Skip to main content
Log in

In search of physical meaning: defining transient parameters for nonlinear viscoelasticity

  • Original Contribution
  • Published:
Rheologica Acta Aims and scope Submit manuscript

Abstract

A complete set of model-independent viscoelastic functions for understanding responses to transient nonlinear rheological tests is presented, using large-amplitude oscillatory shear strain as a model nonlinear protocol. The derivation makes no assumptions about symmetries, and is therefore applicable to the responses to any input, allowing researchers to unambiguously define time-dependent moduli, viscosities, compliances, fluidities, and normal stress coefficients. A legend for interpreting the dynamic trajectories in modulus space is provided, along with explicit definitions of the rates at which the moduli change. These provide a quantitative mechanism to identify when, and by how much, a material response stiffens, softens, thickens, or thins while being deformed. In addition to providing analytical expressions for the moduli, the derivation requires the definition of a conceptually new term. This means there exist three, not two, time-dependent nonlinear viscoelastic functions by which any response can be fully described. The third function accounts for nonlinear properties such as yield stresses and the shifting of the strain equilibrium. This complete analysis scheme is unique in making a distinction between the strains in the lab and material frames. The quantitative sequence of physical process analysis, which is fully developed in this work, allows for comprehensive physical interpretations of responses to transient deformations of any kind to be made, including the steady alternance responses to large-amplitude oscillatory shear (LAOS), time-dependent oscillatory shear startup responses, and thixotropic and anti-thixotropic responses.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  • Bacabac RG, Smit TH, Mullender MG, Dijcks SJ, Van Loon JJWA, Klein-Nulend J (2004) Nitric oxide production by bone cells is fluid shear stress rate dependent. Biochem Biophys Res Commun 315:823–829

    Article  Google Scholar 

  • Barnes HA, Walters K (1985) The yield stress myth? Rheol Acta 24:323–326

    Article  Google Scholar 

  • Berret JF, Roux D, Porte G (1994). Isotropic-to-nematic transition in wormlike micelles under shear. Journal de Physique II, EDP Sciences, 4(8):1261–1279

  • Calabrese MA, Wagner NJ, Rogers SA (2016) An optimized protocol for the analysis of time-resolved elastic scattering experiments. Soft Matter 12:2301

    Article  Google Scholar 

  • Cho KS, Ahn KH, Lee SJ (2005) A geometrical interpretation of large-amplitude oscillatory shear response. J Rheol 49(3):747–758

    Article  Google Scholar 

  • Cox WP, Merz EH (1958) Correlation of dynamic and steady flow viscosities. J Polym Sci 28:619–622

    Article  Google Scholar 

  • de Souza Mendes PR (2009) Modeling the thixotropic behavior of structured fluids. J Non-Newtonian Fluid Mech 164:66–75

    Article  Google Scholar 

  • de Souza Mendes PR (2011) Thixotropic elasto-viscoplastic model for structured fluid. Soft Matter 7:2471–2483

    Article  Google Scholar 

  • de Souza Mendes PR, Thompson RL (2012) A critical overview of elasto-viscoplastic thixotropic modeling. J Non-Newtonian Fluid Mech 187-188:8–15

    Article  Google Scholar 

  • de Souza Mendes PR, Thompson RL (2013) A unified approach to model elastic-viscoplastic thixotropic yield-stress materials and apparent yield-stress fluids. Rheol Acta 52:673–694

    Article  Google Scholar 

  • de Souza Mendes PR, Thompson RL, Alicke AA, Leite RT (2014) The quasilinear large-amplitude viscoelastic regime and its significance in the rheological characterization of soft matter. J Rheol 58:537–561

    Article  Google Scholar 

  • Dealy JM, Morris J, Morrison F, Vlassopoulos D (2013) Official symbols and nomenclature of the Society of Rheology. J Rheol 57:1047–1055

    Article  Google Scholar 

  • Dodge JS, Krieger IM (1971) Oscillatory shear of nonlinear fluids. I Preliminary investigation Trans Soc Rheol 15(4):589–601

    Google Scholar 

  • Evans, A. G. (1974) Slow crack growth in brittle materials under dynamic loading conditions. International Journal of Fracture, 10:251–259

  • Ewoldt RH (2013) Defining nonlinear rheological material functions for oscillatory shear. J Rheol 57:177

    Article  Google Scholar 

  • Ewoldt HE, Bharadwaj NA (2015) Constitutive model fingerprints in medium-amplitude oscillatory shear. J Rheol Acta 59:557

  • Ewoldt RH, Hosoi AE, McKinley GH (2008) New measures for characterizing nonlinear viscoelasticity in large-amplitude oscillatory shear. J Rheol 52(6):1427–1458

    Article  Google Scholar 

  • Ewoldt RH, Winter P, Maxey J, McKinley GH (2010) Large amplitude oscillatory shear of pseudoplastic and elastoviscoplastic materials. Rheol Acta 49:191–212

    Article  Google Scholar 

  • Ferry JD (1980) Viscoelastic properties of polymers, 3rd edn. Wiley, New York

    Google Scholar 

  • Frenet F (1852) Sur les courbes à double courbure. Journal de mathématiques pures et appliquées 1re série, tome 17:437–447

    Google Scholar 

  • Giacomin AJ, Bird RB, Johnson LM, Mix AW (2011) Large amplitude oscillatory shear flow from the corotational Maxwell model. J Non-Newtonian Fluid Mech 166:1081–1099

    Article  Google Scholar 

  • Gurnon AK, Lopez-Barron CR, Eberle APR, Porcar L, Wagner NJ (2014) Spatiotemporal stress and structure evolution in dynamically sheared polymer-like micellar solutions. Soft Matter 10:2889

    Article  Google Scholar 

  • Harris J, Bogie K (1967) The experimental analysis of non-linear waves in mechanical systems. Rheol Acta 6(1):3–5

    Article  Google Scholar 

  • Hyun K, Wilhelm M (2009) Establishing a new mechanical nonlinear Q coefficient from FT-rheology: first investigation of entangled linear and comb polymer model systems. Macromolecules 42:411–422

    Article  Google Scholar 

  • Hyun K, Wilhelm M, Klein CO, Cho KS, Nam JG, Ahn KH, Lee SJ, Ewoldt RH, McKinley GH (2011) A review of nonlinear oscillatory shear tests: analysis and applications of large amplitude oscillatory shear (LAOS). Prog Polym Sci 36:1697–1753

    Article  Google Scholar 

  • Kim J, Merger D, Wilhelm M, Helgeson ME (2014) Microstructure and nonlinear signatures of yielding in a heterogeneous colloidal gel under large amplitude oscillatory shear. J Rheol 58:1359–1390

    Article  Google Scholar 

  • Klein CO, Spiess HW, Calin A, Balan C, Wilhelm M (2007) Separation of the nonlinear response into a superposition of linear, strain hardening, strain softening, and wall slip response. Macromolecules 40:4250–4259

    Article  Google Scholar 

  • Koumakis N, Brady JF, Petekidis G (2013) Complex oscillatory yielding of model hard-sphere glasses. PRL 110:178301

    Article  Google Scholar 

  • Läuger J, Stettin H (2010) Differences between stress and strain control in the non-linear behavior of complex fluids. Rheol Acta 49(9):909–930

    Article  Google Scholar 

  • Lettinga MP, Holmqvist P, Ballesta P, Rogers S, Kleshchanok D, Struth B (2012) Nonlinear behavior of nematic platelet dispersions in shear flow. PRL 109:246001

    Article  Google Scholar 

  • Lonetti B, Kohlbrecher J, Willner L, Dhont JKG, Lettinga MP (2008) Dynamic response of block copolymer wormlike micelles to shear flow. J Phys Condens Matter 20:404207

    Article  Google Scholar 

  • Lopez-Barron CR, Porcar L, Eberle APR, Wagner NJ (2012) Dynamics of melting and recrystallization in a polymeric micellar crystal subjected to large amplitude oscillatory shear flow. PRL 108:258301

    Article  Google Scholar 

  • Lyklema J, van Olphen H (1979) Terminology and symbols in colloid and surface chemistry part 1.13. Definitions, terminology and symbols for rheological properties. Pure & Appl Chem 51:1213–1218

    Google Scholar 

  • Mewis J, Wagner NJ (2012) Colloidal suspension rheology. Colloidal Suspension Rheology Cambridge University Press, Cambridge, England, 2011

  • Onogi S, Masuda T, Matsumoto T (1970) Non-linear behavior of viscoelastic materials. I Disperse systems of polystyrene solution and carbon black J Rheol 14(2):275–294

    Google Scholar 

  • Park JD, Ahn KH, Lee SJ (2015) Structural change and dynamics of colloidal gels under oscillatory shear flow. Soft Matter 11:9262

    Article  Google Scholar 

  • Pearson DS, Rochefort WE (1982) Behavior of concentrated polystyrene solutions in large-amplitude oscillating shear fields. J. Polym. Sci., Polym. Phys Ed 20(1):83–98

    Article  Google Scholar 

  • Phillippoff W (1966) Vibrational measurements with large amplitudes. Trans Soc Rheol 10:317–334

    Article  Google Scholar 

  • Pipkin AC (1972) Lectures on viscoelasticity theory. Springer, New York

    Book  Google Scholar 

  • Poulos AS, Stellbrink J, Petekidis G (2013) Flow of concentrated solutions of starlike micelles under large-amplitude oscillatory shear. Rheol Acta 52:785–800

    Article  Google Scholar 

  • Pressley, A. (ed) (2010) Elementary differential geometry, Springer London

  • Rehage H, Hoffmann H (1988) Rheological properties of viscoelastic surfactant systems. J Phys Chem 92(16):4712–4719

    Article  Google Scholar 

  • Rogers SA (2012) A sequence of physical processes determined and quantified in LAOS: an instantaneous local 2D/3D approach. J Rheol 56(5):1129–1151

    Article  Google Scholar 

  • Rogers SA, Lettinga MP (2012) A sequence of physical processes determined and quantified in large amplitude oscillatory shear (LAOS): application to theoretical nonlinear models. J Rheol 56(1):1–25

    Article  Google Scholar 

  • Rogers SA, Erwin BM, Vlassopoulos D, Cloitre M (2011) A sequence of physical processes determined and quantified in LAOS: application to a yield stress fluid. J Rheol 55(2):435–458

    Article  Google Scholar 

  • Rogers S, Kohlbrecher J, Lettinga MP (2012) The molecular origin of stress generation in worm-like micelles, using a rheo-SANS LAOS approach. Soft Matter 8:7831

    Article  Google Scholar 

  • Saengow CA, Giacomin J, Kolitawong C (2015) Exact analytical solution for large-amplitude oscillatory shear flow. Macromol Theory Simul 24:352–392

    Article  Google Scholar 

  • Serret J-A (1851) Sur quelques formules relatives à la théorie des courbes à double courbure. Journal de mathématiques pures et appliquées 1re série, tome 16:193–207

    Google Scholar 

  • Sharma V, McKinley GH (2012) An intriguing empirical rule for computing the first normal stress difference from steady shear viscosity data for concentrated polymer solutions and melts. Rheol Acta 51:487–495

    Article  Google Scholar 

  • Tee TT, Dealy JM (1975) Nonlinear viscoelasticity of polymer melts. Trans Soc Rheol 19(4):595–615

    Article  Google Scholar 

  • Thompson RL, Alicke AA, de Souza Mendes PR (2015) Model-based material functions for SAOS and LAOS analyses. J Non-Newtonian Fluid Mech 215:19–30

    Article  Google Scholar 

  • van der Vaart K, Rahmani Y, Zargar R, Hu Z, Bonn D, Schall P (2013) Rheology of concentrated soft and hard-sphere suspensions. J Rheol 57:1195

    Article  Google Scholar 

  • Wang Y-C, Gunasekaran S, Giacomin AJ (2001) The lodge rubberlike liquid behavior for cheese in large amplitude oscillatory shear. Appl Rheol 11(6):312–319

    Google Scholar 

  • Yoshimura AS, Prud’homme RK (1987) Response of an elastic Bingham fluid to oscillatory shear. Rheol Acta 26:428–436

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by start-up funds from the Department of Chemical and Biomolecular Engineering at the University of Illinois at Urbana-Champaign. I am grateful for helpful and insightful comments from Florian Nettesheim, Roney Thompson, and Paulo de Souza Mendes.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Simon A. Rogers.

Appendices

Appendix 1

Dimensions, units, stress-controlled experiments, and time-dependent viscosities, compliances, and fluidities

The derivations presented in the main body of the text have been carried out with strain-controlled experiments in mind, where one measures the stress response to oscillating strains and strain-rates and seeks to define moduli. Nonlinear viscoelastic functions in the form of moduli or viscosities from strain-controlled experiments, or compliances or fluidities from responses to stress-controlled tests, in addition to normal stresses from any tests can be derived via a generic route presented in the “A generic definition of time-dependent nonlinear viscoelastic functions” section. In this appendix, the different native spaces in which responses reside are discussed. The generic \( \left[\begin{array}{ccc}\hfill x\hfill & \hfill y\hfill & \hfill z\hfill \end{array}\right] \) notation will be employed to refer to the particular spaces. The ability to provide a single derivation for all nonlinear viscoelastic parameters is brought about by making the analogy between the linear-regime description of a rheological response and Eq. 7, the equation of the general form of a plane, which is repeated here:

$$ ax+ by+ cz+ d=0. $$
(72)

Equation 7 is the general form of the equation of a plane in \( \left[\begin{array}{ccc}\hfill x\hfill & \hfill y\hfill & \hfill z\hfill \end{array}\right] \)-space that has a normal \( \boldsymbol{n}=\left[\begin{array}{ccc}\hfill a\hfill & \hfill b\hfill & \hfill c\hfill \end{array}\right] \) and a vertical displacement of d. A trajectory in this plane therefore has a binormal vector given by \( \boldsymbol{B}=\left[\begin{array}{ccc}\hfill a\hfill & \hfill b\hfill & \hfill c\hfill \end{array}\right] \).

The displacement term, d, reduces to zero in the linear regime and can therefore be neglected in our discussion of spaces without loss of generality. In a generalization of the procedure that leads to Eqs. 5 and 6 from Eqs. 3 and 4, Eq. 72 is interpreted as describing a single response dimension, z(t), as a sum of two terms each consisting of a single viscoelastic shear material function multiplied by one of the excitation dimensions, x(t) or y(t). The formalism need not be restricted to shear conditions, though, and corresponding tension, compression, and torsional parameters can be derived without hindrance.

Application of a sinusoidal strain (Eq. 1) can equally be thought of as application of a cosinusoidal strain rate (Eq. 2). The material functions that describe the stress response to a strain rate are viscosities, meaning that the stress response can be generically described by

$$ \sigma \left(\omega, t\right)={\eta}^{\mathit{\prime}}\left(\omega \right)\dot{\gamma}(t)+{\eta}^{\mathit{{\prime\prime}}}\left(\omega \right)\omega \gamma (t) $$
(73)

Comparison with Eq. 72 suggests that \( \left[\begin{array}{ccc}\hfill \dot{\gamma}(t)\hfill & \hfill \omega \gamma (t)\hfill & \hfill \sigma (t)\hfill \end{array}\right] \) is the natural space in which to define time-dependent dynamic viscosities. Further, comparing the coefficients of the strain and rate directions in Eqs. 3 and 73 leads to the familiar scaling of dynamic moduli being equal to the product of the angular frequency and the dynamic viscosities.

Correspondingly, the generic linear-regime descriptions of the measured strain and strain-rate responses to an applied oscillatory stress define the dynamic compliances and fluidities:

$$ \gamma \left(\omega, t\right)={J}^{\mathit{\prime}}\left(\omega \right)\sigma (t)-{J}^{\mathit{{\prime\prime}}}\left(\omega \right)\frac{1}{\omega}\dot{\sigma}\left(\boldsymbol{t}\right) $$
(74)
$$ \dot{\gamma}\left(\omega, t\right)={\phi}^{\prime}\left(\omega \right)\sigma (t)+{\phi}^{{\prime\prime}}\left(\omega \right)\frac{1}{\omega}\dot{\sigma}(t) $$
(75)

Comparisons of Eqs. 74 and 75 with Eq. 72 suggest that the spaces in which dynamic compliances and fluidities are naturally defined are therefore \( \left[\begin{array}{ccc}\hfill \sigma (t)\hfill & \hfill -\dot{\sigma}(t)/\omega \hfill & \hfill \gamma (t)\hfill \end{array}\right] \) and \( \left[\begin{array}{ccc}\hfill \sigma (t)\hfill & \hfill \dot{\sigma}(t)/\omega \hfill & \hfill \dot{\gamma}(t)\hfill \end{array}\right] \), respectively. The x dimensions of these spaces are the same, while the y dimensions are different by a factor of −1.

Under the general construction of Eq. 13, the dimensions of the space must all have the same units, which are those of the perceived applied perturbation (in the formalism suggested here, the perceived applied perturbation is always set as being the x-dimension). In the case of moduli (Eq. 3), the perceived applied perturbation is in the strain, and therefore, the dimensions of the native space are unit-less. If the strain-rate is thought of as being the perturbing quantity (Eq. 72), then the dimensions of the native space take units of inverse time. Interestingly, in the cases of compliances and fluidities (Eqs. 73 and 74), both native spaces take units of stress, reflecting the identical perturbing parameter.

In the soft matter community, a perturbation of stress-rate is rarely considered and so is hardly used by the vast majority of soft matter scientists. The official Society of Rheology list of symbols and nomenclature (Dealy et al. 2013), for instance, does not contain an entry for shear stress-rate. Stress-rate, however, is not a completely unused parameter. Evans (1974) linked the rate at which stress was applied to the growth of cracks in brittle materials, and Bacabac et al. (2004) linked the stress rate of interstitial fluid through the lacuno-canalicular network to the adaptive response of bone cells. Viscoelastic material functions infrequently used by rheologists that are defined in terms of the stress-rate could be incorporated into the current framework in an identical manner by following the procedure outlined above.

Viewing responses to strain- and stress-controlled stimuli as existing within different spaces provides an intuitive explanation as to why nonlinear compliances and moduli are not simply inverses of each other, or clearly related. In the linear regime, however, the response of a given material or model is independent of whether the excitation is strain- or stress-controlled. The well-known relations between linear-regime shear moduli and compliances can therefore be deduced in the familiar way (Eqs. 27, 28, 29 and 30 of Chapter 1 of Ferry 1980).

Appendix 2

The Frenet–Serret apparatus

The language used throughout this work to describe the space curves that result from oscillatory excitations is that of the Frenet–Serret apparatus. The Frenet–Serret frame consists of the mutually orthogonal tangent, principal normal, and binormal vectors. Here, definitions of those vectors are presented, along with the scalars, referred to as the curvature and torsion of a response, that define how the frame reorients along a curve. The collection of the three vectors defining the frame and the two scalars defining how the frame reorients comprises the Frenet–Serret apparatus. Only issues salient to this work will be presented in this appendix. A full treatment of the geometrical concepts applied to space curves can be found elsewhere (Chapter 2 of Pressley 2010).

The space curve of a response within the three-dimensional Euclidean space ℝ3 is defined as

$$ \boldsymbol{A}=\left[\begin{array}{ccc}\hfill {A}_x\hfill & \hfill {A}_y\hfill & \hfill {A}_z\hfill \end{array}\right]=\left[\begin{array}{ccc}\hfill x\hfill & \hfill y\hfill & \hfill z\hfill \end{array}\right], $$
(76)

where x, y, and z refer to the specific dimensions of the space as discussed in Appendix 1. The time dependence of all terms is implied.

Geometrically, it is more natural to define derivatives of space curves with respect to arc length. However, because the rheological data collected from oscillatory experiments are recorded discretely at evenly spaced time intervals, the natural derivatives of interest in the current case are those with respect to time. The definitions contained in this appendix will use the usual dot notation to represent differentiation with respect to time, as opposed to the dash notation which would indicate differentiation with respect to arc length.

From the perspective of an observer traveling along the space curve (or one who imagines they are), the tangent vector points in the direction of instantaneous travel and is defined as

$$ \boldsymbol{T}=\frac{\dot{\boldsymbol{A}}}{\left\Vert \dot{\boldsymbol{A}}\right\Vert } $$
(77)

The normalized temporal derivative of the tangent vector is the principal normal vector:

$$ \boldsymbol{N}=\frac{\dot{\boldsymbol{T}}}{\left\Vert \dot{\boldsymbol{T}}\right\Vert }=\frac{\dot{\boldsymbol{A}}\times \left(\ddot{\boldsymbol{A}}\times \dot{\boldsymbol{A}}\right)}{\left\Vert \dot{\boldsymbol{A}}\right\Vert \left\Vert \ddot{\boldsymbol{A}}\times \dot{\boldsymbol{A}}\right\Vert } $$
(78)

The tangent and principal normal vectors span the osculating plane, the plane that has a second-order contact with the space curve at the point t. An n-th order contact is defined by the two objects having the same (n − 1)th derivative. Second-order contact between the osculating plane and the trajectory is therefore the case where the first derivatives of each are the same.

The orientation of the osculating plane is defined by the binormal vector, which is therefore orthogonal to both the tangent and principal normal vectors:

$$ \boldsymbol{B}=\boldsymbol{T}\times \boldsymbol{N}=\frac{\dot{\boldsymbol{A}}\times \ddot{\boldsymbol{A}}}{\left\Vert \dot{\boldsymbol{A}}\times \ddot{\boldsymbol{A}}\right\Vert } $$
(79)

The Frenet–Serret frame, made up of the set of T, N, and Bvectors, is therefore a right-handed orthonormal basis that spans ℝ3:

$$ \boldsymbol{B}=\boldsymbol{T}\times \boldsymbol{N};\boldsymbol{T}=\boldsymbol{N}\times \boldsymbol{B};\boldsymbol{N}=\boldsymbol{B}\times \boldsymbol{T} $$
(80)

A visual representation of a generic linear viscoelastic response is shown in Fig. 9 along with the Frenet–Serret frame at two instances separated by a quarter of a period.

Fig. 9
figure 9

A generic linear viscoelastic response is shown as a black solid line. The Frenet–Serret frame, which is defined by the tangent (red), principal normal (green), and binormal (blue) vectors, is displayed at two instances a quarter of a period apart. While the tangent and principal normal vectors change direction, the binormal vector remains oriented in the same direction throughout the oscillatory period (color figure online)

There are two scalars that define how the Frenet–Serret frame reorients along an arbitrary space curve. The two scalars are the curvature, κ, and the torsion, τ. Generally speaking, the curvature measures the extent to which a curve is not a straight line, and the torsion measures the extent to which the curve does not sit within a plane.

The curvature is defined as the projection of the derivative of the tangent vector onto the principal normal vector, and can therefore be thought of as the rate at which the tangent vector rotates about the binormal vector:

$$ \dot{\boldsymbol{T}}=\left\Vert \dot{\boldsymbol{A}}\right\Vert \kappa N $$
(81)

The term \( \left|\right|\dot{\boldsymbol{A}}\left|\right| \)results from the desire to work with the temporal derivatives, rather than the arc-length derivatives which naturally account for the changes in the magnitude of A. In the case of oscillatory rheology, the mutually orthogonal lab-frame oscillatory inputs, x and y, ensure that the response sits on the surface of a cylinder, which in turn means the curvature is always non-zero.

The torsion is defined as the negative projection of the derivative of the binormal vector onto the principal normal vector, and can therefore be thought of as the rate at which the binormal vector rotates about the tangent vector:

$$ \tau =-\left\Vert \dot{\boldsymbol{A}}\right\Vert \boldsymbol{N}\cdot \dot{\boldsymbol{B}}=\frac{\left(\dot{\boldsymbol{A}}\times \ddot{\boldsymbol{A}}\right)\cdot \overset{\dddot{}}{\boldsymbol{A}}}{\left\Vert \dot{\boldsymbol{A}}\times \ddot{\boldsymbol{A}}\right\Vert {}^2} $$
(82)

The geometric interpretations of the curvature and torsion are emphasized in Fig. 1. The tangent vector points in the direction of travel around the curve, and the circular arrows indicating curvature and torsion point in the directions of positive values.

To calculate the derivative of the principal normal vector, combine Eqs. 80, 81, and 82:

$$ \dot{\boldsymbol{N}}=\dot{\boldsymbol{B}}\times \boldsymbol{T}+\boldsymbol{B}\times \dot{\boldsymbol{T}}=-\left\Vert \dot{\boldsymbol{A}}\right\Vert \tau N\times \boldsymbol{T}+\boldsymbol{B}\times \left\Vert \dot{\boldsymbol{A}}\right\Vert \kappa N=\left\Vert \dot{\boldsymbol{A}}\right\Vert \left(-\kappa T+\tau B\right) $$
(83)

Equations 81, 82, and 83 are compactly summarized by Eq. 21:

$$ \left[\begin{array}{c}\hfill \dot{\boldsymbol{T}}\hfill \\ {}\hfill \dot{\boldsymbol{N}}\hfill \\ {}\hfill \dot{\boldsymbol{B}}\hfill \end{array}\right]=\left\Vert \dot{\boldsymbol{A}}\right\Vert \left[\begin{array}{ccc}\hfill 0\hfill & \hfill \boldsymbol{\kappa} \hfill & \hfill 0\hfill \\ {}\hfill -\boldsymbol{\kappa} \hfill & \hfill 0\hfill & \hfill \boldsymbol{\tau} \hfill \\ {}\hfill 0\hfill & \hfill -\boldsymbol{\tau} \hfill & \hfill 0\hfill \end{array}\right]\left[\begin{array}{c}\hfill \boldsymbol{T}\hfill \\ {}\hfill \boldsymbol{N}\hfill \\ {}\hfill \boldsymbol{B}\hfill \end{array}\right] $$
(84)

Appendix 3

An alternative illustration of the need for the displacement term

The definitions presented in Eqs. 18, 19, 22 and 23 made no assumptions regarding the form of the excitation. If this condition is relaxed and a calculation is made where sinusoidal excitations are assumed, an interesting result is obtained that confirms the need for the displacement term. An alternative way to proceed from Eq. 16 that makes use of the symmetries of sinusoidal excitations and their derivatives is to note that the binormal vector is defined by the vector cross-product of the first and second temporal derivatives of the measured trajectory, \( \dot{\boldsymbol{A}} \) and \( \ddot{\boldsymbol{A}} \) (see Appendix 2), and can therefore be written in terms of the temporal derivatives of A γ , \( {A}_{\dot{\gamma}/\omega} \), and A σ :

$$ \boldsymbol{B}\propto \dot{\boldsymbol{A}}\times \ddot{\boldsymbol{A}}=\left[\begin{array}{ccc}\hfill {\dot{\boldsymbol{A}}}_{\dot{\gamma}/\omega}\overset{\cdot \cdot }{A_{\sigma}}-{\ddot{A}}_{\dot{\gamma}/\omega}\dot{A_{\sigma}}\hfill & \hfill -\left(\dot{A_{\gamma}}\overset{\cdot \cdot }{A_{\sigma}}-\overset{\cdot \cdot }{A_{\gamma}}\dot{A_{\sigma}}\right)\hfill & \hfill \dot{A_{\gamma}}{\ddot{A}}_{\dot{\gamma}/\boldsymbol{\omega}}-\overset{\cdot \cdot }{A_{\gamma}}{\dot{A}}_{\dot{\gamma}/\boldsymbol{\omega}}\hfill \end{array}\right] $$
(85)

In the derivatives of the sinusoidal and cosinusoidal inputs, Eqs. 1 and 2 can be simplified as follows:

$$ \dot{A_{\gamma}}={\omega A}_{\dot{\gamma}/\omega};\overset{\cdot \cdot }{A_{\gamma}}=-{\omega}^2{\boldsymbol{A}}_{\gamma};{\ \dot{A}}_{\dot{\gamma}/\omega}=-{\omega A}_{\gamma};{\ \ddot{A}}_{\dot{\gamma}/\omega}=-{\omega}^2{A}_{\dot{\gamma}/\boldsymbol{\omega}} $$
(86)

The relations Eqs. 84 and 85 can therefore be used to simplify the expression of the displacement term given in Eq. 11 that involves only the A σ component and its second derivative:

$$ \frac{B_{\gamma}}{B_{\sigma}}{A}_{\gamma}+\frac{B_{\dot{\gamma}/\omega}}{B_{\sigma}}{A}_{\dot{\gamma}/\omega}+{A}_{\sigma}=\frac{1}{\omega^2}\overset{\cdot \cdot }{A_{\sigma}}+{A}_{\sigma} $$
(87)

The remarkably simple form of Eq. 86, coupled with Eq. 13, allows the generic description of shear stress responses to oscillatory strains given in Eq. 16 to be rewritten as

$$ \sigma (t)={G_t}^{\boldsymbol{\hbox{'}}}(t)\gamma \left(\boldsymbol{t}\right)+{G_t}^{\mathit{\hbox{'}}\boldsymbol{\hbox{'}}}(t)\dot{\gamma}(t)/\omega +\frac{1}{\omega^2}\ddot{\sigma}(t)+\sigma (t) $$
(88)

Canceling the σ(t) terms on either side of the equality leads to

$$ {G_t}^{\mathit{\hbox{'}}}(t)\gamma (t)+{G_t}^{\mathit{\hbox{'}}\mathit{\hbox{'}}}(t)\dot{\gamma}(t)/\omega =-\frac{1}{\omega^2}\ddot{\sigma}(t) $$
(89)

The second derivative of the stress in Eq. 89 eliminates any offsets, making this a statement that the viscoelastic parameters can only account for changes in the measured response relative to the inputs. Yield stresses and strains can therefore not be described by viscoelastic parameters such as moduli, viscosities, compliances, or fluidities defined in this way. It is noted here that Eq. 87 is derived from the general form of Eq. 16, which can be used to describe any response. The only assumptions that go into the derivation of Eq. 87 are listed in Eq. 84, which is valid for any oscillatory deformation. Equation 87 is therefore also applicable in a general sense. It is also possible to rephrase Eq. 87 in the form of Eq. 7, that is, in the general form of the equation of a plane. The nonlinear response moduli are directly interpretable in terms of the components of the binormal vector of the response in the same manner as before, but with the restriction of being in \( \left[\begin{array}{ccc}\hfill \gamma (t)\hfill & \hfill \dot{\gamma}(t)/\omega \hfill & \hfill \ddot{\sigma}(t)/{\omega}^2\hfill \end{array}\right] \)-space, and that the binormal of the trajectory, or the normal of the osculating plane, is equal to \( \left[\begin{array}{ccc}\hfill {G_t}^{\hbox{'}}(t)\hfill & \hfill {G_t}^{\hbox{'}\hbox{'}}(t)\hfill & \hfill 1\hfill \end{array}\right] \). It is also clear from Eq. 89 why there is a need to include the information about the displacement of the osculating plane in the full description of the response trajectory as given by Eq. 16—the moduli, which define the orientation of the plane, can only be used to reconstruct the second derivative of the stress, and not the whole stress response.

The linear-regime representation of the shear stress response to a sinusoidally varying strain is given by

$$ \sigma (t)={G}^{\mathit{\prime}}\left(\omega \right)\gamma (t)+{G}^{\mathit{\hbox{'}\hbox{'}}}\left(\omega \right)\dot{\gamma}(t)/\omega $$
(90)

The second derivative of this generic expression is (applying the results of Eq. 85)

$$ \ddot{\sigma}(t)=-{\omega}^2\left[{G}^{\mathit{\prime}}\left(\omega \right)\gamma (t)+{G}^{\mathit{\hbox{'}\hbox{'}}}\left(\omega \right)\dot{\gamma}(t)/\omega \right] $$
(91)

By comparing Eqs. 88 and 90, an identical conclusion to that previously drawn regarding the equality of the time-dependent response parameters and linear-regime material parameters is reached:

$$ \begin{array}{ccc}\hfill {G_t}^{\mathit{\hbox{'}}}(t)={G}^{\mathit{\hbox{'}}}\left(\omega \right)\hfill & \hfill \mathrm{and}\hfill & \hfill {G_t}^{\mathit{\hbox{'}}\mathit{\hbox{'}}}(t)={G}^{\mathit{\hbox{'}\hbox{'}}}\left(\omega \right)\hfill \end{array} $$
(92)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rogers, S.A. In search of physical meaning: defining transient parameters for nonlinear viscoelasticity. Rheol Acta 56, 501–525 (2017). https://doi.org/10.1007/s00397-017-1008-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00397-017-1008-1

Keywords

Navigation