoelastic shear properties of articular cartilage and 159. Gaikwad, R. M., Woodworth, C. D., Sokolov, I., 186. de Almeida, P. et al. Cytoskeletal stiffening in synthetic
the effects of glycosidase treatments. J. Orthop. Res. Subba-Rao, V. & Iyer, S. Atomic force microscopy hydrogels. Nat. Commun. 10, 609 (2019).
11, 771–781 (1993). detects differences in the surface brush of normal A seminal work showing that hydrogels with
135. Nickien, M., Thambyah, A. & Broom, N. D. How a and cancerous cells. Nat. Nanotechnol. 4, 389–393 cytoskeletal-​like stress stiffening can be obtained.
decreased fibrillar interconnectivity influences stiffness (2009). 187. Dhume, R. Y. & Barocas, V. H. Emergent structure-​
and swelling properties during early cartilage 160. Wong, R. et al. AFM-​based analysis of human metastatic dependent relaxation spectra in viscoelastic fiber
degeneration. J. Mech. Behav. Biomed. Mater. 75, cancer cells. Nanotechnology 19, 384003 (2008). networks in extension. Acta Biomater. 87, 245–255
390–398 (2017). 161. Lekka, M. Discrimination between normal and (2019).
136. Brommer, H. et al. Functional consequences cancerous cells using AFM. Bionanoscience 6, 65–80 188. Lee, H. P., Gu, L., Mooney, D. J., Levenston, M. E.
of cartilage degeneration in the equine (2016). & Chaudhuri, O. Mechanical confinement regulates
metacarpophalangeal joint: quantitative assessment 162. Maciaszek, J. L. & Lykotrafitis, G. Sickle cell trait cartilage matrix formation by chondrocytes.
of cartilage stiffness. Equine Vet. J. 37, 462–467 human erythrocytes are significantly stiffer than normal. Nat. Mater. 16, 1243–1251 (2017).
(2005). J. Biomech. 44, 657–661 (2011). 189. Kim, B. S., Nikolovski, J., Bonadio, J. & Mooney, D. J.
137. Kamiya, A. & Togawa, T. Adaptive regulation of wall 163. Lin, H.-H. et al. Mechanical phenotype of cancer cells: Cyclic mechanical strain regulates the development of
shear stress to flow change in the canine carotid cell softening and loss of stiffness sensing. Oncotarget engineered smooth muscle tissue. Nat. Biotechnol.
artery. Am. J. Physiol. 239, H14–H21 (1980). 6, 20946–20958 (2015). 17, 979–983 (1999).
138. Alkhouli, N. et al. The mechanical properties of human 164. Gilkes, D. M. et al. Hypoxia-​inducible factors mediate 190. Cochis, A. et al. Bioreactor mechanically guided 3D
adipose tissues and their relationships to the structure coordinated RhoA-​ROCK1 expression and signaling mesenchymal stem cell chondrogenesis using a
and composition of the extracellular matrix. AJP in breast cancer cells. Proc. Natl Acad. Sci. USA 111, biocompatible novel thermo-​reversible methylcellulose-
Endocrinol. Metab. 305, E1427–E1435 (2013). E384–E393 (2014). based hydrogel. Sci. Rep. 7, 45018 (2017).

www.nature.com/natrevmats
Reviews

191. Chu, S.-Y. et al. Mechanical stretch induces hair 217. Yip, A. K. et al. Anisotropic traction stresses and focal 241. Siamantouras, E., Hills, C. E., Squires, P. E. & Liu, K. K.
regeneration through the alternative activation adhesion polarization mediates topography-​induced Quantifying cellular mechanics and adhesion in renal
of macrophages. Nat. Commun. 10, 1524 (2019). cell elongation. Biomaterials 181, 103–112 (2018). tubular injury using single cell force spectroscopy.
192. Lee, J. K. et al. Tension stimulation drives tissue 218. Saruwatari, L. et al. Osteoblasts generate harder, Nanomedicine 12, 1013–1021 (2016).
formation in scaffold-​free systems. Nat. Mater. 16, stiffer, and more delamination-​resistant mineralized 242. Sun, S., Song, Z., Cotler, S. J. & Cho, M. Biomechanics
864–873 (2017). tissue on titanium than on polystyrene, associated and functionality of hepatocytes in liver cirrhosis.
193. Tsimbouri, P. M. et al. Stimulation of 3D osteogenesis with distinct tissue micro- and ultrastructure. J. Biomech. 47, 2205–2210 (2014).
by mesenchymal stem cells using a nanovibrational J. Bone Miner. Res. 20, 2002–2016 (2005). 243. Hozic, A., Rico, F., Colom, A., Buzhynskyy, N. &
bioreactor. Nat. Biomed. Eng. 1, 758–770 (2017). 219. Kalyan Phani, M., Kumar, A., Arnold, W. & Samwer, K. Scheuring, S. Nanomechanical characterization
194. Wisdom, K. M. et al. Matrix mechanical plasticity Elastic stiffness and damping measurements in of the stiffness of eye lens cells: a pilot study.
regulates cancer cell migration through confining titanium alloys using atomic force acoustic microscopy. Invest. Ophthalmol. Vis. Sci. 53, 2151–2156 (2012).
microenvironments. Nat. Commun. 9, 4144 (2018). J. Alloys Compd. 676, 397–406 (2016). 244. Kolipaka, A. et al. Magnetic resonance elastography to
195. Miotto, M. et al. 4D corneal tissue engineering: 220. Brown, A. L. et al. 22 week assessment of bladder estimate brain stiffness: measurement reproducibility
achieving time-​dependent tissue self-​curvature acellular matrix as a bladder augmentation material and its estimate in pseudotumor cerebri patients.
through localized control of cell actuators. in a porcine model. Biomaterials 23, 2179–2190 Clin. Imaging 51, 114–122 (2018).
Adv. Funct. Mater. 29, 1807334 (2019). (2002). 245. Arani, A. et al. Measuring the effects of aging and sex
196. Loebel, C., Mauck, R. L. & Burdick, J. A. Local 221. Barak, M. M. & Black, M. A. A novel use of 3D printing on regional brain stiffness with MR elastography in
nascent protein deposition and remodelling guide model demonstrates the effects of deteriorated healthy older adults. Neuroimage 111, 59–64 (2015).
mesenchymal stromal cell mechanosensing and fate trabecular bone structure on bone stiffness and 246. Ma, Z. et al. In vitro and in vivo mechanical properties
in three-​dimensional hydrogels. Nat. Mater. 18, strength. J. Mech. Behav. Biomed. Mater. 78, of human ulnar and median nerves. J. Biomed. Mater.
883–891 (2019). 455–464 (2018). Res. A 101, 2718–2725 (2013).
This article together with references 194 and 195 222. Ramadan, S., Paul, N. & Naguib, H. E. Standardized 247. Robinson, D. L. et al. Mechanical properties of normal
are important studies that serve as the foundation static and dynamic evaluation of myocardial tissue and osteoarthritic human articular cartilage. J. Mech.
for our definition of biolabile environments and properties. Biomed. Mater. 12, 025013 (2017). Behav. Biomed. Mater. 61, 96–109 (2016).
show the importance of allowing cells to ‘master 223. Yoo, L., Gupta, V., Lee, C., Kavehpore, P. & Demer, J. L. 248. Comley, K. & Fleck, N. A. A micromechanical model
their own fate’. Viscoelastic properties of bovine orbital connective for the Young’s modulus of adipose tissue. Int. J. Solids
197. Qiao, E. L., Kumar, S. & Schaffer, D. V. Mastering their tissue and fat: constitutive models. Biomech. Model. Struct. 47, 2982–2990 (2010).
own fates through the matrix. Nat. Mater. 18, Mechanobiol. 10, 901–914 (2012). 249. Savelberg, H. H. C. M., Kooloos, J. G. M., Huiskes, R.
779–780 (2019). 224. Schachar, R. A., Chan, R. W. & Fu, M. Viscoelastic & Kauer, J. M. G. Stiffness of the ligaments of the
198. Vert, M., Li, S. M., Spenlehauer, G. & Guerin, P. properties of fresh human lenses under 40 years human wrist joint. J. Biomech. 25, 369–376 (1992).
Bioresorbability and biocompatibility of aliphatic of age: implications for the aetiology of presbyopia. 250. Przybylski, G. J., Carlin, G. J., Patel, P. R. & Woo, S. L. Y.
polyesters. J. Mater. Sci. 3, 432–446 (1992). Br. J. Ophthalmol. 95, 1010–1013 (2011). Human anterior and posterior cervical longitudinal
199. Hutmacher, D. W. Scaffolds in tissue engineering bone 225. Ozawa, H., Matsumoto, T., Ohashi, T., Sato, M. & ligaments possess similar tensile properties. J. Orthop.
and cartilage. Biomaterials 21, 2529–2543 (2000). Kokubun, S. Comparison of spinal cord gray matter Res. 14, 1005–1008 (1996).
200. Deringer, V. L., Caro, M. A. & Csányi, G. Machine and white matter softness: measurement by pipette 251. Pintar, F. A. Geometric and mechanical properties of
learning interatomic potentials as emerging tools for aspiration method. J. Neurosurg. Spine 95, 221–224 human cervical spine ligaments. J. Biomed. Eng. 122,
materials science. Adv. Mater. 31, 1902765 (2019). (2001). 623–629 (2000).
201. Ansari, S., Khorshidi, S. & Karkhaneh, A. Engineering 226. Lee, L. M. & Liu, A. P. The application of micropipette 252. Arani, A. et al. Cardiac MR elastography for
of gradient osteochondral tissue: from nature to lab. aspiration in molecular mechanics of single cells. quantitative assessment of elevated myocardial
Acta Biomater. 87, 41–54 (2019). J. Nanotechnol. Eng. Med. 5, 040902 (2014). stiffness in cardiac amyloidosis. J. Magn. Reson.
202. Silva, E. D. et al. Multifunctional magnetic-​responsive 227. Moshtagh, P. R., Pouran, B., Korthagen, N. M., Imaging 46, 1361–1367 (2017).
hydrogels to engineer tendon-​to-bone interface. Zadpoor, A. A. & Weinans, H. Guidelines for an 253. Domian, I. J., Yu, H. & Mittal, N. On materials for
Nanomedicine 14, 2375–2385 (2018). optimized indentation protocol for measurement of cardiac tissue engineering. Adv. Healthc. Mater. 6,
203. Calejo, I., Costa‐Almeida, R., Reis, R. L. & cartilage stiffness: the effects of spatial variation and 1600768 (2017).
Gomes, M. E. A textile platform using continuous indentation parameters. J. Biomech. 49, 3602–3607 254. Eby, S. F. et al. Shear wave elastography of passive
aligned and textured composite microfibers to (2016). skeletal muscle stiffness: influences of sex and age
engineer tendon‐to‐bone interface gradient scaffolds. 228. Uriarte, J. J. et al. Early impairment of lung mechanics throughout adulthood. Clin. Biomech. 30, 22–27
Adv. Healthc. Mater. 8, 1900200 (2019). in a murine model of Marfan syndrome. PLoS One 11, (2015).
204. Ribeiro, V. P. et al. Enzymatically cross-​linked silk e0152124 (2016). 255. Leong, H. T., Hug, F. & Fu, S. N. Increased upper
fibroin-​based hierarchical scaffolds for osteochondral 229. Shi, Y., Glaser, K. J., Venkatesh, S. K., Ben-​Abraham, E. I. trapezius muscle stiffness in overhead athletes with
regeneration. ACS Appl. Mater. Interfaces 11, & Ehman, R. L. Feasibility of using 3D MR elastography rotator cuff tendinopathy. PLoS One 11, e0155187
3781–3799 (2019). to determine pancreatic stiffness in healthy volunteers. (2016).
205. Canadas, R. F. et al. Biochemical gradients to generate J. Magn. Reson. Imaging 41, 369–375 (2015). 256. Brandenburg, J. E. et al. Feasibility and reliability
3D heterotypic-​like tissues with isotropic and 230. Murphy, M. C. et al. Measuring the characteristic of quantifying passive muscle stiffness in young
anisotropic architectures. Adv. Funct. Mater. 28, topography of brain stiffness with magnetic resonance children by using shear wave ultrasound elastography.
1804148 (2018). elastography. PLoS One 8, e81668 (2013). J. Ultrasound Med. 34, 663–670 (2015).
206. Calejo, I., Costa-​Almeida, R., Reis, R. L. & Gomes, M. E. 231. Anvari, A., Dhyani, M., Stephen, A. E. & Samir, A. E. 257. Souron, R. et al. Sex differences in active tibialis
A physiology-​inspired multifactorial toolbox in soft-​ Reliability of shear-​wave elastography estimates anterior stiffness evaluated using supersonic shear
to-hard musculoskeletal interface tissue engineering. of the Young modulus of tissue in follicular thyroid imaging. J. Biomech. 49, 3534–3537 (2016).
Trends Biotechnol. 38, 83–98 (2019). neoplasms. Am. J. Roentgenol. 206, 609–616 258. Wang, L., Yan, F., Yang, Y., Xiang, X. & Qiu, L.
207. Grigoryan, B. et al. Multivascular networks (2016). Quantitative assessment of skin stiffness in localized
and functional intravascular topologies within 232. Dutov, P., Antipova, O., Varma, S., Orgel, J. P. R. O. & scleroderma using ultrasound shear-​wave elastography.
biocompatible hydrogels. Science 364, 458–464 Schieber, J. D. Measurement of elastic modulus of Ultrasound Med. Biol. 43, 1339–1347 (2017).
(2019). collagen type I single fiber. PLoS One 11, e0145711 259. Marinelli, J. P. et al. Quantitative assessment of lung
208. Chimene, D., Lennox, K. K., Kaunas, R. R. & (2016). stiffness in patients with interstitial lung disease using
Gaharwar, A. K. Advanced bioinks for 3D printing: 233. Li, W. et al. Fibrin fiber stiffness is strongly affected MR elastography. J. Magn. Reson. Imaging 46,
a materials science perspective. Ann. Biomed. Eng. by fiber diameter, but not by fibrinogen glycation. 365–374 (2017).
44, 2090–2102 (2016). Biophys. J. 110, 1400–1410 (2016). 260. Mariappan, Y. K. et al. Estimation of the absolute
209. Bertoldi, K., Vitelli, V., Christensen, J. & Van Hecke, M. 234. Collet, J.-P., Shuman, H., Ledger, R. E., Lee, S. & shear stiffness of human lung parenchyma using 1 h
Flexible mechanical metamaterials. Nat. Rev. Mater. Weisel, J. W. The elasticity of an individual fibrin fiber spin echo, echo planar MR elastography. J. Magn.
2, 17066 (2017). in a clot. Proc. Natl Acad. Sci. USA 102, 9133–9137 Reson. Imaging 40, 1230–1237 (2014).
210. Frenzel, T., Kadic, M. & Wegener, M. Three-​ (2005). 261. Booth, A. J. et al. Acellular normal and fibrotic
dimensional mechanical metamaterials with a twist. 235. Aaron, B. B. & Gosline, J. M. Elastin as a random‐ human lung matrices as a culture system for in vitro
Science 358, 1027–1032 (2017). network elastomer: a mechanical and optical analysis investigation. Am. J. Respir. Crit. Care Med. 186,
211. Zhao, Z., Fang, R., Rong, Q. & Liu, M. Bioinspired of single elastin fibers. Biopolymers 20, 1247–1260 866–876 (2012).
nanocomposite hydrogels with highly ordered (1981). 262. Bensamoun, S. F., Robert, L., Leclerc, G. E.,
structures. Adv. Mater. 29, 1703045 (2017). 236. Gosline, J. et al. Elastic proteins: biological roles and Debernard, L. & Charleux, F. Stiffness imaging of the
212. Chen, T., Bakhshi, H., Liu, L., Ji, J. & Agarwal, S. mechanical properties. Phil. Trans. R. Soc. Lond. B kidney and adjacent abdominal tissues measured
Combining 3D printing with electrospinning for rapid Biol. Sci. 357, 121–132 (2002). simultaneously using magnetic resonance
response and enhanced designability of hydrogel 237. Zahn, J. T. et al. Age-​dependent changes in microscale elastography. Clin. Imaging 35, 284–287 (2011).
actuators. Adv. Funct. Mater. 28, 1800514 (2018). stiffness and mechanoresponses of cells. Small 7, 263. Samir, A. E. et al. Shear wave elastography in chronic
213. Ingber, D. E. Cellular tensegrity: defining new rules 1480–1487 (2011). kidney disease: a pilot experience in native kidneys.
of biological design that govern the cytoskeleton. 238. Alcaraz, J. et al. Microrheology of human lung BMC Nephrol. 16, 119 (2015).
J. Cell Sci. 104, 613–627 (1993). epithelial cells measured by atomic force microscopy. 264. Ling, W. et al. Effects of vascularity and differentiation
214. Fabry, B. et al. Scaling the microrheology of living cells. Biophys. J. 84, 2071–2079 (2003). of hepatocellular carcinoma on tumor and liver
Phys. Rev. Lett. 87, 148102 (2001). 239. Nakamura, K. et al. Altered nano/micro-​order stiffness: in vivo and in vitro studies. Ultrasound Med.
215. Stamenović, D. et al. Rheological behavior of living elasticity of pulmonary artery smooth muscle cells Biol. 40, 739–746 (2014).
cells is timescale-​dependent. Biophys. J. 93, 39–41 of patients with idiopathic pulmonary arterial 265. Cha, S. W. et al. Nondiseased liver stiffness
(2007). hypertension. Int. J. Cardiol. 140, 102–107 (2010). measured by shearwave elastography: a pilot study.
216. Ingber, D. E., Wang, N. & Stamenović, D. Tensegrity, 240. Lulevich, V., Yang, H. Y., Isseroff, R. R. & Liu, G. Y. J. Ultrasound Med. 33, 53–60 (2014).
cellular biophysics, and the mechanics of living Single cell mechanics of keratinocyte cells. 266. Leal-​Egaña, A. et al. Tuning liver stiffness against
systems. Rep. Prog. Phys. 77, 046603 (2014). Ultramicroscopy 110, 1435–1442 (2010). tumours: an in vitro study using entrapped cells in

Nature Reviews | Materials


Reviews

tumour-​like microcapsules. J. Mech. Behav. Biomed. pathological thyroid gland in vivo. J. Magn. Reson. Acknowledgements
Mater. 9, 113–121 (2012). Imaging 30, 1151–1154 (2009). The authors gratefully acknowledge financial support from the
267. Lee, D. H., Lee, J. M., Han, J. K. & Choi, B. I. MR 273. Pozzi, R. et al. Point shear-​wave elastography in European Research Council, grant agreement ERC-2012-ADG
elastography of healthy liver parenchyma: normal chronic pancreatitis: a promising tool for staging 20120216-321266 (project ComplexiTE). C.F.G. acknowl-
value and reliability of the liver stiffness value disease severity. Pancreatology 17, 905–910 (2017). edges scholarship grant no. PD/BD/135253/2017 from
measurement. J. Magn. Reson. Imaging 38, 274. An, H., Shi, Y., Guo, Q. & Liu, Y. Test–retest reliability Fundação para a Ciência e Tecnologia. The authors also thank
1215–1223 (2013). of 3D EPI MR elastography of the pancreas. the peer-​reviewers for their constructive comments and
268. Venkatesh, S. K., Wang, G., Teo, L. L. S. & Ang, B. W. L. Clin. Radiol. 71, 1068.e7–1068.e12 (2016). suggestions that helped to shape the manuscript.
Magnetic resonance elastography of liver in healthy 275. Kolipaka, A. et al. Magnetic resonance elastography
Asians: normal liver stiffness quantification and of the pancreas: measurement reproducibility and Author contributions
reproducibility assessment. J. Magn. Reson. Imaging relationship with age. Magn. Reson. Imaging 42, 1–7 C.F.G. and L.G. contributed to all aspects of the article. R.L.R.
39, 1–8 (2014). (2017). and A.P.M. contributed substantially to discussions of the
269. Gangadhar, K., Hippe, D. S., Thiel, J. & Dighe, M. 276. Nenadic, I. et al. Noninvasive evaluation of bladder article content and review or editing of the manuscript before
Impact of image orientation on measurements wall mechanical properties as a function of filling submission. A.P.M. additionally contributed to writing the
of thyroid nodule stiffness using shear wave volume: potential application in bladder compliance manuscript.
elastography. J. Ultrasound Med. 35, 1661–1667 assessment. PLoS One 11, e0157818 (2016).
(2016). 277. Matalia, J. et al. Correlation of corneal biomechanical Competing interests
270. Brezak, R., Hippe, D., Thiel, J. & Dighe, M. K. stiffness with refractive error and ocular biometry in a The authors declare no competing interests.
Variability in stiffness assessment in a thyroid nodule pediatric population. Cornea 36, 1221–1226 (2017).
using shear wave imaging. Ultrasound Q. 31, 278. Last, J. A., Thomasy, S. M., Croasdale, C. R., Russell, P. Publisher’s note
243–249 (2015). & Murphy, C. J. Compliance profile of the human Springer Nature remains neutral with regard to jurisdictional
271. Lam, A. C. L., Pang, S. W. A., Ahuja, A. T. & cornea as measured by atomic force microscopy. claims in published maps and institutional affiliations.
Bhatia, K. S. S. The influence of precompression on Micron 43, 1293–1298 (2012).
elasticity of thyroid nodules estimated by ultrasound 279. Jardeleza, M. S. R., Daly, M. K., Kaufman, J. D., Supplementary information
shear wave elastography. Eur. Radiol. 26, 2845–2852 Klapperich, C. & Legutko, P. A. Effect of trypan blue Supplementary information is available for this paper at
(2016). staining on the elastic modulus of anterior lens https://doi.org/10.1038/s41578-019-0169-1.
272. Bahn, M. M. et al. Development and application of capsules of diabetic and nondiabetic patients.
magnetic resonance elastography of the normal and J. Cataract Refract. Surg. 35, 318–323 (2009). © Springer Nature Limited 2020

www.nature.com/natrevmats

You might also like