Skip to main content
Log in

A viscoelastic bonded particle model to predict rheology and mechanical properties of hydrogel spheres

  • Original Report
  • Published:
Granular Matter Aims and scope Submit manuscript

Abstract

The use of hydrogels has exponentially increased in recent years in many fields, such as biology, medicine, pharmaceuticals, agriculture, and more. These materials are so widely used because their mechanical properties change drastically with the different chemical compositions of the constituent polymer chains, making them highly versatile for different applications. We introduce a numerical simulation tool that relies on the Discrete Element Method to reproduce and predict the behavior of hydrogel spheres. We first use a benchmark test, namely an oscillatory compression test on a single hydrogel, to calibrate the model parameters, obtaining a good agreement on the material’s rheological properties. Specifically, we show that the normal modified storage and loss moduli, E’ and E”, obtained in the simulation match the experimental data with a small relative error, around 3%, for E’ and 11% for E”. This result aligns with recent work on numerical modeling of hydrogels, introducing a novel approach with bonded particles and a viscoelastic constitutive relation that can capture a wide range of applications thanks to the higher number of elements. Moreover, we validate the model on a particle-particle compression test by comparing the simulation output with the contact force in the compression direction, again obtaining promising results.

Graphical abstract

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
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

References

  1. Chang, E.P., Holguin, D.: Electrooptical light-management material: Low-refractive-index hydrogels. J. Adhes. 83, 15–26 (2007). https://doi.org/10.1080/00218460601102803

    Article  Google Scholar 

  2. Vashuk, E.V., Vorobieva, E.V., Basalyga, I.I., Krutko, N.P.: Water-absorbing properties of hydrogels based on polymeric complexes. Mater. Res. Innov. 4, 350–352 (2001). https://doi.org/10.1007/s100190000115

    Article  Google Scholar 

  3. Li, Y., Neoh, K., Kang, E.: Poly(vinyl alcohol) hydrogel fixation on poly(ethylene terephthalate) surface for biomedical application. Polymer 45, 8779–8789 (2004). https://doi.org/10.1016/j.polymer.2004.10.077

    Article  Google Scholar 

  4. Zhai, M., Xu, Y., Zhou, B., Jing, W.: Keratin-chitosan/n-zno nanocomposite hydrogel for antimicrobial treatment of burn wound healing: Characterization and biomedical application. J. Photochem. Photobiol. B: Biol. 180, 253–258 (2018). https://doi.org/10.1016/j.jphotobiol.2018.02.018

    Article  Google Scholar 

  5. Deng, Y., et al.: Dual physically cross-linked \(\kappa \)-carrageenan-based double network hydrogels with superior self-healing performance for biomedical application. ACS Book Mater. Interfaces 10, 37544–37554 (2018). https://pubs.acs.org/doi/10.1021/acsami.8b15385

  6. Susilowati, E., Maryani, Ashadi: Green synthesis of silver-chitosan nanocomposite and their application as antibacterial material. J. Phys.: Conf. Ser. 1153, 012135 (2019). https://doi.org/10.1088/1742-6596/1153/1/012135

  7. Bhatnagar, A., Kumar, R., Singh, V.P., Pandey, D.S.: Hydrogels:a boon for increasing agricultural productivity in water-stressed environment. Curr. Sci. 111, 1773 (2016). https://doi.org/10.18520/cs/v111/i11/1773-1779

  8. Nguyen, V.N., et al.: Rheological characterization of mechanical properties of chemically crosslinked microspheres. J. Appl. Polym. Sci. 128, 3113–3121 (2012). https://doi.org/10.1002/app.38510

    Article  Google Scholar 

  9. Wang, D., et al.: The structural, vibrational, and mechanical properties of jammed packings of deformable particles in three dimensions. Soft Matter 17, 9901–9915 (2021). https://doi.org/10.1039/d1sm01228b

    Article  ADS  Google Scholar 

  10. Zeng, X., et al.: Real-time quantitative measurement of mechanical properties of spherical hydrogels during degradation by hydrodynamic loading and numerical simulation. Polym. Degrad. Stab. 202, 110055 (2022). https://doi.org/10.1016/j.polymdegradstab.2022.110055

    Article  Google Scholar 

  11. Potyondy, D., Cundall, P.: A bonded-particle model for rock. Int. J. Rock Mech. Min. Sci. 41(8), 1329–1364 (2004)

    Article  Google Scholar 

  12. Giannis, K., et al.: Stress based multi-contact model for discrete-element simulations. Granul. Matter 23 (2021). https://doi.org/10.1007/s10035-020-01060-8

  13. Ghods, N., Poorsolhjouy, P., Gonzalez, M., Radl, S.: Discrete element modeling of strongly deformed particles in dense shear flows. Powder Technol. 401, 117288 (2022). https://doi.org/10.1016/j.powtec.2022.117288

    Article  Google Scholar 

  14. Lu, W.-M., Tung, K.-L., Hung, S.-M., Shiau, J.-S., Hwang, K.-J.: Compression of deformable gel particles. Powder Technol. 116, 1–12 (2001). https://doi.org/10.1016/s0032-5910(00)00357-0

    Article  Google Scholar 

  15. Mascara, M., Mayrhofer, A., Radl, S., Kloss, C.: Implementation and validation of a bonded particle model to predict rheological properties of viscoelastic materials. Particuology 89, 198–210 (2024). https://doi.org/10.1016/j.partic.2023.11.001

    Article  Google Scholar 

  16. Feng, H., Pettinari, M., Stang, H.: Study of normal and shear material properties for viscoelastic model of asphalt mixture by discrete element method. Constr. Build. Mater. 98, 366–375 (2015). https://doi.org/10.1016/j.conbuildmat.2015.08.116

    Article  Google Scholar 

  17. Wang, Y., Alonso-Marroquin, F.: A finite deformation method for discrete modeling: particle rotation and parameter calibration. Granul. Matter 11, 331–343 (2009). https://doi.org/10.1007/s10035-009-0146-2

    Article  Google Scholar 

  18. Boots, J.N.M., Fokkink, R., van der Gucht, J., Kodger, T.E.: Development of a multi-position indentation setup: Mapping soft and patternable heterogeneously crosslinked polymer networks. Rev. Sci. Instrum. 90, 6 (2019). https://doi.org/10.1063/1.5043628

    Article  Google Scholar 

  19. Asadi, V., Ruiz-Franco, J., van der Gucht, J., Kodger, T.E.: Tuning moduli of hybrid bottlebrush elastomers by molecular architecture. Mater. Des. 234, 112326 (2023). https://doi.org/10.1016/j.matdes.2023.112326

    Article  Google Scholar 

  20. Majidi, B., Taghavi, S., Fafard, M., Ziegler, D., Alamdari, H.: Discrete element method modeling of the rheological properties of coke/pitch mixtures. Materials 9, 334 (2016). https://doi.org/10.3390/ma9050334

    Article  ADS  Google Scholar 

  21. Dondi, G., et al.: Modeling the dsr complex shear modulus of asphalt binder using 3d discrete element approach. Constr. Build. Mater. 54, 236–246 (2014). https://doi.org/10.1016/j.conbuildmat.2013.12.005

    Article  Google Scholar 

  22. Barnes, H.A.: A Handbook of Elementary Rheology. Institute of Non-Newtonian Fluid Mechanics, University of Wales (2000)

  23. Osswald, T., Rudolph, N.: Polymer Rheology. Hanser (2015)

  24. Ren, J., Sun, L.: Generalized maxwell viscoelastic contact model-based discrete element method for characterizing low-temperature properties of asphalt concrete. J. Mater. Civil Eng. 28 (2016). https://doi.org/10.1061/(asce)mt.1943-5533.0001390

  25. Feng, H., Pettinari, M., Stang, H.: 8th rilem international symposium on testing and characterization of sustainable and innovative bituminous materials, RILEM, pp. 423–433. Elsevier (2015)

  26. Cox, W.P., Merz, E.H.: Correlation of dynamic and steady flow viscosities. J. Polym. Sci. 28, 619–622 (1958). https://doi.org/10.1002/pol.1958.1202811812

    Article  ADS  Google Scholar 

  27. Bakri, T., Nabergoj, R., Tondl, A.: Multi-frequency oscillations in self-excited systems. Nonlinear Dyn. 48, 115–127 (2006). https://doi.org/10.1007/s11071-006-9077-1

    Article  MathSciNet  Google Scholar 

  28. Cuccia, N.L., Pothineni, S., Wu, B., Harper, J.M., Burton, J.C.: Pore-size dependence and slow relaxation of hydrogel friction on smooth surfaces. Proc Natl Acad Sci 117, 11247–11256 (2020). https://doi.org/10.1073/pnas.1922364117

    Article  ADS  Google Scholar 

  29. Castro, F.J., Radl, S.: A combined sph-dem approach for extremely deformed granular packings: validation and compression tests. Comp. Part. Mech. 11, 185–196 (2023). https://doi.org/10.1007/s40571-023-00616-8

    Article  Google Scholar 

  30. Brodu, N., Dijksman, J.A., Behringer, R.P.: Multiple-contact discrete-element model for simulating dense granular media. Phys. Rev. E 91 (2015). https://doi.org/10.1103/physreve.91.032201

Download references

Acknowledgements

This work was funded by the EU Horizon 2020 MSCA ITN program CALIPER with grant number 812638. The Department of Physical Chemistry and Soft Matter of Wageningen University and Research has supported this work.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Michael Mascara or Stefan Radl.

Ethics declarations

Conflicts of interest

There are no conflicts to declare.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: Validation of the bonded generalized Maxwell model

Appendix: Validation of the bonded generalized Maxwell model

The simulation is defined to resemble the single particle compression-extension experiment under dry conditions as accurately as possible, where the effect of the discretization needs to be investigated. When discretizing a larger object with DEM spheres, both the spatial configuration and the particles size play a role. In general, the higher the number of particles, the higher the computational cost of the simulation, but the better accuracy is achieved and vice-versa. Some models try to apply a so-called ”coarse-graining”, which aims to reduce the number of particles in the simulation without losing the information a larger system gives. Similarly, in this case, we want to check the effect of particle size and their configuration on the variables computed in a test simulation. We start by checking the effect of particles size by means of a simple test case. We simulate an oscillatory compression-extension motion, similar to the one used in the cyclic compression experiment, but applied to a cube this time. The cube is formed by a Face Centered Cube lattice, as shown in Fig. 15.

Fig. 15
figure 15

Geometrical representation of the FCC lattice cube used for the convergence study and validation of the model

Three different particle sizes are used to study the effect on the output. The size is computed such that the total volume of the cube is kept constant according to the number of particles on one side of the cube. The particles aligned vertically have a distance of \(\delta l = 1.001\cdot \sqrt{2}\cdot d\) between each other, accounting for a total length of the cube side of \(L = \delta l\cdot (N-1)+d=0.005\) m, with N being the number of particles on the side of the cube and d being the diameter of the particles. It is straightforward to conclude that, given the fixed length of the cube and, hence, the volume, the relative diameter of the DEM particles can be calculated by changing the number of particles.

If one blindly computed the system’s stress response and compared it to the analytical solution, one would obtain a great difference in magnitude. This is because the lattice configuration used here, together with the direction of the deformation, would produce both normal and shear force components that are not aligned with the deformation, which means that the shear component of the force must be reduced using the ratio introduced in Section 2.2 between the normal and shear parameters.

To determine the optimal parameter \(\alpha \), the classical approach used in determining DEM simulation parameters is used, which consists of simulating with a first guess parameter, evaluating the macro-properties to compare with analytical or experimental solutions, computing a residual and repeating the loop until the residual is minimized. The latter is a well-known and established method to calibrate parameters in a model. However, it lacks generality as it is usually optimized for specific applications, and the procedure must be repeated each time something in the original setup is changed. The cost function used in this case is similar to the one used in (9), with the difference that now we do not have a frequency sweep but only one frequency, and instead of comparing the modified storage and loss modulus, we are comparing the amplitude and the phase shift of the stress response, since those are the quantities that we want to minimize the difference of, giving

$$\begin{aligned} f_{cost} = \left( \dfrac{\sigma _{0,an}-\sigma _{0,sim}}{\sigma _{0,an}}\right) ^2 + \left( \dfrac{\varphi _{an}-\varphi _{sim}}{\varphi _{an}}\right) ^2, \end{aligned}$$
(A.1)
Fig. 16
figure 16

Zoomed particular of the stress response of the FCC lattice cube under cyclic compression, when using the parameters of Table 1, with relative errors for phase and amplitude showing on the inset graph. All errors are measured against the analytical solution given by (A.2), resulting in values well below 3%, making the model valid for predicting the stress response in oscillatory deformation, regardless of particle size. To note the relative errors for both phase shift and amplitude in the inset graph. Stress is normalized w.r.t. the amplitude of the analytical stress \(\sigma _{0,an}\)

Fig. 17
figure 17

Stress response of the BCC lattice for the same parameters of the FCC case. Here, the value of \(\alpha \) found for the FCC lattice can not be used, showing how this parameter needs to be optimized for each specific system when changing either the lattice configuration and/or the particle size

where the analytical solution, in this case, is known as [23]

$$\begin{aligned} \sigma (t) = \sum _j^4\dfrac{k_j\gamma _0\omega \tau _j}{1+(\omega \tau )^2}(\omega \tau _j\sin {\omega t}+\cos {\omega t}). \end{aligned}$$
(A.2)

The stress in the DEM simulation is computed as the average force on the top layer of particles of the cube, divided by the face area

$$\begin{aligned} \sigma _{sim}(t) = \dfrac{1}{L^2}\sum _i^{N_{top}}f_i(t). \end{aligned}$$
(A.3)

with \(N_{top}\) being the number of particles on the top layer of the cube. Once the stresses are known, the analytical and the simulated phase lags are computed by fitting the relative stresses with a function of the form \(y=A\sin (\omega t+\varphi )\).

To perform the calibration, the smallest particle size is used, returning a value of \(\alpha = 9.1\). This parameter is then used with the three different particles sizes to produce the plot of Fig. 16, which is a zoomed-in version of the total stress response, to observe better the effect of the particles’ size on the amplitude and the phase of the stress signal. In detail, no effect is observed on the phase shift error \(\chi _{\varphi }\), as shown in the inset graph of Fig. 16, with a relative error that is almost constant and below \(2\%\). A more significant effect is observed for the amplitude error \(\chi _{\sigma }\). In conclusion, smaller particles give more accurate results, without the need to change the model parameters.

To check the effect of the particles spatial configuration, a body-centered cube lattice is now used to compute the stress response of the same volume. The same model parameters are used as in the FCC configuration, and the same test is performed together with the same post processing to compute the stress, resulting in the plot of Fig. 17. In this case, the particle size used is computed according to \(d=\dfrac{L}{(N-1)*1.001+1}\), giving \(d=5\cdot 10^{-4}\) m for N=10. As it can be easily observed, a large difference in the amplitude is observed between analytical and simulated stress. This suggests that the scaling of the shear parameters w.r.t. the normal ones needs to be adjusted, resulting in another calibration process to obtain the optimal \(\alpha \) parameter. In conclusion, when changing the spatial configuration, one needs to adjust the \(\alpha \) parameter accordingly but, once that is found, a change in particle size would improve the accuracy of the model without the need to re-calibrate its parameters.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mascara, M., Shakya, C., Radl, S. et al. A viscoelastic bonded particle model to predict rheology and mechanical properties of hydrogel spheres. Granular Matter 26, 64 (2024). https://doi.org/10.1007/s10035-024-01429-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10035-024-01429-z

Keywords

Navigation