You are on page 1of 9

Article https://doi.org/10.

1038/s41467-024-45964-y

Elasticity-controlled jamming criticality


in soft composite solids
1
Received: 25 September 2023 Yiqiu Zhao , Haitao Hu1, Yulu Huang1, Hanqing Liu 2
, Caishan Yan 1
,
1
Chang Xu , Rui Zhang 1, Yifan Wang 3 & Qin Xu 1
Accepted: 8 February 2024

Soft composite solids are made of inclusions dispersed within soft matrices.
Check for updates They are ubiquitous in nature and form the basis of many biological tissues.
In the field of materials science, synthetic soft composites are promising can-
1234567890():,;
1234567890():,;

didates for building various engineering devices due to their highly program-
mable features. However, when the volume fraction of the inclusions increases,
predicting the mechanical properties of these materials poses a significant
challenge for the classical theories of composite mechanics. The difficulty arises
from the inherently disordered, multi-scale interactions between the inclusions
and the matrix. To address this challenge, we systematically investigated the
mechanics of densely filled soft elastomers containing stiff microspheres. We
experimentally demonstrate how the strain-stiffening response of the soft
composites is governed by the critical scalings in the vicinity of a shear-jamming
transition of the included particles. The proposed criticality framework quan-
titatively connects the overall mechanics of a soft composite with the elasticity
of the matrix and the particles, and captures the diverse mechanical responses
observed across a wide range of material parameters. The findings uncover a
novel design paradigm of composite mechanics that relies on engineering the
jamming properties of the embedded inclusions.

Dispersing nano-to-micron-sized particles within a soft polymeric gel mechanics are solely determined by the interactions between an iso-
forms soft composite solids that are widely used in various engineering lated inclusion and the surrounding matrix, which allows the effective
materials, including synthetic tissue1, wearable biomedical devices2,3, shear modulus to be described by the classical Eshelby theory14.
and soft robots4. In addition to reinforcing the polymer matrix5, the Further, modified effective medium theories have been extended to
dispersed particles can enable diverse functional features such as ani- systems with finite-density inclusions, where neighboring particles
sotropic elasticity6, shape-memory effects7,8, and stimuli-responsive interact via their induced strain fields11,15. However, this assumption of
behaviors9,10. Due to the great compliance of soft polymeric gels, the matrix-mediated, short-range interactions breaks down in the dense
embedded particles can undergo moderate displacement within the limit, where the overall stress response may involve networks of direct
matrix without causing internal fractures9. This particle rearrangement contacts16,17 or long-range rearrangements of dispersed particles12. Due
may alter both the strain couplings among neighboring inclusions11 and to the inherently disordered and heterogeneous microstructures of
the stress fields over a large length scale9,12. Compared with classical stiff dense soft composites, predicting their mechanics is challenging for
composite materials13, the current understanding of the multi-scale classical composite theories.
interactions within soft composites remains very limited. To address these issues, we systematically investigated the strain
The complexity of composite mechanics increases exponentially stiffening of soft elastomers containing a high volume fraction of stiff
with the volume fraction of the inclusions. In a dilute composite, the microspheres. Inspired by the concepts of both granular jamming18–21

1
Department of Physics, The Hong Kong University of Science and Technology, Hong Kong SAR, China. 2Theoretical Division, Los Alamos National Laboratory,
Los Alamos, NM 87545, USA. 3School of Mechanical and Aerospace Engineering, Nanyang Technological University, Singapore 639798, Singapore.
e-mail: yiqiuzhao@ust.hk; qinxu@ust.hk

Nature Communications | (2024)15:1691 1


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41467-024-45964-y

and rigidity transitions in disordered systems22–24, we demonstrate that modulus of the elastomer matrix Gm. First, at a fixed Gm = 1.28 kPa, the
the mechanical responses of soft composites are governed by relative shear modulus, Gr = G/Gm, grows more rapidly with ε as ϕ
elasticity-controlled scalings near a continuous phase transition. In the increases from 0.44 to 0.67. Second, at a fixed ϕ = 0.60, the strain
absence of matrix elasticity, the transition coincides with the shear- stiffening becomes more pronounced while Gm decreases from 1.28 to
jamming of the included particles. This novel criticality framework 0.04 kPa.
captures the stiffening responses for a variety of material parameters We define Gr, max as the relative shear modulus at the maximally
where the classical theories break down. The results provide a new stiffened states and Gr,0 as the relative shear modulus at ε = 0. Within
approach to understand the nonlinear mechanical responses of var- experimental uncertainty, Gr, max appears at approximately ε = 0.2
ious multi-phase soft materials. regardless of ϕ and Gm. Therefore, we estimated Gr, max for all the
samples using the values of Gr at ε = 0.2. For ε > 0.2, Gr decreases with ε,
Results and the composites were unable to fully recover their original shapes
Strain-stiffening responses of soft composite solids after the compressions were removed. This plasticity is likely caused
We prepared compliant polydimethylsiloxane (PDMS) elastomers fil- by internal fractures between the elastomer and the particles27. Since
led with stiff polystyrene (PS) microspheres having an average dia- the adhesion energy at gel interfaces is approximately independent of
meter of 30 μm (Fig. 1a and Supplementary Fig. 8). While the shear the crosslinking density28, the plasticity onset (ε ≈ 0.2) remains nearly
modulus of the PS spheres is Gp = 1.6 GPa (Supplementary Fig. 1), the unchanged for various Gm. In contrast, the plots of Gr(ε) appear to be
shear modulus of the PDMS matrix was systematically varied from highly reproducible when the compressions are released at ε < 0.2.
Gm = 0.04 to 4 kPa by tuning the crosslinking density (ref. 25 and In this study, we focus exclusively on the stiffening regime between
Supplementary Fig. 7). The mechanical properties of the soft compo- ε = 0 and 0.2.
sites were characterized using a rheometer equipped with a parallel- Figure 1d shows both Gr, max (solid points) and Gr,0 (hollow points)
plate shear cell (Fig. 1b). The top plate controls the gap size (d) and as a function of ϕ as Gm varies between 0.04 and 1.28 kPa. For ϕ < 0.4,
applies axial compressive strains (ε) (Supplementary Fig. 9). Due to the only Gr,0 was reported since no strain-stiffening was found. For com-
incompressibility of crosslinked PDMS gels26, the volume of the sample parison with the classical theories of composite mechanics, we plotted
remains unchanged under axial compression (Supplementary Fig. 12 the predictions from the Eshelby theory14 and the Mori–Tanaka
and Supplementary Movie 1), which gives rise to pure shear. At each approximation scheme15, which align well with the Gr,0 measured in the
given ε, the rheological properties of the composites were measured dilute limit (ϕ < 0.2). However, for dense composites (ϕ > 0.4), the
using an oscillatory shear with a small amplitude (δγa = 0.01 %). At an classical theories significantly deviate from the measured Gr, max and
angular frequency (ω = 0.1 rad/s), the storage modulus has reached a Gr,0, and also fail to describe the strain-dependent shear modulus Gr(ε).
low-frequency plateau (Supplementary Fig. 11), which indicates the These mismatches suggest that potential mechanisms, such as direct
shear modulus of soft composites (G). This resulting G(ε) represents contact between inclusions16,17, were overlooked in the classical mod-
the linear elastic response of the soft composites in differently sheared els of the mechanics of dense soft composites.
states (Supplementary Fig. 10).
The dense soft composites exhibit characteristic strain-stiffening Signatures of jamming-controlled elasticity
responses under the axial compressions (Fig. 1c). The stiffening degree We re-examine the super-exponential rise of Gr, max in Fig. 1d. As Gm
is determined by both the particle volume fraction ϕ and the shear decreases, the growth of Gr, max becomes increasingly more divergent

PDMS 0.67

PS 30
0.60
0.65
μm
0.60
0.63

0.60
0.57
0.55
0.51
0.49
0.44

Rp Rs

Fig. 1 | Strain-stiffening of PS-PDMS soft composites under volume-conserving In addition, the hollow red squares and hollow yellow triangles represent the
compressions. a Schematic of the cross-section of PS-PDMS composites. For a results of Gr(ε) at the same ϕ = 0.60 but for different matrix moduli, Gm = 0.12 kPa
predetermined volume fraction ϕ, polydisperse PS spheres with an average dia- and Gm = 0.04 kPa, respectively. d Comparison between the experimentally mea-
meter of approximately 30 μm are well dispersed in a crosslinked PDMS matrix. sured Gr and the predictions from the classical theories of composite mechanics.
b Schematic of the experimental setup used to characterize the strain-stiffening of The solid and hollow points indicate Gr, max and Gr,0, respectively, versus ϕ for
the soft composites. The top plate moves down in a stepwise manner to apply an samples with varying Gm. The error bars represent the standard deviation from
axial strain ε. At each ε, the linear shear modulus G was measured through an measuring two to five independently fabricated samples. The two dashed gray lines
oscillatory shear with a strain amplitude of δγa = 10−4 and an angular frequency of represent the predictions from the Eshelby theory and the Mori–Tanaka approx-
ω = 0.1 rad/s. c Plots of the relative shear modulus, Gr = G/Gm, against ε for various imation. The pink area represents the range of the volume fraction where strain-
particle volume fractions ϕ and matrix shear moduli Gm. The blue hollow circles stiffening was observed.
indicate the results for a constant Gm = 1.28 kPa as ϕ increases from 0.44 to 0.67.

Nature Communications | (2024)15:1691 2


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41467-024-45964-y

the Krieger–Dougherty relation29,30

ηðϕÞ
ηr ðϕÞ = = ð1  ϕ=ϕJ Þγ , ðϕ < ϕJ Þ ð1Þ
ηs

with a fixed exponent γ = 2 and a fitted jamming volume fraction


ϕJ = 0.594 ± 0.003. A similar scaling has been identified in the simu-
lations of over-damped granular systems near jamming18,31. For ϕ > ϕJ,
Flowing Jammed we did not observe homogeneous steady shear flow at any shear rate.
states states
Instead, the suspensions were consistently jammed under contin-
uous shear (Supplementary Fig. 5). To quantify the mechanical
responses of these shear-jammed states, we initially prepared fully
relaxed suspensions without rigidity at ε = 0. Subsequently, we
applied an axial strain ε > 0 to induce jamming in the suspensions
and measure their shear moduli. The measurement protocol is
detailed in Supplementary Information Section I.A.3. The right panel
of Fig. 2a represents the nonzero shear moduli of the shear-jammed
PS-PDMS suspensions (Gs) measured at ε = 0.2 in the regime of ϕ > ϕJ.
Since no significant change in Gs was found when ε was further
increased (Supplementary Fig. 4), ϕJ = 0.594 represents the lowest
particle volume fraction required to achieve shear-jamming in the
PS-PDMS suspensions.
In Fig. 2b, we plot ηr(ϕ) from Eq. (1) together with Gr, max ðϕÞ of the
composites for a comparison. The traces of Gr, max gradually converge to
ηr as Gm decreases, suggesting that Gr, max ≈ ð1  ϕ=ϕJ Þγ for ϕ < ϕJ as Gm
approaches zero, and the actual shear modulus Gmax scales linearly with
Gm in this limit. In contrast, for ϕ > ϕJ, Gmax becomes independent of Gm
(Fig. 2c) and is close to the value of Gs measured independently from
the jammed suspensions (Fig. 2a), suggesting a particle-dominated
response. Considering the contrasting mechanical behaviors exhibited
for the ranges ϕ < ϕJ and ϕ > ϕJ, it is likely that the shear-jamming point
of the suspensions controls a crossover of the mechanical properties of
the composites.

Fig. 2 | Signatures of jamming-controlled elasticity. a Rigidity transition of PS


Elasticity-controlled criticality near jamming
particles suspended in un-crosslinked silicone oil. The black triangles show the Since the plots of Gmax ðϕÞ in Fig. 2c resemble the critical behaviors near
relative viscosity ηr = η/ηs in the low-stress Newtonian regime for different particle a continuous phase transition23,24,32,33, we next investigate the scalings
volume fractions ϕ = 0.45, 0.49, 0.53, 0.55. The dashed black curve indicates the of the composite shear modulus (Gmax ) near (ϕ = ϕJ, Gm = 0). Motivated
best fit of the experimental results to Eq. (1) where ϕJ = 0.594. In the regime of by the observation that Gr, max approaches ηr as Gm → 0 (Fig. 2b) and the
ϕ > ϕJ, the suspensions were initially unjammed at ε = 0, and then shear jammed at classical analogy between the effective shear modulus and the shear
ε = 0.2. The red crosses represent their shear moduli (Gs) in the shear-jammed viscosity in multi-phase systems34–36, we conjecture the scaling law
states (ε = 0.2). The error bars represent the standard deviation from five inde-
pendent measurements. b Plots of Gr, max against ϕ for different Gm values. To Gmax ðϕÞ ηðϕÞ
compare the absolute values of Gr, max with ηr, the fit to Eq. (1) obtained in panel a is lim = = ð1  ϕ=ϕJ Þγ , ðϕ < ϕJ Þ ð2Þ
Gm !0 Gm ηs
also shown by the dashed gray line in the same plot. c The actual shear modulus
Gmax is plotted against ϕ for different Gm values based on the results in (b). The
with γ = 2 and ϕJ = 0.594. To demonstrate the validity of this scaling
solid lines in both panels b and c are predictions from the scaling model based on
jamming criticality (Eqs. (4) and (5)). The error bars in b and c represent the stan-
assumption, we plot Gr, max against ϕJ − ϕ in Fig. 3a with different Gm
dard deviation from measuring two to five independently fabricated samples. values, where the results show the best agreement with Eq. (2) for the
softest matrix. We further consider how Gmax varies with Gm at ϕ = ϕJ.
In Fig. 3b, Gmax is plotted at ϕ = 0.59 ≈ ϕJ against Gm, which can be fitted
to the power-law scaling
near ϕ ≈ 0.6. Since a soft composite solid will asymptotically become a
granular suspension as the matrix elasticity approaches zero, we Gmax ∼ G1=δ ð3Þ
m , ðϕ = ϕJ Þ
hypothesize an underlying connection between the shear-jamming of
dense suspensions and the strain-stiffening of soft composites in the with a fitted exponent 1/δ = 0.6 ± 0.1.
limit of Gm → 0. Considering the scalings in Eqs. (2) and (3), we compare the soft
To validate this assumption, we first characterize the shear composites near ϕJ with a ferromagnetic system near the Curie tem-
rheology of a concentrated PS suspension in the PDMS base solution perature (Tc). The material parameters of the soft composites
without any crosslinkers. We define the relative viscosity (ηr) as the ðGmax , Gm , ϕ  ϕJ Þ are directly analogous to (M, H, T−Tc) in the Ising
ratio of the viscosity of the suspension (η) to that of the PDMS base model. By assuming a scale-invariant free energy at the critical point
(ηs = 1.0 Pa s): ηr = η/ηs. The left panel in Fig. 2a shows ηr measured (ϕ = ϕJ, Gm = 0), we propose a universal scaling form
within a Newtonian regime where the shear stress ranges from 1 to
10 Pa. This value effectively estimates the suspension viscosity in the !
Gm
quasi-static limit. The details of the rheological measurements are Gmax = j1  ϕ=ϕJ jβ f ± ð4Þ
provided in Supplementary Fig. 2. The results are well described by j1  ϕ=ϕJ jΔ

Nature Communications | (2024)15:1691 3


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41467-024-45964-y

material constants c1 = 1.4 and c2 = 1.3 for the PS–PDMS composites.


With all the essential parameters (ϕJ, β, Δ, c1, and c2), we can cal-
culate Gmax for a given ϕ and Gm. For instance, the colored solid
lines in Fig. 2b and c represent the theoretical predictions from
Eqs. (4) and (5).

Criticality near a strain-dependent jamming transition


To describe the entire strain-stiffening regime, it is necessary to
expand the scaling analysis to include the axial strains ranging from
ε = 0 to 0.2. Since the shear-jamming point of granular materials
depends on strain37–45, we next explore an extension to our model
by incorporating a strain-dependent jamming volume fraction ϕJ(ε)
for 0 ≤ ε ≤ 0.2.
Motivated by the previous simulation showing the similar sym-
metry between shear-jamming and isotropic jamming transitons40, we
assume that the critical exponents (β = 3 and Δ = 5) and the material
parameters (c1 = 1.4 and c2 = 1.3) of dense composites remain constant
for different ε. Therefore, Eq. (5) is rewritten as

~ = g ðm ~ Δ=β
~ ε Þ = c1 m ~ ðΔ1Þ=β ∓m
~ ε, ð6Þ
h ε ± ε ∓c2 mε

~  G G1 j1  ϕ=ϕ ðεÞjΔ and m ~ ε  GðεÞG1 β


where h ε m p J p j1  ϕ=ϕJ ðεÞj .
For each ε, we search for an optimal ϕJ(ε) that allows the composite
shear modulus G(ε) measured with different Gm and ϕ to be collapsed
onto Eq. (6) (the dashed gray line in Fig. 4a). As a consequence, we are
able to overlay G(ε) measured within the range of 0 ≤ ε ≤ 0.2 by plotting
Fig. 3 | Elasticity-controlled criticality near jamming. a Plots of Gr, max against G(ε)/(Gp∣1 − ϕ/ϕJ(ε)∣β) versus Gm/(Gp∣1−ϕ/ϕJ(ε)∣Δ). The resulting ϕJ(ε) in
ϕJ−ϕ for Gm = 0.12, 0.36, and 1.28 kPa, respectively. The dashed gray line indicates Fig. 4b can be fitted to a form that describes the shear-jamming phase
the scaling law of Eq. (2). b Plots of Gmax versus Gm for composites with boundary of granular materials39,42,46
ϕ = 0.59 ≈ ϕJ, where the dashed black line represents the scaling law of Eq. (3).
c Scaling collapse of Gmax , normalized by ∣1−ϕ/ϕJ∣β, as a function of Gm/∣1−ϕ/ϕJ∣Δ ϕJ ðεÞ = ϕm + ðϕ0  ϕm Þeε=ε
*
ð7Þ
with ϕJ = 0.594, β = 3, and Δ = 5. The solid markers represent the experimental
results obtained for ϕ > ϕJ, and the open markers represent the results obtained for with ϕ0 = 0.688 ± 0.004, ϕm = 0.594 ± 0.002, and a characteristic
ϕ < ϕJ. The data points are labeled with different colors based on ϕ. The dashed red
strain scale ε* = 0.035 ± 0.003. While ϕm agrees with ϕJ = 0.594 mea-
and blue curves are the best fits to the equations of state (Eq. (5)) for the experi-
sured under the steady-state rheology of the PS-PDMS suspensions
mental results within ϕ > ϕJ and ϕ < ϕJ, respectively. All error bars represent stan-
shown in Fig. 2a, ϕ0 is consistent with the simulated random close
dard deviations from measuring two to five independently fabricated samples.
packing of spheres having the same size distribution as our samples
(Supplementary Fig. 6). Although Eq. (7) was obtained from the scaling
where β = γ/(δ − 1) = 3.0 ± 0.7 and Δ = δβ = 5.0 ± 1.1, and the crossover behaviors of soft composites, it effectively predicts the line of rigidity
scaling functions f+ and f− apply to the regimes of ϕ > ϕJ and ϕ < ϕJ, transitions for the PS-PDMS suspensions in our experiments (Supple-
respectively. The derivations of Eq. (4) and the relationships between mentary Fig. 4).
the exponents are described in the “Methods” section. A similar scaling To test the universality of the scaling model, we further examined
was previously applied to study fibrous networks near central force a different composite system made by dispersing glass beads in PDMS
rigidity transitions, where the bending rigidity plays a similar role as Gm matrices. The size of these glass beads is similar to that of the
in soft composites23,33. PS particles but their shear modulus is ten times higher; that is,
To test our scaling ansatz (Eq. (4)), the mechanical responses Gp = 15.8 GPa. The results of the glass–PDMS composites are collapsed
(Gmax ) measured for different Gm and ϕ are plotted in Fig. 3c using the onto the same plot in Fig. 4a with the same critical exponents β = 3
rescaled variables Gmax =j1  ϕ=ϕJ jβ and Gm /∣1−ϕ/ϕJ∣Δ. The range of Gm and Δ = 5 but different material constants c1 = 0.9 and c2 = 0.8.
spans two orders of magnitude, from 0.04 to 4.0 kPa, while ϕ increases The difference in c1 and c2 is likely due to the high bonding energy
from 0.45 to 0.67 around ϕJ = 0.594. Consistent with Eq. (4), the data between glass and PDMS. The resulting ϕJ(ε) was also fitted to Eq. (7)
points for ϕ > ϕJ and ϕ < ϕJ are nicely collapsed onto two distinct with ϕ0 = 0.676 ± 0.003, ϕm = 0.613 ± 0.003, and ε* = 0.040 ± 0.007.
branches. The ϕ < ϕJ branch exhibits a slope close to 1, indicating that We again found that ϕm = 0.613 is consistent with the shear-jamming
Gmax ∼ Gm . The ϕ > ϕJ branch reaches a plateau independent of Gm, point of the glass–PDMS suspensions and that ϕ0 = 0.676 is consistent
suggesting that Gmax is dominated by the particle phase. In the limit of with the predicted random close packing. The variation in ϕm may
ϕ ≈ ϕJ, a critical regime emerges where the two branches become stem from the difference in frictional coefficients and polydispersities
β=Δ
indistinguishable and both follow the same scaling, Gmax ∼ G m ∼ G0:6 m . between the PS and glass particles.
To model Gmax analytically, we derived an explicit form of the With the given parameters Gm, ϕ and ε, we can calculate the shear
equations of state modulus of soft composites as
!
~ = g ðmÞ
h ~ Δ=β ∓c2 m
~ = c1 m ~ ðΔ1Þ=β ∓m
~ ð5Þ Gm =Gp
±
Gðε, ϕ, Gm Þ = Gp j1  ϕ=ϕJ ðεÞjβ f ± , ð8Þ
j1  ϕ=ϕJ ðεÞjΔ
where g± are the inverse functions of f±. The reduced variables
~  G G1 j1  ϕ=ϕ jΔ and m~  Gmax G1 β
h m p J p j1  ϕ=ϕJ j were used to where ϕJ(ε) is given by Eq. (7), and the functions f± can be evaluated by
simplify the notation. The derivation is detailed in the “Methods” numerically solving the inverse functions g± in Eq. (6). In Fig. 4c, we
section. By fitting the data in Fig. 3c to Eq. (5), we obtain the compared the measured shear moduli of two sets of PS-PDMS samples

Nature Communications | (2024)15:1691 4


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41467-024-45964-y

Jammed

0.6
Gla
ss
PS

Unjammed

Particle- l
tica
-Dominant Cri

1
- nt
trix ina
a
M om
-D Glass PS

Fig. 4 | Criticality near a strain-dependent jamming transition. a Collapse of the (open black circles) and glass-PDMS systems (gray uptriangles). The error bars
rescaled composite shear modulus G(ε)/(Gp∣1−ϕ/ϕJ(ε)∣β) as a function of the indicate the fitting uncertainties. The solid black and gray curves represent the best
rescaled matrix shear modulus Gm/(Gp∣1−ϕ/ϕJ(ε)∣Δ) at different axial strains (ε) with fits of ϕJ(ε) to Eq. (7) for these two material systems, respectively. The pink area
two fixed critical exponents β = 3 and Δ = 5. The data points include the experi- indicates the shear-jammed phase of the PS-PDMS suspensions. c Plots of the shear
mental results obtained for two composite systems: PS-PDMS and glass-PDMS soft modulus of PS-PDMS composites (G) as a function of both ε and ϕ. The blue and
composites. The dotted gray (and pink) curves represent the best fit to Eq. (6) for pink connected points represent the experimental results for Gm = 1.28 and
the PS-PDMS (and glass-PDMS) composites. The vertical dashed line (Gm/ 0.12 kPa, respectively. The blue and pink surfaces represent the theoretical pre-
Gp = ∣1−ϕ/ϕJ(ε)∣Δ) approximates the crossover boundary from the critical regime to dictions from Eq. (8) for these two Gm values.
the particle- or matrix-dominated regime. b Plots of the fitted ϕJ(ε) for PS-PDMS

with Gm = 0.12 and 1.28 kPa, respectively, to the theoretical predictions


from Eq. (8).
The phase diagram in Fig. 5 summarizes the fundamental aspects
of our criticality framework. The Gm = 0 plane represents the granular
suspensions consisting of particles in uncrosslinked polymers. The
solid red curve within the plane, ϕ = ϕJ(ε), denotes the boundary of the
shear-jamming transition21,37. Soft composites exist in the 3D space
characterized by Gm > 0, and the vertical planes in Fig. 5 represent the
cross sections of this space at different strains. While there is no
rigidity transition in this space with Gm > 0, the mechanics are deter-
mined by the critical scalings near ϕJ(ε). When Gm /Gp ≪ ∣1−ϕ/ϕJ(ε)∣Δ, a
soft composite resides either in a matrix-dominated regime if ϕ < ϕJ or
in a particle-dominant regime if ϕ > ϕJ. As Gm approaches zero,
Gðε, ϕÞ = Gm ð1  ϕ=ϕJ ðεÞÞγ for ϕ < ϕJ, or Gðε, ϕÞ = CGp j1  ϕ=ϕJ ðεÞjβ for
ϕ > ϕJ, where C is a prefactor depending on the material parameters c1
and c2. When 1 ≫ Gm/Gp ≫ ∣1−ϕ/ϕJ(ε)∣Δ, a soft composite is anticipated
Fig. 5 | Phase diagram of the mechanical responses of soft composite solids and β=Δ β=Δ 1β=Δ
granular suspensions. The Gm = 0 plane represents the suspensions consisting of to be in the critical regime, where G = c1 G m Gp .
particles dispersing in uncrosslinked polymers. The solid red line in the Gm = 0
plane signifies the shear-jamming transition (ϕJ(ε)) of dense suspensions37. The 3D Discussion
space defined by Gm > 0 represents soft composites consisting of particles dis- The study reveals the essential role of shear-jamming in the mechanics
persing in crosslinked polymeric elastomers. The mechanical properties of dense of soft composites in the dense limit, a regime where the system
soft composites under different strains ε are controlled by the scalings (Eq. (8)) near becomes highly responsive and promises wide-ranging applications,
the critical line ϕJ(ε). The dashed red lines Gm/Gp = ∣1−ϕ/ϕJ(ε)∣Δ indicate the cross- yet remains challenging to model using conventional tools from con-
over boundary from the matrix- or particle-dominated regime to the critical tinuous mechanics. We show that the strain-stiffening of soft compo-
regime. The solid arrow (A → B → C) illustrates a representative strain-stiffening sites can be interpreted as a manifestation of the criticality near a
process of soft composites with a particle volume fraction ϕm < ϕ < ϕ0. With the strain-dependent jamming point of dense suspensions (Fig. 5). The
increase in the applied strain ε, the mechanical response of the composites crosses
efficacy of our scaling model reveals the unique mechanical features of
over from the matrix-dominated regime (ε = 0) to the critical regime (ε = ε1), and
soft composites. As Gm decreases to the order of 101–102 Pa, the PDMS
finally to the particle-dominated regime (ε = ε2).
matrix consists of both a weakly crosslinked network and a substantial

Nature Communications | (2024)15:1691 5


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41467-024-45964-y

amount of uncrosslinked free chains. The characteristic pore size of be expressed as


1=3
the network can be estimated as a ∼ ðk B T=GÞ ∼ 50 nm47. Therefore,
y
y G
the particles included in the PDMS matrix can potentially move to FðΦ, GÞ = jΦjyΦ F~ ± ðjΦj
d
Φ GÞ: ð10Þ
create direct contacts without causing fractures in the network. Con-
sequently, the contact network within soft composites may resemble where F~ + and F~  are different forms of the free energy in the regimes
that in shear-jammed granular systems as Gm approaches zero. of Φ > 0 and Φ < 0, respectively.
From the perspective of materials science, the study will benefit For a composite with Gm > 0, the parameters fG, Gm g are analo-
materials design in tissue engineering. Strain-stiffening has been gous to {M, H} in the Ising model. We define Gm  Gm =Gp as the
widely observed in both biological48 and synthetic tissues10,16, with the dimensionless shear modulus of the elastomer matrix. To transform
prevailing interpretations attributing them to the nonlinear mechanics the variable by substituting fΦ, Gg with fΦ, Gm g, we minimize the fol-
of the fibrous networks in the matrix. The significance of direct con- lowing Legendre transformation function:
tacts between inclusions and the associated jamming transition in soft
matrices began attracting attention only recently16,49. A key difference LðΦ, Gm Þ = minfFðΦ, GÞ  Gm Gg: ð11Þ
G
between our experiments and previous studies10,48,49 is that the strain
stiffening in our study occurs without increasing the volume fraction In soft composites, FðΦ, GÞ and LðΦ, Gm Þ are in direct analogy to
and thus cannot be explained by the model in ref. 16. The connection the Helmholtz free energy and Gibbs free energy in thermodynamic
between strain-stiffening in incompressible soft composites and shear- systems. The explicit evaluation of Eq. (11) leads to
jamming in dense suspensions offers a new scheme for designing the
0
tissue-like mechanics of soft composites. Gm = jΦjΔ F~ ± ðjΦjβ GÞ, ð12Þ

Methods where Δ ≡ (d − yG)/yΦ, and β ≡ yG/yΦ. By defining f± as the inverse


0
Material preparation functions of F~ ± , we obtain the scaling form of the equations of state
Both the PS and glass particles are micron-sized spheres with size shown in Eq. (5) of the main text:
distributions that can be described by the log-normal function
  2 
1 lnðr=r 0 Þ G = jΦjβ f ± ðGm jΦjΔ Þ: ð13Þ
f ðrÞ = pffiffiffiffi
1ffi
2π σr
exp  2 σ . For the PS particles, r0 = 12 μm and
σ = 0.6. For the glass particles, r0 = 20 μm and σ = 0.5. The shear For Φ < 0, we have
modulus of the particles, Gp, was measured by compressing individual
beads between two flat substrates using a nanoindenter (Bruker, G ∂G 0
lim ∼ = jΦjβΔ f ± ðjΦjΔ Gm Þ ∼ jΦjβΔ : ð14Þ
Hysitron TI-980). The resulting force-displacement curves were fitted Gm !0 Gm ∂Gm
to the Hertzian contact model (see Supplementary Fig. 1b). The results
showed that Gp = 1.6 and 15.8 GPa for the PS and the glass particles, Compared with Eq. (2) in the main text, we have γ = Δ − β.
β=Δ
respectively. In addition, Eq. (13) suggests that f ± ðGm jΦjΔ Þ / ðGm jΦjΔ Þ at
the critical point Φ = 0 to prevent the divergence of free energy.
The PDMS matrix was made by mixing a silicone base vinyl- Therefore, we have
terminated polydimethylsiloxane (DMS-V31, Gelest Inc) with copoly-
β=Δ
mer crosslinkers (HMS-301, Gelest Inc) and a catalyst complex in GðΦ = 0Þ ∼ Gm : ð15Þ
xylene (SIP6831.2, Gelest Inc.). We prepared two mixture solutions,
Gelest Part A and Gelest Part B, before curing. In particular, Part A Compared with Eq. (3) in the main text, we obtain δ = Δ/β.
consisted of a silicone base with 0.005 wt% catalyst, and Part B con- We next derive the explicit form of the equation of states in Eq. (5)
sisted of a silicone base with 10 wt% crosslinkers. By changing the in the main text. Based on the scale-invariant expression of Eq. (9), the
weight ratio of A to B from 14.5:1 to 8:1, we varied Gm from 0.04 expansion of FðΦ,GÞ should comprise terms Φa Gb with ayΦ + byG = d.
to 4 kPa. Therefore, F can be expressed as
We prepared disk-shaped composite samples with 10 mm radius
and 10 mm height in an acrylic mold covered with a para-film. To fully X dai yΦ X Δai
+1
FðΦ, GÞ = μi, ± jΦjai G yG
= μi, ± jΦjai G β , ð16Þ
relax the internal structures, we used a vortex mixer (BV1000, i i
Benchmark Scientific Inc.) to vibrate the samples immediately after
mixing all the components. For ϕ > 0.5, we compressed the samples where ai > 0, and μi,± are the expansion coefficients for Φ > 0 and Φ < 0.
using a glass plate to flatten the top surface. Each sample was then left By evaluating the variation in Eq. (11), we obtain
to cure at room temperature for at least 48 h.
X Δ  ai + β Δai
Gm = μi, ± jΦjai G β : ð17Þ
Criticality analysis i
β
We first show how the scaling form of the equations of the state shown
in the main text Eq. (4) can be obtained by minimizing a scale-invariant The above equation can be further simplified by including only
phenomenological free energy. Denote the singular part of the free three terms to describe the key experimental observations. First,
energy of a dense granular suspension (Gm = 0) under a given axial λ = μ0(Δ + β)/β > 0 when a0 = 0 to ensure that the free energy is mini-
strain ε as FðΦ,GÞ, where G  G=Gp is the dimensionless shear modulus, mum at G = 0 while Φ = 0. Second, μ01, ± = μ1, ± ðΔ + β  1Þ=β ≠ 0 when
and Φ ≡ ϕ/ϕJ(ε)−1 is the reduced volume fraction. For a given length a1 = 1 to ensure that ∂GðΦÞ=∂Φ ≠ 0 at Φ = 0. Finally, because G / Gm
scale l, we assume that the free energy is self-similar near the critical in the matrix-dominated regime, we have μ02, ± = 2μ2, ± ≠ 0 when
point Φ = 0, a2 = Δ − β. As a consequence, Gm can be simplified as

d y y Δ Δ1
FðΦ, GÞ = l Fðl Φ Φ, l G GÞ, ð9Þ Gm = λG β + μ01, ± jΦjG β + μ02, ± jΦjΔβ G: ð18Þ

where d = 3 is the space dimension, and yΦ and yG are the scaling Due to the intrinsic nature of a continuous phase transition at
1
dimensions of Φ and G, respectively. Considering l = jΦj yΦ , Eq. (9) can Φ = 0, we have μ0i, ± = ∓μ0i with μ0i >0 for both i = 1 and 2. By defining

Nature Communications | (2024)15:1691 6


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41467-024-45964-y

the reduced variables m ~  G =jΦjΔ , Eq. (18) can be


~  G=jΦjβ and h 14. Eshelby, JohnDouglas & Peierls, RudolfErnst The determination of
m
rewritten as the elastic field of an ellipsoidal inclusion, and related problems.
Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 241, 376–396 (1957).
~ =c m
h
Δ
~ β ∓c2 m
Δ1
~ β ∓c3 m,
~ ð19Þ 15. Mori, T. & Tanaka, K. Average stress in matrix and average elastic
1
energy of materials with misfitting inclusions. Acta Metall. 21,
where c1 = λ, c2 = μ01 , and c3 = μ02 . In the regime of Φ < 0, we experi- 571–574 (1973).
mentally observed h= ~ m ~ = 1 as m
~ ! 0, suggesting that c3 = 1. Therefore, 16. Shivers, J. L. et al. Compression stiffening of fibrous networks with
we finally obtain stiff inclusions. Proc. Natl Acad. Sci. USA 117, 21037–21044 (2020).
17. Phan-Thien (Phan Thiên Nhân), N., Kim, S. & Wang, S. Finite defor-
~ =c m
h
Δ
~ β ∓c2 m
Δ1
~
~ β ∓m, ð20Þ mation of a random array of rigid spheres in an elastic matrix at high
1
concentration. Phys. Fluids 33, 113314 (2021).
which is Eq. (5) in the main text. 18. Olsson, P. & Teitel, S. Critical scaling of shear viscosity at the jam-
In the particle-dominated regime, when Φ > 0 and h ~ = 0, the ming transition. Phys. Rev. Lett. 99, 178001 (2007).
nonzero solution of m ~ from Eq. (20) gives the prefactor C in the scaling 19. Liu, A. J. & Nagel, S. R. The jamming transition and the marginally
of the shear modulus G = CGp j1  ϕ=ϕJ jβ . The value C can be obtained jammed solid. Annu. Rev. Condens. Matter Phys. 1, 347–369 (2010).
by solving 20. Zaccone, A. & Scossa-Romano, E. Approximate analytical descrip-
tion of the nonaffine response of amorphous solids. Phys. Rev. B 83,
Δ Δ1
c1 C β1  c2 C β
1
 1 = 0, ð21Þ 184205 (2011).
21. Behringer, R. P. & Chakraborty, B. The physics of jamming for
and is thus determined by both c1 and c2. granular materials: a review. Rep. Prog. Phys. 82, 012601 (2018).
In the critical regime, as Φ = 0 and both m ~ ! 1 and h ~ ! 1, 22. Lacasse, Martin-D., Grest, G. S., Levine, D., Mason, T. G. & Weitz, D.
~ ~ Δ=β β=Δ β=Δ 1β=Δ
Eq. (20) reduces to h = c1 m , which gives G = c1 G m Gp . A. Model for the elasticity of compressed emulsions. Phys. Rev. Lett.
76, 3448–3451 (1996).
Data availability 23. Broedersz, C. P., Mao, X., Lubensky, T. C. & MacKintosh, F. C.
All the data supporting the findings of this study are available within Criticality and isostaticity in fibre networks. Nat. Phys. 7, 983–988
the main text and the Supplementary Information. The source data (2011).
used in generating all the figures are provided in the Source Data 24. Bi, D., Lopez, J. H., Schwarz, J. M. & Manning, M. L. A density-
file. Source data are provided with this paper. independent rigidity transition in biological tissues. Nat. Phys. 11,
1074–1079 (2015).
References 25. Zhao, W., Zhou, J., Hu, H., Xu, C. & Xu, Q. The role of crosslinking
1. Gosselin, E. A., Eppler, H. B., Bromberg, J. S. & Jewell, C. M. density in surface stress and surface energy of soft solids. Soft
Designing natural and synthetic immune tissues. Nat. Mater. 17, Matter 18, 507–513 (2022).
484–498 (2018). 26. Pritchard, R. H., Lava, P., Debruyne, D. & Terentjev, E. M. Precise
2. Koydemir, HaticeCeylan & Ozcan, A. Wearable and implantable determination of the Poisson ratio in soft materials with 2D digital
sensors for biomedical applications. Annu. Rev. Anal. Chem. 11, image correlation. Soft Matter 9, 6037–6045 (2013).
127–146 (2018). 27. Francqueville, Foucaultde, Gilormini, P., Diani, J. & Vandenbroucke,
3. Ray, T. R. et al. Bio-integrated wearable systems: a comprehensive A. Relationship between local damage and macroscopic response
review. Chem. Rev. 119, 5461–5533 (2019). of soft materials highly reinforced by monodispersed particles.
4. Miriyev, A., Stack, K. & Lipson, H. Soft material for soft actuators. Mech. Mater. 146, 103408 (2020).
Nat. Commun. 8, 596 (2017). 28. Style, R. W., Hyland, C., Boltyanskiy, R., Wettlaufer, J. S. & Dufresne,
5. Bergström, J. örgenS. & Boyce, M. C. Mechanical behavior of par- E. R. Surface tension and contact with soft elastic solids. Nat.
ticle filled elastomers. Rubber Chem. Technol. 72, 633–656 Commun. 4, 2728 (2013).
(1999). 29. Guazzelli, E. & Pouliquen, O. Rheology of dense granular suspen-
6. Wang, X. et al. Mechanical nonreciprocity in a uniform composite sions. J. Fluid Mech. 852, 1 (2018).
material. Science 380, 192–198 (2023). 30. Morris, J. F. Shear thickening of concentrated suspensions: Recent
7. Testa, P. et al. Magnetically addressable shape-memory and developments and relation to other phenomena. Annu. Rev. Fluid
stiffening in a composite elastomer. Adv. Mater. 31, 1900561 Mech. 52, 121–144 (2020).
(2019). 31. Hatano, T. Scaling properties of granular rheology near the jam-
8. Xia, Y., He, Y., Zhang, F., Liu, Y. & Leng, J. A review of shape memory ming transition. J. Phys. Soc. Jpn. 77, 123002 (2008).
polymers and composites: mechanisms, materials, and applica- 32. Cardy, J. Scaling and Renormalization in Statistical Physics Vol. 5
tions. Adv. Mater. 33, 2000713 (2021). (Cambridge University Press, 1996).
9. Huang, S. et al. Buckling of paramagnetic chains in soft gels. Soft 33. Sharma, A. et al. Strain-controlled criticality governs the nonlinear
Matter 12, 228–237 (2016). mechanics of fibre networks. Nat. Phys. 12, 584–587 (2016).
10. Xie, Q. et al. Astral hydrogels mimic tissue mechanics by aster-aster 34. Smallwood, H. M. Limiting law of the reinforcement of rubber.
interpenetration. Nat. Commun. 12, 4277 (2021). J. Appl. Phys. 15, 758–766 (1944).
11. Puljiz, M. & Menzel, A. M. Forces and torques on rigid 35. Ju, J. W. & Chen, T. M. Effective elastic moduli of two-phase com-
inclusions in an elastic environment: resulting matrix-mediated posites containing randomly dispersed spherical inhomogeneities.
interactions, displacements, and rotations. Phys. Rev. E 95, Acta Mech. 103, 123–144 (1994).
053002 (2017). 36. Torquato, S. Random Heterogeneous Materials: Microstructure and
12. Fang, Y. et al. Dynamic and programmable cellular-scale granules Macroscopic Properties (Springer, 2002).
enable tissue-like materials. Matter 2, 948–964 (2020). 37. Bi, D., Zhang, J., Chakraborty, B. & Behringer, R. P. Jamming by
13. Hull, D. & Clyne, T. W. An Introduction to Composite Materials, shear. Nature 480, 355–358 (2011).
Cambridge Solid State Science Series 2nd edn (Cambridge 38. Vinutha, H. A. & Sastry, S. Disentangling the role of structure and
University Press, 1996). friction in shear jamming. Nat. Phys. 12, 578–583 (2016).

Nature Communications | (2024)15:1691 7


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article https://doi.org/10.1038/s41467-024-45964-y

39. Kumar, N. & Luding, S. Memory of jamming—multiscale models for Author contributions
soft and granular matter. Granul. Matter 18, 58 (2016). Y.Z., Y.W. and Q.X. designed the project. Y.Z., H.H., C.Y. and C.X. con-
40. Baity-Jesi, M., Goodrich, C. P., Liu, A. J., Nagel, S. R. & Sethna, J. P. ducted the experimental measurements. Y.Z. and Q.X. analyzed the
Emergent so(3) symmetry of the frictionless shear jamming transi- experimental data. H.L., Y.Z. and Q.X. built the scaling model. Y.H. and
tion. J. Stat. Phys. 167, 735–748 (2017). R.Z. performed the simulations on the isotropic jammed states. Y.Z., H.L.
41. Han, E., James, N. M. & Jaeger, H. M. Stress controlled rheology of and Q.X. wrote the manuscript.
dense suspensions using transient flows. Phys. Rev. Lett. 123,
248002 (2019). Competing interests
42. Zhao, Y., Barés, J., Zheng, H., Socolar, J. E. S. & Behringer, R. P. The authors declare no competing interests.
Shear-jammed, fragile, and steady states in homogeneously
strained granular materials. Phys. Rev. Lett. 123, 158001 (2019). Additional information
43. Seto, R., Singh, A., Chakraborty, B., Denn, M. M. & Morris, J. F. Shear Supplementary information The online version contains
jamming and fragility in dense suspensions. Granul. Matter 21, supplementary material available at
82 (2019). https://doi.org/10.1038/s41467-024-45964-y.
44. Jin, Y. & Yoshino, H. A jamming plane of sphere packings. Proc. Natl
Acad. Sci. USA 118, e2021794118 (2021). Correspondence and requests for materials should be addressed to
45. Pan, D., Wang, Y., Yoshino, H., Zhang, J. & Jin, Y. A review on shear Yiqiu Zhao or Qin Xu.
jamming. Phys. Rep. 1038, 1–18 (2023).
46. Han, E., Wyart, M., Peters, I. R. & Jaeger, H. M. Shear fronts in shear- Peer review information Nature Communications thanks Dong Wang
thickening suspensions. Phys. Rev. Fluids 3, 073301 (2018). and the other anonymous reviewer(s) for their contribution to the peer
47. Xu, Q., Wilen, L. A., Jensen, K. E., Style, R. W. & Dufresne, E. R. review of this work. A peer review file is available.
Viscoelastic and poroelastic relaxations of soft solid surfaces. Phys.
Rev. Lett. 125, 238002 (2020). Reprints and permissions information is available at
48. van Oosten, A. S. G. et al. Emergence of tissue-like mechanics from http://www.nature.com/reprints
fibrous networks confined by close-packed cells. Nature 573,
96–101 (2019). Publisher’s note Springer Nature remains neutral with regard to jur-
49. Song, J., Deiss-Yehiely, E., Yesilata, S. & McKinley, G. H. Strain stif- isdictional claims in published maps and institutional affiliations.
fening universality in composite hydrogels and tissues. Preprint at
https://arxiv.org/abs/2307.11687. (2023). Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
Acknowledgements adaptation, distribution and reproduction in any medium or format, as
We thank Bulbul Chakraborty, Yilong Han, Hisao Hayakawa, Ryohei Seto, long as you give appropriate credit to the original author(s) and the
and Xiang Cheng for the insightful discussions. The work was supported source, provide a link to the Creative Commons licence, and indicate if
by the Early Career Scheme No. 26309620 (Q.X.), the General Research changes were made. The images or other third party material in this
Fund No. 16307422 (Q.X.) and No. 16300221 (R.Z.), and the Collaborative article are included in the article’s Creative Commons licence, unless
Research Fund No. C6004-22Y (Q.X.) and No. C6008-20E (Q.X.) from indicated otherwise in a credit line to the material. If material is not
the Hong Kong Research Grants Council (RGC). We also appreciate the included in the article’s Creative Commons licence and your intended
support of the Partnership Seed Fund from the Asian Science and use is not permitted by statutory regulation or exceeds the permitted
Technology Pioneering Institutes of Research and Education League No. use, you will need to obtain permission directly from the copyright
ASPIRE2021#1 (Q.X. and Y.W.). Yiqiu Zhao acknowledges the support holder. To view a copy of this licence, visit http://creativecommons.org/
from the RGC postdoctoral fellowship PDFS2324-6S02 (Y.Z.). Hanqing licenses/by/4.0/.
Liu is supported by the U.S. Department of Energy, Office of Science,
Nuclear Physics program, and by the Quantum Science Center. © The Author(s) 2024

Nature Communications | (2024)15:1691 8


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:

1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at

onlineservice@springernature.com

You might also like