André Pincotac,
Jeffrey Chin
ab,
Ryan Murphy
d,
F. John Burpo
ab,
Caspar Yi
ab,
Edward Chen
ab,
H. Daniel Bahaghighata,
Benjamin Thompsone,
Simuck F. Yuk
ab,
Gareth H. McKinley
c,
Enoch A. Nagelli
*ab and
Matthew Armstrong
*ae
aDepartment of Chemistry and Life Science, United States Military Academy, West Point, NY 10996, USA. E-mail: enoch.nagelli@westpoint.edu
bPhotonics Research Center, United States Military Academy, West Point, NY 10996, USA
cDepartment of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
dNIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA
eDepartment of Chemical and Biomolecular Engineering, University of Delaware, Newark, DE 19716, USA. E-mail: armstm@udel.edu
First published on 25th March 2025
Interest in novel energy storage and conversion methods has prompted a broad interest in potential applications of conductive, complex materials such as graphene oxide slurries. Investigating the complex rheological, material, and chemical properties of chemically exfoliated graphene oxide suspensions is a potential means to address that interest. The morphological size and clustering, rheology, and electronic conductivity are determined to characterize the properties of graphene oxide (GO) suspensions from variable centrifugation speeds. The evolution of viscosity is then analyzed under oscillatory shear, steady shear, and transient shear characteristics. The resulting microstructure is then analyzed via neutron scattering analysis and imaged with scanning electron microscopy. Small-Angle Neutron Scattering (SANS) of a 500g centrifuged GO suspension determined that particle structure is locally flat sheet-like at lengths below 100 nm, crumpled aggregates of GO sheets with surface roughness at length scales from 200 nm to 2 μm, and a dense mass fractal of overlapping GO sheets extending up to length scales of 20 μm. Increased centrifugation force of the 1000g GO suspension corresponded with lower zero-shear viscosity, yield stress, and less pronounced thixotropic behavior. Rheo-dielectric measurements were conducted on 1000g and 500g GO suspensions to determine the ohmic resistance, electronic conductivity, and specific capacitance. The more fluid-like microstructure of 1000g with smaller monodispered thinning GO sheets in suspension had lower ohmic resistance and higher electronic conductivity compared to the 500g GO suspension with more polydispersed larger aggregates. The 1000g GO suspension had the highest specific capacitance of 4.63 mF cm−2 at the highest shear rate of 700 s−1 due to the higher frequency of particle–particle collisions during shear within the network of smaller and more intrinsically conductive GO sheets to store charge. Therefore, the results of this study have implications for future studies in flowable carbon nanomaterials in flow battery and flow capacitor technologies.
The influence of the surface chemistry, crystal structure, and particle dimension of microstructure on the conductivity and viscosity of a nanoparticle suspension has been demonstrated in CB particles.2 When mixed into an aqueous dispersion, CNPs begin to aggregate, forming a slurry, leading to complex interactions between the individual particles. Optimizing particle agglomeration is necessary to ensure the slurry holds charge with minimal possible viscosity.2 At high viscosities, the utility of CNP slurries could be limited by the decreased ease of pumping and increased material quality requirements due to internal friction.7 Forming hierarchical structures with CB–CNT composite assemblies can act to reduce viscosity as the hybrid chains possess the advantages of both CB and CNTs, conferring decreased shear forces of CB and increased relative capacitance of CNTs at lower particle concentrations.8 Viscosity can also be tuned by coating CB with a dispersal agent, inhibiting agglomeration to promote uniform particle dispersion in the fluid. However, dispersal agents can decrease the conductive properties of the CNP slurry.9
Alternate methods of improving the conductivity of nanoparticle slurries can include the integration of other materials into the CNP dispersion. A common example of this can be seen in the mechanical mixing of poly(methyl methacrylate) (PMMA) with CB, a process that acts to increase the localization of the CNPs via the formation of small gaps for electron flow between the introduced PMMA molecules.10 While decreased shearing resistance can negatively affect the flow characteristics of liquid-phase conductive polymer composites (CPCs), covalent bonding in solid CPC structures positively affects mechanical strength while simultaneously improving electron transport for conductive polymer materials.11,12 Regardless of the particular CNP or CPC chosen for use, maximum faradaic efficiency can only be achieved by deliberately tailoring device electrodes to the unique properties of the particular slurry.13,14
Recent investigations have revealed GNPs as another viable nanomaterial for energy storage applications due to their expansive surface area affording them a high charge capacity and capacity for rapid electron exchange.1,15,16 Scalable methods for the synthesis of GO include chemical vapor deposition, direct liquid-phase exfoliation (electrochemical intercalation and shearing of graphite), and, most commonly, the Hummers' method.17–19 The most popular variant of the Hummers' method is the “modified” or “improved” Hummers' method.20–24 The aqueous GO suspension can also be reheated or evaporated to further concentrate the graphene sheets and mixed with other conductive materials such as Nafion polymer, creating more effective conductive pathways by altering the microstructure or interparticle interactions of graphene.25 However, commercially sourced GNPs come at far greater expense than CB, with the particles incurring average costs exceeding $15 USD per gram compared to $0.01 USD per gram.26–28
CNPs applied to the storage, transport, and conduction of charge must be stored within an electrolyte dispersion to allow the free flow of its constituent CNPs without the deleterious influence of the van der Waals forces and permit the agglomeration of the CNPs.29 When charged, a conductive slurry of nanoparticles can be further aggregated via water addition or, more commonly, sonication.19,30 An example of sonication can be seen in the fraying and exfoliation of nanoparticles from CNT/ZnO chains into electrolyte fluid.30 These free particles implant themselves on the exposed edges of micelles, isolating them and decrease overall fluid viscosity. Nonetheless, it is vital to balance the maximization of CNP density and the minimization of CNP chain degradation in the slurry.30 Careful sonication of the CNT dispersion is necessary as extended sonication or high frequencies can destroy the nanotubules, rendering them amorphous masses of nanofiber.31,32
As CNP agglomerates build up, the viscosity of the nanoparticle slurry generally increases, demonstrating complex thixo-elasto-visco-plastic (TEVP) flow properties.33 Thixotropy describes the time-dependent evolution of the viscosity of a fluid under shear effects on microstructure. Viscoelasticity represents the simultaneity of viscous and elastic characteristics in a material. Plasticity characterizes the threshold of non-zero yield stress above which a fluid undergoes permanent, irreversible deformation; below the threshold, the material deforms elastically.34–37 Effectively examining the rheology of CNP slurries necessitates a model of the steady-state and transient behavior of a suspension. The Generalized Newtonian Fluid (GNF) models such as the Casson, Carreau–Yasuda, and Herschel–Bulkley models can represent many aspects of steady-state flow behavior but prove unable to discern the transient viscoelastic behaviors of complex flowing materials such as GO dispersions.38,39 As such, accurately modeling the transient response phenomena produced by thixotropy and viscoelasticity requires more sophisticated constitutive models.40,41
The thinness of the monoatomic layers of GO permits exceptionally high electron mobility, facilitating the uninhibited flow of electrons through graphene sheets and resulting in outstanding conductivity, with low resistivities (corresponding to high conductivities) on the order of a mΩ cm.42 In exploring GO slurries under different flow conditions, the interplay between rheology and bulk electrochemical properties becomes a focal point, specifically the potential impact of shear-induced microstructure buildup and breakdown on ion diffusion. As GO sheets fragment in response to imposed shear stresses, the observed shear-thinning behavior may influence the ability of ions to conduct through the slurry. The dynamic stacking and restacking of numerous GO sheets in the slurry may also further contribute to the overall conductivity of the material.42,43
The flowability of GO slurries, crucial for redox flow batteries, also influences the material capacitance and electrronic conductivity. The effective particle concentration of the slurry is subject to alteration under varying flow conditions and evolving microstructure, making it a key parameter in the electrochemical performance of the material. The maintenance of a well-structured and flowable electrode comprised of nanoparticles with robust mechanical stability, characteristic of GO, is imperative for sustaining strong electrochemical performance over time.43,44 Moreover, the flowability characteristics of GO slurries can have broader implications for battery design. These include the optimal coating of substrates within flow batteries to form electrodes and the formation of conductive films.43,44 For electrochemical flow battery and flow capacitor applications, carbon nanomaterials such as GO are used as a conductive additive to electrolyte resulting in a flowable slurry electrode, therefore there is a fundamental need to study the flow rate dependent charge transfer through particle–particle interaction.45–47 For the optimization of the electrochemical performance of flowable electrodes in energy storage applications, it is critical to increase the rate of charge transfer in the slurry while lowering the viscosity. There have been recent studies that have focused on suspensions of CB to exhibit shear-thinning properties and an increase in conductivity by characterizing the shear-induced microstructure by probing the rheological and frequency-based dielectric properties with applied shear rates.48–52
The electrical percolation and mechanical percolation of suspensions can be a result of different particle–particle interactions and length scales. Results of early studies on CB showed both increasing and decreasing conductivity with increasing shear rate.49,51 The increase in conductivity was attributed to an increase in the frequency of particle–particle collisions due to shear.48,50,51 Contrary results indicating a decrease in conductivity was due to a decrease in the size of aggregates or the volume fraction of particles in the fluid.51,52 Percolation theory suggests a decrease in conductivity is a result of a decrease in agglomerate size or volume fraction of particles in a fluid.53 The conductivity with the shear of micron-sized particles such as CB has been well investigated. There are limited studies in probing the conductivity changes and microstructural properties of nano-sized CNPs such as GO in response to shear. Electron transfer can occur through a network microstructure through tunnelling or hopping during shear. Electron hopping occurs in short distances of about 1 nm at ambient temperatures, thus, nano-sized particles are the ideal material to measure it.54 Therefore, understanding the shear-dependency on viscosity, conductivity, and microstructure of a suspension can improve the performance of flowable slurries for electrochemical energy storage applications. The frequency-dependent dielectric response of GO suspensions under shear together with the changes in mechanical properties and Rheo-SANS measurements provide an insight into the role of microstructure on the conductivity of the particles in suspension. To our knowledge, there are limited multi-disciplinary studies on probing the length scales of the microstructure, transient rheological properties, and electronic conductivity and capacitance of nanometer-sized GO suspensions with shear rates.
To bridge the current gap in the robust quantitative characterization of CNP slurry electrochemistry in flow, this investigation links GO slurries' electrochemical characteristics and mechanical behavior through the collection and analysis of morphological, rheological, and electrochemical performance results. This study focuses on investigating the properties of GO in aqueous solutions at atmospheric pressure and room temperature conditions relevant operating conditions to electrochemical energy storage device applications. Fig. S9† contains Scanning Electron Microscopy (SEM) images of GO sheets after heated drying at 80 °C showing less prevalence of aggregated GO sheets settling onto substrate likely caused by the high temperature drying process. Section II details the mathematics of the ethixo-mHAWB model and the experimental parameter fitting procedure used to fit and predict the rheological data. Section 3 provides an overview of the materials and methods, with a discussion of the experimental process. Section 4 discusses the results of experiments, followed by the discussion. The remainder of the manuscript serves to conclude the study.
The elasticity and plasticity terms are resolved by splitting total strain and its time derivative, the total rate of strain, into two separate terms.59,62,64 Kinematic hardening theories of plasticity can be used to capture the plastic behavior of a material, integrating both isotropic hardening (IH) and kinematic hardening (KH) into the model description.64–69 The employment of representative equation parameters for the two hardening terms in the ethixo-mHAWB (modified Horner–Armstrong–Wagner–Beris) model used here is vital in properly modeling the dynamics of fluid systems such as slurries.69–71
Characterizing the rheology of CNP slurries plays a vital role in enhancing our understanding of the microstructural responses of slurries to stress and strain under different conditions. As a viscoelastic material, the response can be analyzed (i) at steady-state, (ii) small amplitude oscillatory shear (SAOS), representing a fairly limited region of viscoelasticity where outputs vary linearly with inputs, and (iii) large amplitude oscillatory shear (LAOS), which captures the larger spans of the viscoelastic domain.72–76 The breakdown process of the suspension provides insight into time-dependent flow characteristics. Nonlinear measurements explored by LAOS are decomposed using Fourier transforms, which typically utilize sine inputs for controlled strain and cosine inputs for controlled stress.72,77–80
The structure of CNPs can be analyzed via rheo-ultra-small-angle neutron (Rheo-USANS) scattering. This details both flow stability and the threshold at which dense agglomerates evolve into more open, smaller variants (transition state) which is informative in assessing the conductive behavior of slurries. This is expressed by the inverse Bingham number, Bi−1, which compares the bulk stress imposed on a suspension to its yield stress. At lower shear rates, Bi−1 < 1, the Bingham number indicates the accumulation of particles in relatively large, dense particles that tend to sediment under dominant yield stress, producing analytic inconsistencies. At Bi−1 > 1, corresponding to high shear rates, the aggregates are smaller, demonstrating an apparent shear-thinning as the yield stress is overcome, indicating the shear-independence of the fractal dimensions of the agglomerates.81
With the relative increase in the availability of transient experimental data for complex materials, models have been developed to represent the spectrum of TEVP behavior.56 One such model, the HAWB model, integrates a generalized White–Metzner–Cross model to better represent the viscoelasticity of deformable particles within a suspension. From the time-varying viscosity Maxwell model (TVM) model, a further improved model introducing the viscoelastic response of a material was developed in the mHAWB method.41 A subsequent modification to the kinetic equations of mHAWB to include a shear aggregation term for the structure parameter would result in the enhanced thixotropic mHAWB (ethixo-mHAWB) model shown here.41,71
A full, and more detailed ethixo-mHAWB model description can be found in the following ref. 40, 41, 71 and 73. Here, we will provide an overview of the steady state and transient constitutive equations to ground the subsequent discussion of the fitted data in the rheological underpinnings of the model. The model is a White–Metzner viscoelastic model, integrated with the Cross model to describe the viscosity as a function of shear rate.40,41,56 This baseline framework for the stress tensor evolution and the rate-dependent viscosity describes the suspension with individual GO particles, or aggregates before they begin forming larger-scale agglomerates during periods of low shear rate.
![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
![]() | (5) |
![]() | (6) |
![]() | (7) |
σtot = σyx,ve + σyx,A,ve + σyx,A,e | (8) |
![]() | (9) |
This thixotropic modeling incorporates the separating of strain and shear rate into elastic and plastic components from literature37,40,41,64–66 where γ = γe + γp and =
e +
p and is further described by,
![]() | (10) |
γmax = γ0,Aλ, | (11) |
![]() | (12) |
The elastic contribution from the GO microstructure to total stress is
![]() | (13) |
![]() | (14) |
γ(t) = γ0![]() | (15) |
![]() ![]() | (16) |
![]() | (17) |
![]() | (18) |
Ultra-small-angle neutron scattering (USANS) measurements were performed on the BT5-USANS instrument at the NCNR.88 The slit-smeared USANS intensity was measured over a significantly lower q-range, spanning from approximately 0.00003 Å−1 < q < 0.003 Å−1, which provided an overlapping q-region with the 13 m (lenses) SANS configuration between 0.003 Å−1 and 0.001 Å−1. The slit-smeared USANS data were de-smeared to compare the SANS and USANS on the same absolute intensities. All data were normalized to an absolute intensity I(q) with units of inverse cm−1 sr−1 (denoted here only as cm−1 as the inverse steradian is redundant in this application) after subtracting the scattering from the empty cell and blocked beam using the NCNR reduction procedures in Igor software.89 Lastly, the q-independent incoherent scattering (B) was determined at the high-q plateau and subtracted from the total intensity to give the q-dependent coherent scattering, I(q) − B. The combined SANS and USANS curves were fit using the DREAM algorithm in SasView (version 5.0.5) using a linear-weighted sum of the “lamellar” model and the “correlation length” model.90 The details of the model equation, fit parameters, and parameter uncertainties are discussed further in the ESI.†
![]() | (19) |
![]() | (20) |
![]() | (21) |
The small amplitude oscillatory shear is predicted, and Fcost,SAOS computed as follows,
![]() | (22) |
The qualitative features of the apparent power-law slopes are described in Table S2 of ESI and by eqn (S2)–(S4)† are discussed in three different regions of q, demarcated by vertical dotted lines in Fig. 1c. These regions correspond to high-q (0.006 < q < 0.5 Å−1), mid-q (0.0002 < q < 0.006 Å−1), and low-q (0.00003 < q < 0.0002 Å−1). Within the high-q region, an observed power-law of I(q) ∼ q−2 indicates scattering from a uniformly flat, sheet-like region of the GO primary particles. The best-fit thickness determined from the lamellar model parameter (δ = 1.1 nm) agrees with the expected thickness of single GO sheets. In the mid-q region at length scales larger than 100 nm, a different power-law exponent (mq = −3.2 to −3.6) is observed up to an apparent correlation length of 540 nm. This region corresponds to scattering from a surface fractal with a relatively roughened topology which suggests that the graphene sheets do not remain flat in dispersion at length scales above 100 nm and instead begin to fold, crumple, or overlap. At the low-q region below the correlation length, the power-law scaling exponent (n = −3.2) continues up to the largest length scale probed by USANS (2π/qmin ≈ 20 μm). This surface-fractal scaling corresponds to scattering from the rough surface of larger GO clusters. Since a low-q plateau (Guinier region) is not observed, the maximum cluster size can be inferred to have a lower limit of its radius of gyration, Rg,max > 20 μm. Overall, the in situ structure determined from SANS and USANS agrees with the dried structure observed from SEM. Both techniques indicate the percolation of large clusters containing sheet-like graphene oxide primary particles.
Contrast variation experiments were performed to determine the neutron scattering length density (SLD) of GO particles. These experiments were accomplished by collecting SANS measurements of five different solvent ratios (20% to 100% by volume D2O in H2O) at the same GO concentration. The resulting particle SLD was estimated as ρp = 8.3 ± 0.4 × 10−6 Å−2, as shown in Fig. S2.† Using the estimated particle SLD and the known solvent SLD (ρH2O = −5.6 × 10−7 Å−2), the total scattering invariant of this two-component system could be computed and used to estimate the total volume fraction of graphene oxide sheets in the 500g sample. The scattering invariant was estimated to be 2.3 × 10−4 cm−1 Å−3, and the total volume fraction of scattering centers (0.15% by volume GO) was estimated after determining the particle SLD and scattering invariant. Lastly, the specific surface area of dispersed particles in solvent (particle surface area per particle volume) is determined by Sv = CP(2π)−1(ρP – ρH2O)−2, in which CP is the Porod constant determined at the limiting high-q region from a linear fit of I(q) q4 as a function of q4 (CP = 2.9 ± 1.8 × 10−4 cm−1 Å−4). Assuming the GO particles retain a density of 0.26 g cm−3, the specific surface area of GO particles in dispersion was estimated to be Sv = 220 ± 140 m2 g−1. Details of the neutron contrast variation measurements, scattering invariant, volume fraction, Porod constant, and specific surface area estimations are described in the ESI.†
Parameter | Description | 500g GO | 1000g GO |
---|---|---|---|
μ0,s (Pa s) | Zero shear viscosity of suspension & ind. particles | 0.071 | 0.011 |
μ∞,s (Pa s) | Infinite shear viscosity of suspension & particles | 0.006 | 0.004 |
τC (s) | Cross time constant | 0.047 | 0.052 |
tr1 (s) | Ratio of shear breakdown to Brownian build-up constant | 0.384 | 0.991 |
tr2 (s) | Ratio of shear aggregation to Brownian build-up constant | 0.133 | 0.058 |
μA (Pa s) | Agglomerate viscosity | 0.263 | 0.079 |
σy0 (Pa) | Yield stress | 0.229 | 0.024 |
d | Power law for shear aggregation (fixed parameter) | 0.5 | 0.5 |
m | Power law for effect of aggregate viscosity (fixed parameter) | 1.5 | 1.5 |
γ∞ (Pa s) | Infinite strain (fixed parameter) | 1 | 1 |
τλ (s) | Time constant of thixotropic evolution | 1.112 | 2.590 |
GA (Pa) | Elastic modulus of agglomerates | 1.013 | 8.845 |
Gs (Pa) | Elastic modulus of suspension and particles | 4.689 | 5.250 |
Fcost (SS) | Normalized cost function of steady state fit | 0.0101 | 0.0056 |
n (points) | Number of data points of steady state fit | 16 | 16 |
RSS | Residual sum of squares for steady state fit | 0.0260 | 0.0081 |
k (par.) | Number of parameters fit to steady state data | 7 | 7 |
Fcost (trans) | Cost function of LAOS transient parameter fit | 0.0024 | 0.0011 |
n (points) | Number of data points of transient parameter fit | 3992 | 3992 |
RSS | Residual sum of squares for transient fit | 89.441 | 19.026 |
k (par.) | Number of parameters (or cycles) fit to LAOS, transient data | 3 | 3 |
An analysis of Table 1 below quantifies what is shown graphically in Fig. 2 and 3. There is a significant drop in yield stress, and the infinite, zero, and aggregate viscosities. The yield stress drops by an order of magnitude due to the ostensible decrease in the amount of structure present in the 500g vs. the 1000g GO samples. In addition, analysis of the Fcost suggests that the use of the ethixo-mHAWB model for this system seems to have been more justified for the 1000g GO system.40,41,60,94,95 Which possesses lower zero-shear (or fully structured) and infinite-shear (or fully-unstructured) viscosities when compared to the 500g slurry.
Fig. 4a and c display the frequency sweeps (performed at γ0 = 10), while Fig. 4b and d show the amplitude sweeps (performed at ω = 12.6 rad s−1), for the 500g and 1000g graphene oxide systems, showing the data and ethixo-mHAWB model predictions. Qualitatively, the ethixo-mHAWB model seems to capture material viscoelasticity as the frequency increases at constant strain amplitude per Fig. 4a and c, and as the strain amplitude increases at constant frequency per Fig. 4b and d. However, Table 2 makes it clear that based on SAOS Fcost values, the model more accurately predicts the 1000g frequency and amplitude sweeps. This demonstrates again that the lower centrifugation magnitude of the 500g sample has led to a markedly different microstructure from the 1000g sample and that we are nearing the limits of the ethixo-mHAWB model's accuracy for the higher viscosity 500g system.
Parameters | 500g GO | 1000g GO |
---|---|---|
Fcost,SAOS | ||
Freq sweep; γ0 = 10 (–) | 0.0097 | 0.0121 |
I3/I1 | 0.0100 | 0.0090 |
Amp sweep; ω = 12.566 rad s−1 | 0.0710 | 0.0084 |
I3/I1 | 0.0111 | 0.0117 |
![]() |
||
Fcost,LAOS | ||
(ω = 1) | 0.0203 | 0.0033 |
γ0 = 0 | 0.0260 | 0.0035 |
γ0 = 1 | 0.0230 | 0.0022 |
γ0 = 5 | 0.0087 | 0.0025 |
γ0 = 10 | 0.0166 | 0.0031 |
γ0 = 50 | 0.0275 | 0.0051 |
γ0 = 100 | 0.0671 | 0.0069 |
Fcost, averaged | 0.0337 | 0.0079 |
Fig. 5 and Table 2 below show the results of the LAOS predictions of the six sets of LAOS data at ω = 100 rad s−1 and γ0 = 5, 10, 50, and 100. Qualitatively the model predicts the 1000g LAOS data better than the 500g data. This is again corroborated by the LAOS Fcost values comparison in Table 2.
Fig. 6a and b are the Cole–Cole plots of the 500g and 1000g graphene oxide suspensions, respectively. Fig. 6c shows the color mapping of δt, a solid-like metric and Fig. 6d shows the Cole–Cole legend.93 The Cole–Cole figures are once again demonstrating significant differences in elastic and viscous properties between the 500g and 1000g samples. For example, the magnitude of the stress itself is different by one order of magnitude, and Fig. 6a shows a much more solid-like substance at all the frequency and strain amplitude combinations (from the δt color comparison), and the fact that the 500g demonstrates more persistent solid-like behavior during the oscillation at higher strains. As we follow a cycle from steps 1 through 4 over a half-period, we can see that from 1 to 2 to 3 we are in a more solid-like configuration, and moving to point 4 takes us to the more liquid-like configuration of the microstructure, as it is broken down over this portion of the LAOS cycle, with an observable structural recovery as the shear rate decreases.
![]() | ||
Fig. 6 (a) Cole–Cole plots of 500g; and (b) Cole–Cole plots of 1000g GO suspensions for six sets of LAOS collected at ω = 1 rad s−1 and γ0 = 0.5, 1, 5, 10, 50 and 100 ![]() |
Fig. 7a–d show the accumulated contour mapping of δt and λ over six sets of LAOS at ω = 1 rad s−1 and γ0 = 5, 10, 50 and 100 for both the 500g and 1000g graphene oxide samples, respectively. The y-axis on each of the subfigures is from 0–6.27 s, the amount of time for a complete LAOS period at ω = 1 rad s−1. We note that there are portions of more ‘liquid-like’ and ‘solid-like’ behavior shown for each figure. The λ values shown in the contour maps below are from the structure predictions of the ethixo-mHAWB model of each LAOS prediction, while the aforementioned δt values are computed with the Sequence of Physical Processes (SPP), using key parameters drawn from previous studies.45,70,82,93,94,96,97
All rheological measurements were performed with an ARESG2 strain-controlled rheometer from TA instruments equipped with the cup and bob geometry. In this analysis, the bob possesses an outer diameter of 27 mm, an inner diameter of 26 mm, and a length of 34 mm. For rheological experiments, 10 mL of the 500g, and 1000g GO suspension in water was loaded into the cup portion of the ARESG2. The electrical percolation threshold of the GO suspensions in H2O was determined by probing the frequency-dependent resistance to determine the conductivity. Since the GO suspensions are in a water fluid phase, the contribution of the measured resistance is electronic charge transfer without any ionic contribution.
The samples consisted of 1000g GO in H2O (∼0.9 wt% GO in H2O) and 500g sample (∼0.1 wt% GO in H2O after 500g centrifugation) within a frequency range of 20 Hz to 2 MHz. A stock solution was first made with a low centrifugation speed of 100g for 5 min to remove any undissolved large aggregates by using the resulting supernatant. There was an observed yield stress just by inverting the vial of the GO suspension and appeared to be strongly shear thinning. From the rheo-dielectric measurements, the impedance (Z) (ohms), the frequency (f) (Hz) and phase angle (φ) (radians) were determined. A brief conceptual overview of the EIS method to determine the Nyquist plot can be found in ESI Section 3.† From this data, the resulting complex plane plots of Nyquist plots of imaginary versus real impedance can be calculated using the following equations:
Re(Z) = Z![]() | (23) |
Im(Z) = Z![]() | (24) |
Since the focus of the data analysis is the determination of the real ohmic resistance for each condition, determining where the x-intercept occurs for the calculation of conductivity is the most important aspect of the Nyquist plots. Analysis of the Nyquist plots determined the ohmic resistance contribution with 0 s−1, 0.1 s−1, 1 s−1, 10 s−1, 100 s−1, and 700 s−1 shear rates. Since each rheo-dielectric measurement involves a frequency sweep from 20 Hz to 2 MHz five times at each shear rate, each Nyquist plot contains the real impedance as the x-intercepts corresponding to the ohmic resistance of the GO suspension. The x intercept of the 1000g GO sample occurred at around a frequency of 10023.74 ± 0.01 Hz and the 500 g GO sample occurred at a frequency of around 12
619.15 ± 0.01 Hz. The plus or minus (±) Hz refers to the accuracy of dielectric measurement from instrumentation. Table 3 shows each average contribution. The ohmic resistance of the 500g and 1000g at zero applied shear rate is the electronic resistance of the innate microstructure of the GO suspension. The ohmic resistance decreased in the 1000g GO suspension with increasing shear rate. However, the ohmic resistance increased with increasing shear rate in the 500g GO suspension. These changes in ohmic resistance for both 1000g and 500g GO suspensions is due to the difference in the microstructure during steady shear. Determining the ohmic resistance of a flowable electrode suspension can help inform slurry electrode design to minimize ohmic power loss and maximize the electronic conductivity necessary for charge transfer in various flowable electrochemical energy storage devices. Overall, the 1000g GO suspension had a lower ohmic resistance is attributed to the higher loading of GO sheets with 0.9 wt% compared to the more dilute 0.1 wt% 500g GO since interparticle interactions tend to be strongest and most affected by the imposed flow at higher concentrations resulting in enhanced electrical percolation to facilitate the movement of electrons across finite distances despite being under shear. This is evident from the conductivity results in Table 4 comparing 1000g to 500g GO suspensions. Moreover, the rheology results in Section 4.3 show the 1000g GO suspension as being less viscous than 500g and therefore less solid-like since most of the larger particulate has been separated in the sample centrifugation process.
Shear rate (s−1) | 1000g GO [Ω] | 500g GO [Ω] |
---|---|---|
0 | 0.84321 | 2.31628 |
0.1 | 0.82216 | 2.34627 |
1 | 0.82265 | 2.40244 |
10 | 0.83461 | 2.41894 |
100 | N/A | 2.45011 |
700 | 0.83231 | 2.38442 |
Shear rate (s−1) | Conductivity [S cm−1] | |
---|---|---|
1000g GO | 500g GO | |
0 | 0.81867 | 0.29803 |
0.1 | 0.83963 | 0.29422 |
1 | 0.83913 | 0.28734 |
10 | 0.82710 | 0.28538 |
100 | N/A | 0.28175 |
700 | 0.82939 | 0.28951 |
From this ohmic resistance data, the conductivity of each GO dispersion within the rheometer geometry is calculated by taking the inverse of the resistivity. Since the diameter of the bob is 27.7 mm, the cross-sectional area A of the bob is calculated at 6.02628 cm2. Eqn (S1) (seen in the accompanying ESI)† can then be used to solve for resistivity using the length of the bob l as given by instrument specifications and resistance R from Table 3 above. Taking the inverse of the resistivity value from eqn (S1)† results in the conductivity value for each test in [S cm−1] shown below in Table 4 and Fig. 10. A comparison of reported electronic conductivity values of GO from literature is included in Table S4.† Moreover, as shown in Table 4, the conductivity of the 1000g GO suspension increased with shear rate while the conductivity of the 500g GO suspension generally decreased with shear rate. As confirmed with the rheology results in Section 4.3, the 1000g GO suspension contains more of a finer microstructure comprised of less aggregates and smaller GO sheets because of the centrifugation process. The separation process resulted in a supernatant with more monodispersed GO particles with fewer graphene layer to create the 1000g suspension. These thinner GO sheets have higher intrinsic particle conductivity due to less resistance of the delocalized electrons allowing for more electron mobility through the conjugated double bond structure. Therefore, the 1000g GO suspension has smaller electrical percolation microstructure with enhanced charge transfer through shear-induced collisions of the thinner GO sheets. The thinner and smaller GO particles form a liquid-like microstructure in the 1000g suspension with shorter distances for electron transfer leading to a higher conductivity. The conductivity of the 1000g suspension increased with shear due to more intimate particle–particle points of contact through the collisions taking place between the smaller and thinner GO sheets that are in suspension. However, the conductivity of the 500g GO suspension was four times less with applied shear compared to the 1000g. This lower conductivity in 500g is from the presence of aggregates in suspension and with shear there are less points of contact between GO sheets in the microstructure. The 500g suspension is more polydispersed with larger GO particles forming larger aggregates under shear and more void space in suspension because of the fewer points of contact between the aggregates resulting in higher ohmic resistance for electron transfer and lower electronic conductivity.
Another method used to analyze the internal resistance contributions is to calculate ASR or area-specific resistance (Ω cm2). Table 5 shows the ASR values for each test under different shear rates and concentrations by multiplying the measured resistance values from Table 3 and the calculated bob area using the given instrument parameters. The ASR of the GO suspensions assess the efficacy and performance of GO as a slurry electrode for flowable electrochemical applications.
Shear rate (s−1) | ASR [Ω cm2] | |
---|---|---|
1000g GO | 500g GO | |
0 | 5.08142 | 13.95855 |
0.1 | 4.95459 | 14.13926 |
1 | 4.95754 | 14.47776 |
10 | 5.02961 | 14.57722 |
100 | N/A | 14.76506 |
700 | 5.01576 | 14.36921 |
From the results of the Nyquist plots in Fig. 8, the specific areal electrochemical capacitance (Csp) in mF cm−2 of the 500g and 1000g GO suspensions was determined. Under a constant shear rate across a cross-sectional area within a frequency range of 2 MHz to 20 Hz, the Csp was determined using the following eqn (25) where f is the frequency in Hz, Z” is the imaginary component of impedance, and A is the geometric area of the rheometer bob in cm2:101,102
![]() | (25) |
The specific areal capacitance as a function of frequency for the GO suspensions under shear of 0 s−1, 0.1 s−1, 1 s−1, 10 s−1, 100 s−1, and 700 s−1 is plotted in Fig. 9. Across all the applied shear rates in Fig. 9, the specific areal capacitances of the GO suspensions were the highest for 500g GO at 12.6 kHz and the highest for 1000g at 10 kHz. With increasing frequency, the time for charge transfer decreases since the distance for charges to move also decreases.103,104 In Fig. 9a, the specific areal capacitance of the 500g and 1000g GO suspensions with a zero shear rate at a low frequency of 502 Hz is 0.027 mF cm−2 and 0.028 mF cm−2, respectively. With a sweep to higher frequencies, there is a noticeable large increase in specific areal capacitance of 0.97 mF cm−2 for 500g GO at 12.6 kHz and 0.67 mF cm−2 for 1000g at 10 kHz which is due to the charge build up in the governing percolating microstructure of GO sheets as primary particles in suspension.
![]() | ||
Fig. 10 Conductivity in S cm−1 comparison for 1000g GO (blue) and 500g GO (red) at different shear rates. |
The Csp peaks at 10 kHz and 12.6 kHz for 1000g and 500g GO suspensions, respectively, are the characteristic flip frequencies which determine the Debye relaxation time for the 1000g and 500g GO suspensions to discharge to an equilibrium within an applied electric field.105,106 The transition from long to short distance charge carrier mobility occurs at these characteristic maximum frequencies.105–107 The Csp in both the 500g and 1000g GO suspensions is mostly constant at frequencies below and above the characteristic flip frequency (∼104 kHz) at zero applied shear. The lower Csp at frequencies below the characteristic maximum flip frequency (<104 kHz) is due to the charge carriers having to move longer distances through the microstructure suspension.105,106 After applying shear to the 500g and 1000g GO suspension, the charge storage capacitance lowers when shifting to higher frequencies beyond the maximum due to charge carrier confinement within short distances.105,106 This asymmetric specific capacitance is due to a non-Debye broader relaxation or discharge time because of the charge diploes on the GO particles.105,106 Despite the lower conductivity compared to the 1000g suspension as shown in Table 3, there was a larger capacitance increase in the 500g GO suspension. This is due to the charge storage within the aggregates of GO in the 500g suspension without any leakage current at low shear rates. There is a shift to a higher flip frequency for the more viscous 500g with more aggregates in suspension since the charge transfer within the microstructure does not occur instantly since it takes time to align dipole moments within the larger GO aggregates.
The specific areal capacitance decreased drastically when sweeping to higher frequencies to 2 MHz indicating a dissipation of this charge build-up in the microstructure of the percolating large clusters of GO sheets. This governing microstructure of the GO sheets in suspension is disrupted by applying a steady shear rate to the GO suspensions thus, impacting the GO sheet–sheet interaction for charge transfer and storage. Therefore, an increase in shear rate of 0.1 s−1, 1 s−1, 10 s−1, and 100 s−1 applied to both the 500g and 1000g GO suspensions decreased the specific areal capacitance at low frequencies such as 502 Hz. Unlike the pristine 500g and 1000g suspensions, there was not a large increase in capacitance with sweeping to higher frequencies such as 10 kHz for 1000g to 12.6 kHz for 500g. The capacitances of the 500g and 1000g in the table insets in Fig. 9b–9e were lower in magnitude at 10 kHz and 12.6 kHz compared to the capacitance of the undisturbed GO suspensions in Fig. 9a. As shown in Fig. 9f, at the largest applied shear rate of 700 s−1, the specific areal capacitance of the 1000g GO suspension was the largest in magnitude across all applied shear rates with 4.63 mF cm−2. This dispersed microstructure present in the 1000g GO suspension allows for bulk charge transport and not limited to local electron transfer events among particle aggregates. The large specific capacitance at the highest shear is due to the disruption of a system spanning microstructure network of thin GO sheets resulting in an increased frequency of particle–particle collisions. At the highest shear rate of 700 s−1, there are smaller aggregates with more frequent collisions in suspension with higher inter-aggregate distance. The specific capacitance differences between the 1000g and 500g suspensions is dependent on the time scales for electron transfer as a function of the frequency-dependent dielectric response through electron transfer through either localized sites amongst particles or bulk transfer consistent with percolation and the presence of aggregates. This was evident for the 500g GO suspension which had the lowest capacitance of 0.15 mF cm−2 in higher shear rates and at a higher frequency of 12.6 kHz. Overall, there was decrease in the Csp at the characteristic maximum flip frequency with increased shear rate of 0.1 s−1, 1 s−1, 10 s−1, and 100 s−1 for all 500g GO suspensions. The same trend for Csp at the characteristic maximum flip frequency was observed for the 1000g GO suspension with shear rate until a very large increase at the highest shear rate of 700 s−1. Similarly, there was a slight increase of capacitance of the 500g GO suspension of 0.13 mF cm−2 at 100 s−1 and 0.15 mF cm−2 at the highest shear rate of 700 s−1. Therefore, there is better shear alignment with the more concentrated suspensions of GO such as 500g where interparticle interactions in the aggregates are the strongest and most affected by the imposed high shear rate at higher concentrations of sheets leading to a higher capacitance of charge storage. However, the increase in specific capacitance in the 1000g GO suspension is due to the likely formation of smaller aggregates of smaller GO sheets with more frequency of collisions during shear collectively storing charge minimizing leakage current.
The 1000g GO dispersion is found to be more conductive than the 500g GO dispersion across all shear rates. This suggests an inverse relationship between increasing centrifugal force and slurry viscosity and conductivity. The previous SANS and rheological analyses suggest that the 500g sample has more residual structure, aggregates, and agglomerates compared to the 1000g GO sample. Due to the dispersed nature of the particle network present in the 1000g GO sample, it will demonstrate less robust mechanical properties compared to the less aligned, more entangled network of the 500g sample. However, the higher centrifugation forces likely permit better intraparticle alignment of the GO flakes allowing for greater ease of electron flow resulting from enhanced conductive pathways. While it could be assumed that the decreasing size of agglomerates with an increasing shear rate would change the conductivity characteristics, it did not impact conductivity for either GO sample.
As such, we judge from the results that effective conductivity and linked properties such as resistance and impedance are not observably dependent on microstructure agglomeration states in GO dispersions. The observable timescales of structural buildup and breakdown are likely far outpaced by conduction suggesting that the flow of electrons throughout the material is contingent upon the construction of the percolated network of the GO particles themselves rather than interparticle interactions. If it is the case that electron transfer pathways are limited to the particles themselves, the resulting electronic resistance may stem from the gaps between particles in suspension.
By fitting the ethixo-mHAWB model to rheometric data, we have been able to predict the complex behavior of the graphene oxide dispersion while simultaneously translating our understanding of bulk properties to a more concrete understanding of how microstructure contributes to system shear-thinning, thixotropy, and viscoelasticity. This was linked to subsequent analysis of the microstructure by SANS and SEM, revealing the aggregation behavior of the individual graphene oxide sheets from their smallest individual morphologies to the crumpled, irregular forms of larger multi-particle agglomerates. The 500g sample was found to have higher zero-shear viscosity, yield stress, and more pronounced thixotropic behavior. Analysis suggested that the more concentrated and dispersed 1000g sample demonstrated a low impedance and maximum electrochemical capacitance at the characteristic flip frequency at 10023.74 ± 0.01 Hz versus 12
619.15 ± 0.01 Hz. Additionally, it was found that the higher concentration of GO promoted an average of 65% higher conductivity and lower resistance compared to the 500g sample. In this work, the primary particle, aggregate, and agglomerate differences between the microstructures were examined and subsequently revealed the factors most relevant to the flow and microstructural properties of the GO slurry, concurrently demonstrating an inverse correlation between increasing centrifugal force and slurry viscosity and conductivity, informing the use of such GO materials for application in future flow battery technologies. Given this understanding of the relationship between flow and electrochemical performance, the design of these prospective battery technologies can be honed to maximize their effectiveness. New lines of research can be developed in using carbon-based and other flowable nanomaterials for those applications.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra08308c |
This journal is © The Royal Society of Chemistry 2025 |