Hostname: page-component-848d4c4894-ttngx Total loading time: 0 Render date: 2024-05-07T18:09:13.610Z Has data issue: false hasContentIssue false

Effect of hydrodynamic slip on the rotational dynamics of a thin Brownian platelet in shear flow

Published online by Cambridge University Press:  20 May 2021

Catherine Kamal
Affiliation:
School of Engineering and Material Science, Queen Mary University of London, LondonE1 4NS, United Kingdom
Simon Gravelle
Affiliation:
School of Engineering and Material Science, Queen Mary University of London, LondonE1 4NS, United Kingdom
Lorenzo Botto*
Affiliation:
Process and Energy Department, 3ME Faculty of Mechanical, Maritime and Materials Engineering, TU Delft, 2628 CDDelft, The Netherlands
*
Email address for correspondence: l.botto@tudelft.nl

Abstract

The classical theory by Jeffery predicts that, in the absence of Brownian fluctuations, a thin rigid platelet rotates continuously in a shear flow, performing periodic orbits. However, a stable orientation is possible if the surface of the platelet displays a hydrodynamic slip length $\lambda$ comparable to or larger than the thickness of the platelet. In this article, by solving the Fokker–Plank equation for the orientation distribution function and corroborating the analysis with boundary integral simulations, we quantify a threshold Péclet number, ${Pe}_{c}$, above which such alignment occurs. We found that for ${Pe}$ smaller than ${Pe}_{c}$, but larger than a second threshold, a regime emerges where Brownian fluctuations are strong enough to break the platelet's alignment and induce rotations, but with a period of rotation that depends on the value of $\lambda$. For ${Pe}$ below this second threshold, slip has a negligible effect on the orientational dynamics. We use these thresholds to classify the dynamics of graphene-like nanoplatelets for realistic values of $\lambda$ and apply our results to the quantification of the orientational contribution to the effective viscosity of a dilute suspension of nanoplatelets with slip. We find a non-monotonic variation of this term, with a minimum occurring when the slip length is comparable to the thickness of the particle.

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2021. Published by Cambridge University Press

1. Introduction

The flow behaviour of thin plate-like particles is of interest in many industrial and environmental applications, ranging from the processing of composite materials (Kumar, Sharma & Dixit Reference Kumar, Sharma and Dixit2019) to the transport of clay in natural waters (Tawari, Koch & Cohen Reference Tawari, Koch and Cohen2001). Recently, the emergence of graphene and other two-dimensional (2-D) nanomaterials and their use in a variety of liquid-based processes (White et al. Reference White, Hawkins, Miyamoto, Takahara and Sue2015; Del Giudice & Shen Reference Del Giudice and Shen2017; Koltonow et al. Reference Koltonow, Luo, Luo and Huang2017; Karagiannidis et al. Reference Karagiannidis2017) has spurred renewed interest in the dynamics of these extremely thin plate-like colloids when suspended in sheared liquids (Xu & Green Reference Xu and Green2014; Poulin et al. Reference Poulin, Jalili, Neri, Nallet, Divoux, Colin, Aboutalebi, Wallace and Zakri2016; Reddy et al. Reference Reddy, Natale, Prud'homme and Vermant2018; Gravelle, Kamal & Botto Reference Gravelle, Kamal and Botto2021; Silmore, Strano & Swan Reference Silmore, Strano and Swan2021). In applications such as graphene inks or polymer nanocomposites, colloidal 2-D nanomaterials take the form of platelets of nanometric thickness and lateral size spanning from a few nanometres to a few microns (Wick et al. Reference Wick2014). There is a great interest in understanding how the application of an external flow affects the orientation of these highly anisotropic particles in the presence of Brownian fluctuations. The surface of graphene and other 2-D nanomaterials can be characterised by a hydrodynamic slip length $\lambda$ of several nanometres for many common solvents (Maali, Cohen-Bouhacina & Kellay Reference Maali, Cohen-Bouhacina and Kellay2008; Ortiz-Young et al. Reference Ortiz-Young, Chiu, Kim, Voïtchovsky and Riedo2013; Tocci, Joly & Michaelides Reference Tocci, Joly and Michaelides2014), where $\lambda$ is the distance within the solid at which the relative solid–fluid velocity extrapolates to zero (Bocquet & Barrat Reference Bocquet and Barrat2007); the mathematical definition of $\lambda$ is given in § 4.3. For comparison, the typical thickness of 2-D material is $< 1\ \textrm {nm}$. The aim of the current study is to understand the interplay between hydrodynamic slip and Brownian fluctuations in determining the rotational dynamics of a thin plate in an unbounded, simple shear flow under creeping flow conditions.

According to Jeffery's (Reference Jeffery1922) classical theory, derived for the high-Péclet number limit in which Brownian fluctuations are negligible, an oblate ellipsoid with its normal in the flow plane rotates continuously, performing a characteristic tumbling dynamics (figure 1a). Bretherton (Reference Bretherton1962) generalised the high-Péclet, low-Reynolds-numbers theory of Jeffery to axisymmetric particles of general shape. Bretherton's analysis showed that for axisymmetric shapes the rotational dynamics is governed by a scalar parameter that depends, for a no-slip particle, only on the shape of the particle. For instance, for an oblate ellipsoidal particle of major semiaxis $a$ and minor semiaxis $b$, the scalar parameter identified by Bretherton corresponds simply to the geometric aspect ratio $b/a$. However, when hydrodynamic slip is present, Bretherton's scalar parameter (which we shall name $k_{e}$ in the current article) is not only a function of $b/a$ but must also depend on the slip length $\lambda$ (Zhang, Xu & Qian Reference Zhang, Xu and Qian2015; Kamal, Gravelle & Botto Reference Kamal, Gravelle and Botto2020). This brings about the question of how slip affects Jeffery's predictions, and how Brownian fluctuations alter the rotational dynamics when the slip length is large in comparison with the particle thickness. This question is particularly relevant considering the nanometric thickness of 2-D nanomaterial particles.

Figure 1. Sketch illustrating the rotational dynamics of a platelet having (a) zero slip length, as described by Jeffery's theory and (b) slip length $\lambda$ significantly greater than the platelet's half-thickness $b$. The behaviours described in panels (a,b) are characteristic of the infinite Péclet number limit. In this paper we enquire about the effect of Brownian fluctuations on the rotational dynamics.

Hydrodynamic slip at solid boundaries is known to affect the motion of small rigid particles in fluids. Slip reduces the tangential hydrodynamic stress on the particle surface (Sellier Reference Sellier2013), resulting in a reduction in the translational and rotational drag coefficient of spherical or anisotropic particles (Youngren & Acrivos Reference Youngren and Acrivos1975; Loyalka & Griffin Reference Loyalka and Griffin1994; Allison Reference Allison1999; Keh & Huang Reference Keh and Huang2004; Keh & Chang Reference Keh and Chang2008; Sellier Reference Sellier2012; Chang & Keh Reference Chang and Keh2012; Sherwood Reference Sherwood2012). Slip also results in a slowdown of the rotational dynamics of spherical particles freely suspended in a simple shear flow (Luo & Pozrikidis Reference Luo and Pozrikidis2008). A similar slip-dependent slowdown in a shear flow was predicted by Zhang et al. (Reference Zhang, Xu and Qian2015) for a particle with an elliptic cross-section ($b/a \sim 0.5$) and infinite extent in the vorticity direction, in the limit of a slip length smaller than the particle thickness ($\lambda \sim b / 5$).

Recently, we studied the motion of a thin ($b/a \ll 1$) rigid graphene-like nanoparticle with hydrodynamic slip (Kamal et al. Reference Kamal, Gravelle and Botto2020), via a combination of molecular dynamics (MD) and continuum simulations. In contrast to previous studies (Zhang et al. Reference Zhang, Xu and Qian2015), we considered values of the slip length either larger or smaller than the particle thickness. By using numerical and asymptotic solutions of the low-Reynolds-number boundary integral equations governing the hydrodynamic surface stress distribution, we showed that provided that the half-thickness of the platelet $b$ is smaller than $\lambda$, in the limit of infinite Péclet ($Pe$) numbers the platelet aligns at a small angle $\phi _{c}$ with respect to the flow direction, instead of performing full rotational cycles as predicted by Jeffery's theory for ${Pe}\to \infty$ (figure 1b). Our asymptotic analysis revealed that this behaviour is due to a perturbed balance between two hydrodynamic torque contributions. One contribution is due to tangential hydrodynamic stresses acting on a lever arm proportional to $b$. The second contribution is due to normal hydrodynamic stresses acting on a lever arm proportional to $a$. For $\lambda /b \ll 1$ the torque due to tangential stresses is larger than that due to normal stresses, resulting in an expected rotation in the same direction of the vorticity vector (Singh et al. Reference Singh, Koch, Subramanian and Stroock2014). However, we demonstrated that to leading order in $b/a$ slip induces a reduction in the tangential stresses, without affecting the normal stresses significantly. As a consequence, for $\lambda /b \gtrapprox 1$ the torque due to normal stresses dominates, resulting in a rotation in the direction opposite to the vorticity vector (for a particle initially aligned in the flow direction, as illustrated in figure 1b) and the attainment of an equilibrium angle $\phi _{c}$ in correspondence to which the two torque contributions balance exactly (an illustration of such arrested dynamics is given in § 4). This surprising result was confirmed with MD simulations. While the theory we developed was for infinite values of $Pe$ for which the effect of Brownian fluctuations is negligible, we observed a stable orientation in MD simulations of relatively short graphene nanoplatelets for $Pe$ of approximately $100$ (Kamal et al. Reference Kamal, Gravelle and Botto2020). This raises the question of what is the minimum $Pe$ for which a stable orientation can be observed.

In this paper we first recall the essential elements of the high-$Pe$ number theory we developed previously. Then we analyse the rotational dynamics for finite $Pe$ numbers, by considering solutions of the Fokker–Plank equation governing the single-particle orientational distribution function. To simplify the problem, following Leahy, Koch & Cohen (Reference Leahy, Koch and Cohen2015) we assume that the motion of the particle is in the flow plane, hence the orientational distribution is a function only of the angle with respect to the flow direction and time. Earliest work on the use of the Fokker–Plank equation for predicting the orientational distribution of elongated particles in a shear flow is due to Burgers (Reference Burgers1938) and Peterlin (Reference Peterlin1938), who examined the effects of Brownian fluctuations on the dynamics of ellipsoids and rods for ${Pe} \ll 1$. Other authors have examined the range ${Pe} \gg 1$ (Leal & Hinch Reference Leal and Hinch1971; Hinch & Leal Reference Hinch and Leal1973; Leahy et al. Reference Leahy, Koch and Cohen2015) and the range of intermediate ${Pe}$ (Sadron Reference Sadron1953; Scheraga Reference Scheraga1955; Hinch & Leal Reference Hinch and Leal1972; Férec et al. Reference Férec, Heniche, Heuzey, Ausias and Carreau2008), considering ellipsoids (both oblate and prolate) and more general rod-like particles. An important result of these theoretical and numerical analyses is that for elongated particles almost aligned with the flow, Brownian fluctuations increase the average particle's rate of rotation with respect to the high-$Pe$ value, essentially by allowing the particle to escape from the region of small streamwise velocity. Theories for the rotational dynamics of elongated particles based on the seminal work of Jeffery have been confirmed experimentally for a range of $Pe$ numbers by standard rheological measurements (Ivanov, Van de Ven & Mason Reference Ivanov, Van de Ven and Mason1982; Mueller, Llewellin & Mader Reference Mueller, Llewellin and Mader2009), rheo-optics (Frattini & Fuller Reference Frattini and Fuller1986; Fuller Reference Fuller1995; Vermant, Yang & Fuller Reference Vermant, Yang and Fuller2001; Reddy et al. Reference Reddy, Perez-Juste, Pastoriza-Santos, Lang, Dhont, Liz-Marzan and Vermant2011) and by direct observation of the dynamics of single particles (Anczurowski & Mason Reference Anczurowski and Mason1967b; Stover, Koch & Cohen Reference Stover, Koch and Cohen1992; Herzhaft & Guazzelli Reference Herzhaft and Guazzelli1999; Leahy et al. Reference Leahy, Cheng, Ong, Liddell-Watson and Cohen2013). We are not aware of experimental work on the rotational dynamics of plate-like particles specifically focusing on the effect of slip.

The structure of the paper is as follows. In § 2 we analyse Bretheton's equation of motion. We show that this equation can be used to describe the rotational dynamics of plate-like particles with slip, provided that when $\lambda /b$ is larger than a threshold value the effective aspect ratio of the particle is taken to be a complex number. In § 3, we consider numerical solutions of the Fokker–Plank equations for the orientational distribution function. The objective of this section is to illustrate the qualitative features of the particle dynamics and quantify the range of $Pe$ above which the effect of slip-dependent hydrodynamics becomes important. Such analysis enabled us to estimate the threshold $Pe$ giving a stable orientation. In § 4, we apply our theory to the dynamics of nanoplatelets whose geometry mimics that of single and multilayer graphene, for practically relevant values of $\lambda$. In this section, we quantify the threshold shear stress needed to prevent the platelet from completing full rotations. Finally, we analyse how the indefinite alignment of the particles affects the orientational contribution to the effective viscosity of a diluted suspension of platelets with slip.

2. Formulation of the problem

We consider the rotational dynamics of a rigid plate-like particle (henceforth referred to as ‘platelet’) in a simple shear flow. The platelet has length $2a$ and thickness $2b$. We work under conditions of Stokes flow and neglect particle inertia and gravity (i.e., the particle is force and torque free). In our theoretical derivations, we assume that the geometric aspect ratio $k={b}/{a}$ is much smaller than one. A consequence of this is that the rotational dynamics of a three-dimensional (3-D) platelet can be well approximated by the rotational dynamics of a 2-D platelet of infinite extent in the vorticity direction, provided that the motion of the particle occurs in the plane of the flow and the depth of the platelet's extent in the vorticity direction is not small in comparison with its length (Kamal et al. Reference Kamal, Gravelle and Botto2020). We assume that the particle is symmetric about two planes (lines in two dimensions) passing through the particle centre. The particle motion is induced by an undisturbed shear flow field $\boldsymbol {u}_{\infty }=\dot \gamma y \boldsymbol {\hat {e}}_{x}$, where $\boldsymbol {\hat {e}}_{x}$ is the unit normal in the flow direction and $\dot {\gamma }$ is the shear rate.

The probability $p$ of finding the platelet at a certain angle with respect to the flow, in response to both hydrodynamic and Brownian stresses, is governed by a Fokker–Planck (Smoluchowski) equation (Gardiner Reference Gardiner2004). With just one degree of freedom, the Fokker–Planck equation for the orientational probability distribution function simplifies to a second-order partial differential equation,

(2.1)\begin{equation} \dot{p} = \left[D_{r} p_\phi - \dot{\gamma} \varOmega p\right]_{\phi}, \end{equation}

where $\phi$ is the anticlockwise orientation angle with respect to $\boldsymbol {\hat {e}}_{x}$, $t$ is time, $\dot {p}=\partial p/\partial t$, $[\;]_{\phi }=\partial [\;]/\partial \phi$, $\varOmega (\phi )$ is the hydrodynamic angular velocity non-dimensionalised by $\dot {\gamma }^{-1}$ and $D_{r}$ is the rotational diffusion coefficient. The ratio of convective and diffusive fluxes in (2.1) is the Péclet number ${Pe}=\dot {\gamma }/D_{r}$.

2.1. Hydrodynamic angular velocity

The hydrodynamic angular velocity $\dot {\gamma }\varOmega$ is required to close (2.1). In the dilute limit, $\varOmega$ can be evaluated by examining the motion of an isolated freely suspended platelet in the absence of Brownian fluctuations. For a symmetric platelet, $\varOmega$ can be calculated exactly by using Bretherton's (Reference Bretherton1962) equation of motion (Kim & Karrila Reference Kim and Karrila2013). This equation describes the relationship between the time derivatives of the particle orientation vector $\boldsymbol {d}$, and the rate of strain and vorticity tensors associated with the undisturbed flow field, $\boldsymbol {\varOmega _{\infty }}=(\boldsymbol {\nabla } \boldsymbol {u}_{\infty }+\boldsymbol {\nabla } \boldsymbol {u}^{T}_{\infty })$ and $\boldsymbol {E}_{\infty }=(\boldsymbol {\nabla } \boldsymbol {u}_{\infty }-\boldsymbol {\nabla } \boldsymbol {u}^{T}_{\infty })$, respectively, as follows:

(2.2)\begin{equation} \dot{\boldsymbol{d}}=\dot{\gamma}\boldsymbol{\varOmega}\times\boldsymbol{d} = \boldsymbol{\varOmega_{\infty}}\times\boldsymbol{d}+\left(\frac{k_{e}^2-1}{k_{e}^2+1}\right)\left(\boldsymbol{E}_{\infty}\boldsymbol{\cdot}\boldsymbol{d} -\boldsymbol{E}_{\infty}\,\boldsymbol{d}\boldsymbol{d}\,\boldsymbol{d} \right). \end{equation}

Using spherical polar coordinates, the orientation vector can be expressed as $\boldsymbol {d}=\sin {\theta }\cos {\phi } \boldsymbol {\hat {e}}_{x} +\sin {\theta }\sin {\phi } \boldsymbol {\hat {e}}_y+\cos {\theta }\boldsymbol {\hat {e}}_{z}$, resulting in the following coupled ordinary differential equations (Kim & Karrila Reference Kim and Karrila2013):

(2.3)\begin{equation} \dot{\theta}=\left(\frac{1-k_{e}^2}{1+k_{e}^2}\right)\frac{\dot{\gamma}}{4}\sin{2\theta}\sin{2\phi}, \quad \dot{\phi}\equiv \dot{\gamma}\varOmega={-}\frac{\dot{\gamma}}{k_{e}^2+1}\left(k_{e}^2\cos^2{\phi}+\sin^2{\phi}\right). \end{equation}

Note that $\dot {\phi }$ does not depend on $\theta$, hence an analysis of the particular 2-D case considered here corresponding to $\theta ={\rm \pi} /2$ is representative of the $\phi$-dependence for all values of $\theta$. The derivation of Bretherton's equation of motion does not make assumptions regarding the boundary condition at the solid surface and therefore applies to platelets with slip. This equation depends on the value of $k_{e}$, which in turn depends on the slip length as well as on the platelet's shape (Luo & Pozrikidis Reference Luo and Pozrikidis2008; Kamal et al. Reference Kamal, Gravelle and Botto2020). The parameter $k_{e}$ can be calculated for $\theta ={\rm \pi} /2$ as the ratio between the hydrodynamic torques acting on a particle held fixed with its long axis held parallel to the flow, $T(0)$, or perpendicular to the flow, $T({\rm \pi} /2)$, according to the following expression (Cox Reference Cox1971):

(2.4)\begin{equation} k_{e}=\sqrt{\frac{T(0)}{T({\rm \pi}/2)}}. \end{equation}

For a platelet satisfying the no-slip boundary condition, $T(0)/T({\rm \pi} /2)$ is positive and therefore $k_{e}$ is a real number (Willis Reference Willis1977). In this case, when the geometric aspect ratio $k = b/a \ll 1$, the parameter $k_{e}$ follows a power-law relationship $k_{e} \propto k^n$ where $n$ is an exponent that depends on the specific geometry of the platelet. For this reason, $k_{e}$ is often called the ‘effective aspect ratio’. For example, an ellipsoid with a no-slip surface has $k_{e}=k$ (Jeffery Reference Jeffery1922), and a disk with blunt edges has $k_{e}\propto k^{3/4}$ (Singh et al. Reference Singh, Koch, Subramanian and Stroock2014). For real $k_{e}$ the solution to (2.3) is

(2.5a,b)\begin{equation} \tan{\theta}=\dfrac{C k_e}{\sqrt{k_e^2 \cos^2{\phi}+\sin^2{\phi}}}, \quad \tan{\phi}={-}k_{e}\tan{\left(\frac{\dot{\gamma}t}{k_{e}+k_{e}^{{-}1}}\right)}, \end{equation}

where $C$ is a positive integration constant. These equations describe periodic rotations of the particle called ‘Jeffery orbits’; the time period of such rotations is

(2.6)\begin{equation} P=2{\rm \pi}\dot{\gamma}^{{-}1}(k_{e}+k_{e}^{{-}1}). \end{equation}

The value of $C$ depends on $\theta (0)$; for $\theta (0) = {\rm \pi}/2$, $C\to \infty$ and $\theta (t)={\rm \pi} /2$ for all values of $t$. Thus, a particle initially rotating in the plane of the flow will continue rotating in the plane of the flow. This is the situation we are aiming to model in the current paper. For finite $Pe$ numbers, trajectories out of the $\theta = {\rm \pi}/2$ plane are of course possible. The case $\theta = {\rm \pi}/2$ for finite $Pe$ is tractable analytically and has been shown, in the no-slip case, to be relevant to the full 3-D dynamics (Leahy et al. Reference Leahy, Koch and Cohen2015). In § 3 we shall consider the general slip case at finite $Pe$.

An example of an orbit for $\theta = {\rm \pi}/2$ and real $k_{e}$ and the corresponding $\varOmega (\phi )$ are given in figures 2(a) and 2(b), respectively. Since $k_{e} \ll 1$, the time scale for the particle to flip is much smaller than the time spent by the particle near $\phi =0$ (Singh et al. Reference Singh, Koch, Subramanian and Stroock2014). For a no-slip platelet of vanishing thickness, $k_{e} \to 0$ as $k \to 0$ and thus the particle will remain in a fixed position in the ad hoc case $k=0$ (Bretherton Reference Bretherton1962). Otherwise the platelet will rotate (unless the platelet has a specially designed, slightly non-axisymmetric shape, see Borker, Stroock & Koch (Reference Borker, Stroock and Koch2018)).

Figure 2. (a) Rotation angle $\phi$ for a particle with real effective aspect ratio $k_{e} = 0.16$ and ${Pe} = \infty$. (b) Associated angular velocity $\dot {\gamma }\varOmega$; notice that $\varOmega <0 \ \forall \phi$. (c) Rotation angle $\phi$ for a particle with imaginary effective aspect ratio $k_{e} = 0.28 \text {i}$ and ${Pe} = \infty$, showing stabilisation at $\phi _{c} \approx 0.27$. (d) Corresponding non-dimensional angular velocity, with the two equilibrium points at $\phi = \pm \phi _{c}$. The ‘${\bigodot}$’ symbol marks the stable angle $\phi _{c}$; the angle denoted by ‘${\bigotimes}$’ is unstable.

For platelets presenting a surface with slip, however, such alignment is possible (Kamal et al. Reference Kamal, Gravelle and Botto2020). The effect of surface slip is to reduce the value of $k_{e}^2$, such that the $\varOmega (\phi )$ curve shifts upwards, reducing the frequency of rotation with respect to the no-slip case (Luo & Pozrikidis Reference Luo and Pozrikidis2008; Zhang et al. Reference Zhang, Xu and Qian2015; Kamal et al. Reference Kamal, Gravelle and Botto2020). At or above a critical slip length that scales with the thickness of the platelet, $k_{e}^2$ becomes negative, and so $\varOmega (\phi )$ will be shifted onto or above the zero line (figure 2d). Such a solution corresponds to a negative value of $T(0)/T({\rm \pi} /2)$ and a purely imaginary complex value of $k_{e}$ (Kamal et al. Reference Kamal, Gravelle and Botto2020). It is instructive to compare the structure of the $\phi -\theta$ equations (describing full 3-D trajectories) for pure and imaginary values of $k_{e}$. For a purely imaginary $k_{e}$, the solution of (2.3) is

(2.7a,b)\begin{equation} \tan{\theta}=\frac{C|k_{e}|}{|k_{e}|^2\cos^2{\phi}+\sin^2{\phi}},\quad \tan{\phi}=|k_{e}|\tanh{\left(\frac{\dot{\gamma}t}{|k_{e}|^{{-}1}-|k_{e}|}\right)}. \end{equation}

Comparing with (2.5a,b), whilst the equation for $\theta$ remains unchanged with respect to the case of a real $k_{e}$, the equation for $\phi$ now admits a stable non-periodic solution for $t \rightarrow \infty$. The particle relaxes to the stable orientation angle $\phi _{c}=\arctan (|k_{e}|)$, as shown in figure 2(c). For this angle, $\varOmega$ has a positive zero root, indicated by the open circle in figure 2(d). The relaxation to equilibrium occurs on a time scale

(2.8)\begin{equation} P_{R}=\dot{\gamma}^{{-}1}(|{k}_e|^{{-}1}-|k_{e}|), \end{equation}

to be compared with the rotational time period of (2.6). For $|{k}_e|\ll 1$, $P$ and $P_{R}$ are proportional to each other. We note that a stable orientation has been predicted for blood cells in shear flow in the case of small internal-to-external viscosity ratios (due to the so-called ‘tank-treading’ motion, see e.g. Keller & Skalak (Reference Keller and Skalak1982)). The analogy with our case is, however, only qualitative, because there is no single slip parameter that can model the dynamic coupling between inner and outer fluids seen in blood cells.

Surface slip at infinite Péclet number thus has significant effects on the rotational dynamics of a platelet. When the slip length is smaller than the platelet's thickness, slip reduces the frequency of the rotational orbits through a decrease in the value of $k_{e}$. However, when the slip length is greater than a threshold value comparable to the platelet's thickness, the platelet aligns indefinitely at an angle $\phi _{c}$ with the flow.

3. Rotational dynamics at finite Péclet numbers

At finite $Pe$, both the rotational diffusion coefficient $D_{r}$ and the effective aspect ratio $k_{e}$ change due to slip. However, in the limit $k\to 0$, $D_{r}$ becomes approximately independent of $\lambda$ (Sherwood & Meeten Reference Sherwood and Meeten1991; Sherwood Reference Sherwood2012). Therefore, for thin platelets, the rotational dynamics at finite $Pe$ changes primarily due to slip through the dependence of $k_{e}$ on the hydrodynamic angular velocity.

3.1. Random walk trajectories

To illustrate the effects of Brownian fluctuations on the platelet's rotational trajectories, we solve numerically the Langevin representation of (2.1) for different $Pe$ and $k_{e}$. In finite difference form, such Langevin representation describes simply connected paths in the particle orientational space. Using a first-order approximation of the time derivative, the resulting stochastic finite difference equation is (Doi & Edwards Reference Doi and Edwards1988)

(3.1)\begin{equation} \Delta \phi= \dot \gamma \varOmega(\phi) \Delta t + X. \end{equation}

Here, we take $X$ as a Gaussian random variable with zero mean and variance $\left \langle X^2\right \rangle = 2 D_{r} \Delta t$. Solving numerically this equation for several sequences of random numbers provides a set of trajectories distributed according to (2.1). Examples of trajectories calculated from (3.1) for different $Pe$ are shown in figure 3, comparing the cases of real and purely imaginary $k_{e}$. Full rotational cycles, corresponding to a platelet travelling through the full range $\phi \in [-{\rm \pi} /2, {\rm \pi}/2]$, can always be observed if $k_{e}$ is real. When $k_{e}$ is real, the average time period (in units of $Pe$ $t$) for the platelet to perform a full rotational cycle decreases as $Pe$ decreases (figure 3ac). This behaviour has been predicted in previous theoretical studies on the steady-state rotatory dynamics of plate-like particles (Leal & Hinch Reference Leal and Hinch1971; Hinch & Leal Reference Hinch and Leal1972), and is in agreement with experiments where the instantaneous rotations of elongated particles have been visualised directly (Stover et al. Reference Stover, Koch and Cohen1992; Herzhaft & Guazzelli Reference Herzhaft and Guazzelli1999; Leahy et al. Reference Leahy, Cheng, Ong, Liddell-Watson and Cohen2013), or through rheo-optical or neutron diffraction techniques (Frattini & Fuller Reference Frattini and Fuller1986; Jogun & Zukoski Reference Jogun and Zukoski1999; Brown et al. Reference Brown, Clarke, Convert and Rennie2000). In contrast, if $k_{e}$ is purely imaginary, for large $Pe$ (${Pe}=640$), the platelet fluctuates around a stable orientation angle (figure 3df). This angle is approximately the same as the equilibrium angle for $Pe\to \infty$ ($\phi _{c}=\arctan {|k_{e}|}$). When $Pe$ decreases below a threshold value, Brownian fluctuations cause the platelet to perform full rotational cycles. However, the average frequency of these full rotational cycles is significantly reduced in comparison with the no slip case, as seen by the reduction of the number of peaks in figure 3(df) as compared with figure 3(ac) for ${Pe}=64$. For ${Pe} \sim O(1)$, Brownian fluctuations induce a chaotic behaviour whereby the platelet rotates frequently and randomly. In this range, the trajectories for a purely imaginary and a real $k_{e}$ are practically indistinguishable (see ${Pe}=6.4$ in figure 3).

Figure 3. Time evolution of the rotation angle for different $Pe$, as calculated from (3.1), comparing (ac) real $k_{e}$ ($k_{e}=0.11$) to (df) purely imaginary $k_{e}$ ($k_{e}=0.21\textrm {i}$). The red dashed line is the value of $\phi _{c}$ obtained for ${Pe} \to \infty$.

Figure 3 suggests that for large $Pe$, the rotational dynamics depends strongly on whether $k_{e}$ is real or purely imaginary, while for smaller $Pe$ such dependency disappears. Quantifying the values of $Pe$ needed for the rotational dynamics to depend on the value of $k_{e}$ requires solving the Fokker–Planck (2.1).

3.1.1. Time evolution of the orientation distribution function's moments

Taking advantage of the periodicity of $p(\phi ,t)$, any solution of (2.1) can be expressed as a spectral series,

(3.2)\begin{equation} p(\phi,t)=a_0(t)+\sum_{n=1}^{\infty} \left[ a_n(t)\cos{(2n\phi)}+b_n(t)\sin{(2n\phi)}\right]. \end{equation}

Normalisation of $p$ requires $a_0=1/{\rm \pi}$, and assuming $\phi$ to be uniformly distributed at time $t=0$ gives an initial uniform distribution $p(\phi ,0)=1/{\rm \pi}$. In spectral coordinates, this is equivalent to having only one non-zero coefficient $a_0=1/{\rm \pi}$. Numerically, we solve (2.1) by truncating the series at a large value $N$, yielding $2N$ first-order differential equations for the spectral coefficients (Leahy et al. Reference Leahy, Koch and Cohen2015). A truncated value $N=1000$ has been chosen from the analysis of the spectral coefficients $a_n$ and $b_n$. For $n>N$, $a_n, b_n \to 0$ for all values of time $t$ (see figure 4 for $t \to \infty$). The $2N$-coupled equations are solved in time by a Runge–Kutta method (Press et al. Reference Press, Teukolsky, Vetterling and Flannery2007). The mean angle, standard deviation and higher moments can be evaluated directly in terms of the spectral distribution. For example, the mean and variance are

(3.3a,b)\begin{equation} \mu=\left\langle \phi\right\rangle=\sum_{i=1}^{\infty}\frac{\rm \pi}{2n}({-}1)^{1+n}b_n, \quad \sigma^2=\left\langle (\phi-\mu)^2\right\rangle=\frac{{\rm \pi}^3}{12}+\sum_{i=1}^{\infty}\frac{\rm \pi}{2n^2}({-}1)^{n}a_n-\mu^2, \end{equation}

respectively. Here, the close brackets $\left \langle \;\right \rangle$ represent an average over the orientation distribution function $p(\phi ,t)$ for fixed $t$.

Figure 4. Spectral coefficients $a_n$ and $b_n$ for $k_e=0.05\text {i}$ and $Pe=10^{6}$ and $t \to \infty$. The spectral coefficients are truncated at $n=1000$, since $a_n, b_n \to 0$ above this threshold.

Many studies discuss solutions to (2.1) for elongated particles with real $k_{e}$ (Burgers Reference Burgers1938; Peterlin Reference Peterlin1938; Sadron Reference Sadron1953; Scheraga Reference Scheraga1955; Leal & Hinch Reference Leal and Hinch1971; Hinch & Leal Reference Hinch and Leal1972, Reference Hinch and Leal1973; Férec et al. Reference Férec, Heniche, Heuzey, Ausias and Carreau2008; Leahy et al. Reference Leahy, Koch and Cohen2015; Leahy, Koch & Cohen Reference Leahy, Koch and Cohen2017). However, the case of purely imaginary $k_{e}$, to our knowledge, has not yet been analysed. The time evolution of the mean, variance and third-order moment $\left \langle (\phi -\mu )^3\right \rangle$ of the orientation distribution for $k_{e}=0.1\text {i}$ and different $Pe$ is shown in figure 5. Similarly to when $k_{e}$ is real (Leahy et al. Reference Leahy, Koch and Cohen2015), the moments relax in time to a steady value, with a relaxation time that decreases as $Pe$ decreases. Therefore, we can assume that the statistical steady state is reached over a time greater than the relaxation time $P_R=\dot {\gamma }^{-1}(|k_{e}|^{-1}-|k_{e}|)$ of a platelet in absence of Brownian motion.

Figure 5. Mean (a), variance (b) and skewness (c) of the angular probability distribution versus time, for a nanoplatelet with $k_{e}=0.1\text {i}$ and different values of $Pe$.

3.2. Steady-state probability distribution

We solve (2.1) with $\dot {p}=0$ to find the steady-state orientation distribution function $p(\phi )$ using the method described in § 3.1.1. The effect of $Pe$ on $p(\phi )$ comparing real and purely imaginary values of $k_{e}$ is shown in figure 6. For ${Pe} = 1000$, there is a noticeable difference in the shape of the orientation distribution function depending on whether $k_{e}$ is real or purely imaginary. In particular, $p(\phi )$ has a much sharper peak if $k_{e}$ is purely imaginary. This difference can be analysed by considering the limit $Pe\to \infty$. For ${Pe}\to \infty$, (2.1) reduces at steady state to

(3.4)\begin{equation} p\varOmega = c, \end{equation}

for some integration constant $c$. For real $k_{e}$, normalisation of $p$ gives $p=(P\dot {\gamma } | \varOmega |)^{-1}$, where $P$ is the time period (see (2.6)) (Anczurowski & Mason Reference Anczurowski and Mason1967a). Such a probability distribution shown by the (red) dotted line in figure 6 has mean orientation angle $\mu =0$ and finite variance $\sigma ^2$ by the symmetry of $p$ about $\phi =0$. If $k_{e}$ is purely imaginary, however, $1/\varOmega$ diverges as $\phi \to \pm \phi _{c}$, and thus the integrand of $c/\varOmega$ will not be finite unless $k_{e}=0$ or $c=0$, as required by the normalisation constraint for $p$. Thus, the equation $p\varOmega =c$ is only satisfied for $c=0$ and $p(\phi )=\delta (\phi -\phi _{c})$, as expected. As seen in figure 6 for the imaginary $k_{e}$, as $Pe$ increases the probability distribution function indeed converges to a Dirac delta, confirming that for sufficiently large $Pe$ the platelet spends most of the time fluctuating about $\phi _{c}$ (as observed in figure 3df).

Figure 6. Steady-state orientation distribution function $p(\phi )$ for a real $(k_e = 0.1)$ or a purely imaginary $(k_e = 0.1\textrm {i})$ value of $k_{e}$ for different $Pe$ numbers. The red dotted line marks the orientation distribution function for ${Pe} \to \infty$ and $k_{e}=0.1$ (a) and the location of $\phi _{c}$ for $k_{e}=0.1 i$ (b). The values [$\mu$, $\sigma ^2$, $\left \langle (\phi -\mu )^3\right \rangle$] in increasing order of $Pe$ are $[0.17,0.39,-0.36]$, $[0.069,0.21,-0.40]$ and $[0.019,0.14,-0.30]$, respectively, for $k_{e}=0.1$; and $[0.19,0.38,-0.37]$, $[0.094,0.16,-0.53]$ and $[0.075,0.033,-1.1]$, respectively, for $k_{e}=0.1\text {i}$.

For ${Pe} = 10$ the orientation distribution functions for $k_{e}=0.1$ and $k_{e}=0.1\text {i}$ are almost identical, and slightly skewed towards positive values of $\phi$ (the value of $\mu$, $\sigma ^2$ and $\left \langle (\phi -\mu )^3\right \rangle$ for $k_{e}=0.1$ and $k_{e}=0.1\text {i}$, given in the caption of figure 6, are within 10 % of each other). Since $Pe$ is not small in this example, hydrodynamic stresses must play a role in creating such skewness. However, the effect of changes to the value of $k_{e}$ due to surface slip on the probability distribution is evidently minor.

Balancing the orders of magnitude of the convective and diffusive fluxes appearing in (2.1) yields a characteristic angle $\varPhi$. From this, we can define a ‘local’ $Pe$ number, ${Pe}_{loc} = \dot {\gamma }\varOmega (\phi ) \varPhi /D_{r}$ (Leal & Hinch Reference Leal and Hinch1971). The effects of Brownian fluctuations on the orientation of the platelets are considered to be ‘weak’ compared with the hydrodynamic angular velocity if, for $\phi$ near zero, ${Pe}_{loc} \gg 1$ (Leal & Hinch Reference Leal and Hinch1971). This condition translates to the Brownian angular velocity being smaller than, using (2.3), $|\varOmega |\sim k_{e}^2$. For $|\varOmega |\sim k_{e}^2$ we have that $\phi$ can be at most $\varPhi \sim |k_{e}|$, and so $|{Pe}_{loc}|$ can be at most of the order of $|k_{e}^3 {Pe}|$.

Thus, for ${Pe} \gg |k_{e}|^{-3}$ Brownian diffusion is subdominant, and the rotation of the platelet can be approximated by the orientation distribution obtained for negligible Brownian fluctuations.

To quantify the value of $Pe$ for which the orientation distribution does not depend on $k_{e}$ to leading order, one can proceed as follows. For a geometric aspect ratio $k\to 0$, we will verify in § 4 that $|k_{e}|\ll 1$ whatever the value of $\lambda$. Expanding the right-hand side of the equation for $\dot {\phi }$ in (2.3) to leading order in $|k_{e}|$, the hydrodynamic velocity is seen to become independent of $k_{e}$ if $\sin^{2} {\phi _m} \gg |k_{e}^2|\cos^{2} {\phi _m}$. Thus, because we are interested in a critical $Pe$ valid for both $\lambda =0$ and finite $\lambda$, the condition for $\phi _m$ is $\tan {\phi _m}\approx \phi _m \gg \max (|k_{e}(\lambda =0)|,|k_{e}(\lambda )|)$. For the rotational dynamics to be independent of $k_{e}$ to leading order, the local Péclet number must be $O(1)$ over the characteristic angle $\varPhi \sim \phi _m$. This regime is associated with a rotational velocity $\varOmega (\phi _m)\approx \dot {\gamma }\phi _m^2$, so the local Péclet number ${Pe} \phi _m^3 \sim 1$. It follows from the condition $\phi _m\gg \max (|k_{e}(\lambda =0)|,|k_{e}(\lambda )|)$ that ${Pe} \ll \min (|k_{e}^{-3}(\lambda =0)|,|k_{e}^{-3}(\lambda )|)$. Thus, Brownian diffusion will control the orientation of the platelet when ${Pe} \ll \min (|k_{e}^{-3}(\lambda =0)|,|k_{e}^{-3}(\lambda )|)$.

The value ${Pe}_{m} = \min (|k_{e}^{-3}(\lambda =0)|,|k_{e}^{-3}(\lambda )|)$ is an important threshold. For ${Pe} \ll {Pe}_{m}$ the value of $k_{e}$ does not affect to leading order the orientation distribution $p(\phi ,t)$. When ${Pe} \gg {Pe}_{m}$ significant changes to the orientation distribution for either a real or purely imaginary value of $k_{e}$ are expected. Figure 7 demonstrates the importance of this threshold on the statistical moments $\mu , \sigma ^2$ and $\left \langle (\phi -\mu )^3\right \rangle$. The black line corresponds to ${Pe}_{m} = |k_{e}|^{-3}$. For $Pe$ above this line, the statistical moments approach the values predicted by using the probability distribution functions corresponding to $Pe \to \infty$. In this case the moments depend strongly on the value of $k_{e}$. For example, if $k_{e}$ is purely imaginary, then the variance $\sigma ^2 \to 0$ and the mean $\mu \to k_{e}$ for $Pe$ above the ${Pe}_{m}=|k_{e}^{-3}|$ line. If instead $k_{e}$ is real then $\sigma ^2$ is instead finite and $\mu \to 0$. Likewise, if $Pe$ is much smaller than $|k_{e}^{-3}|$ the statistical moments depend only on $Pe$ and become essentially independent of $k_{e}$ (compare, for example, the colour maps for $|k_{e}| \leq 0.1$ and ${Pe}\leq 100$ in figure 7).

Figure 7. Colour maps of the mean $\mu$, variance $\sigma ^2$ and skewness $\left \langle (\phi - \mu )^3\right \rangle$ of the steady-state orientation distribution in the ($Pe$, $k_{e}$) space. The black straight line is ${Pe}=|k_{e}|^{-3}$. For $Pe$ above this line, the difference in the moments between real and purely imaginary values of $k_{e}$ becomes noticeable.

Finally, for ${Pe} \ll 1$, the platelet's rotation is almost entirely due to Brownian diffusion. As ${Pe}\to 0$, the leading-order orientation distribution can be described at all times by a uniform distribution (Burgers Reference Burgers1938; Peterlin Reference Peterlin1938; Scheraga Reference Scheraga1955). Such distribution describes randomly rotating platelets that complete an average rotational cycle with a variance much larger than in the pure hydrodynamic flow limit, as also shown in figure 3.

3.3. Threshold Péclet number for full rotations

When the effective aspect ratio $k_{e}$ is purely imaginary, we can define a critical Péclet number ${Pe}_{c}$ separating the stable region, in which the platelet fluctuates about $\phi _{c}$, from the unstable region in which the platelet performs full rotations. Because the hydrodynamic angular velocity is zero for $\phi =\phi _{c}$ and small in the neighbourhood of this critical angle, in the stable region weak Brownian fluctuations always affect the orientation of the platelet for any finite value of $Pe$. When the platelet is in the unstable region, full rotation cycles can occur because the variance $\sigma ^{2}$ is large enough for the fluctuating platelet to ‘jump out’ of the region where $\varOmega (\phi )$ is positive. The width of this ‘hydrodynamic potential well’ is given by the range of values of $\phi$ for which $\varOmega (\phi )$ is positive. This width can be approximated as $2 \arctan (\theta _c) \approx 2 |k_{e}|$ for $|k_{e}| \ll 1$. Therefore, we estimate that the probability for a given particle to perform full rotary cycles becomes significant for $\sigma$ larger than a threshold $\sigma \sim 2 |k_{e}|$.

Using our numerical procedure for calculating $\sigma ^2$, as given in (3.3a,b), we calculate ${Pe}_{c}$ by first computing $\sigma ({Pe},k_{e})$ and then calculating the value of $Pe$ that gives $\sigma =2|k_{e}|$. We emphasise that full rotations of a given platelet can still occur for ${Pe}$ significantly larger than ${Pe}_{c}$, but these events, associated with the tails of the probability distribution, are rare and should not affect the occurrence of a stable orientation on account of the approximate symmetry of $p(\phi )$ about its mean value. Our results for ${Pe}_{c}$ are shown in figure 8 for a range of purely imaginary $k_{e}$. As $|k_{e}|$ decreases, the width of the ‘hydrodynamic potential well’ decreases, so ${Pe}_{c}$ increases. The data appear to be well fitted by ${Pe}_{c}\propto |k_{e}|^{-3.4}$ for small values of $|k_{e}|$. The exponent $-3.4$ is quite close to the exponent $-3$ obtained by balancing the order of magnitude of the convection and diffusive terms in (2.1). The data are expected to approach this theoretical value for values of $|k_{e}|$ smaller than we could simulate. For $|k_{e}| \gtrapprox 0.1$ the local power-law exponent is definitely not close to $-3$. In this region, the line ${Pe}_{c}\propto |k_{e}|^{-4.8}$ appears to be a better fit to the data.

Figure 8. Threshold value of the Péclet number ${Pe}_{c}$ beyond which no platelet rotation is expected, as a function of $|k_{e}|$ (with $k_{e}$ purely imaginary). Black disks are numerical values, lines are fits for $|k_{e}| \gtrapprox 0.1$ (dot–dashed blue line) and $|k_{e}| \lessapprox 0.1$ (dashed red line), respectively.

Figure 8 does not enable us by itself to draw a conclusion about whether a stable orientation occurs for given values of $\lambda$ and $a$, as both $k_{e}$ and $Pe$ depend on these variables. To draw such conclusion we need to map $k_{e}$ and $Pe$ to specific values of $\lambda$ and $a$. This will be done in the next section, by considering particles of a specific geometry and slip properties.

4. Rotational dynamics of model graphene-like nanoplatelets

Our work is motivated by understanding the dynamics of graphene colloids or similar 2-D nanomaterials. Applying our theory to realistic graphene-like nanoplatelets requires evaluating $k_{e}$ for geometric and slip characteristics typical of this specific type of anisotropic colloid. The calculation of $k_{e}$ for platelets of atomic thickness (in the following we will use the word ‘nanoplatelets’ for brevity) must be done with care, because the classical framework developed for mesoscale colloidal particles may break down when one of the particle dimensions is comparable with the size of the liquid molecules. In previous studies (Gravelle, Kamal & Botto Reference Gravelle, Kamal and Botto2020; Gravelle, Kamal & Botto Reference Gravelle, Kamal and Botto2021; Kamal et al. Reference Kamal, Gravelle and Botto2020), we have performed MD of a rigid graphene nanoplatelet suspended in a shear flow of water and other solvents. We showed that $k_{e}$ can still be predicted using a continuum description, provided that the hydrodynamic stress is computed on a suitable reference surface surrounding the platelet and a slip boundary condition is enforced at this surface. Interestingly, we found that using a single slip parameter $\lambda$ is sufficient to capture the effects of the hydrodynamic slip on the particle rotation and hydrodynamic torque, despite the heterogeneous surface of graphene (which has different slip properties at the edges as compared with the planar region). This feature, which we have proved to be related to the weak sensitivity of the hydrodynamic torque on the tangential components of the traction at the edges, is useful as it simplifies the analysis of the problem by reducing the number of parameters.

4.1. Effective nanoplatelet's geometry

Nanoplatelets of 2D materials, made for example of carbon (C), boron nitride (BN), or molybdenum disulfide ($\textrm {MoS}_2$), are essentially stacks of atomic crystal layers. To achieve a continuum description of the nanoplatelets in flow, one must define the reference surface that best approximates such platelets from a hydrodynamic standpoint. Molecular dynamics calculations reveal that, because of the smoothing of the molecular flow field by the finite-interaction potential near the edges, the reference surface is a cuboid with rounded edges, as sketched in figure 9. The cuboid has a half-thickness $b = \xi + d_{gg}(n-1)/2$, where $\xi$ is the effective radius of a single atom of the nanoplatelet, $d_{gg}$ is the interlayer spacing and $n$ is the number of stacks. For multilayer graphene, $d_{gg} \approx 3.35$ Å (Chung Reference Chung2002). The effective radius $\xi$ depends on the equilibrium distance between the atoms of the solid and the liquid molecules. For graphene in water, $\xi \sim 1.8\text {--}2.5$ Å (Gravelle et al. Reference Gravelle, Joly, Ybert and Bocquet2014), and comparable values can be expected for BN and MoS$_2$ due to their similar atomic structure (Radisavljevic et al. Reference Radisavljevic, Radenovic, Brivio, Giacometti and Kis2011; Tocci et al. Reference Tocci, Joly and Michaelides2014; Luan & Zhou Reference Luan and Zhou2016). The edges form a semicircle shape for $n=1$ (figure 9a) and a flat face with blunt edges as $n$ increases (figure 9bd). The shape of the edge slightly affects the value of $k_{e}$: the blunter the edge, the larger the value of $k_{e}$ for a fixed aspect ratio $a/b$ (cf. Singh et al. (Reference Singh, Koch, Subramanian and Stroock2014) for $\lambda =0$).

Figure 9. (ac) Water density profiles as extracted from MD simulations for a number of layer $n=1$ (a), $2$ (b) and $3$ (c), respectively. The centres of the carbon atoms are located along the grey lines. Details of the molecules dynamics simulations are given in Kamal et al. (Reference Kamal, Gravelle and Botto2020). The colour field is the water density, from white (low density) to blue (high density), and the red dashed lines show the reference surfaces. (d) Sketch of the reference surface for $n=3$, showing the slip length $\lambda$, the effective radius $\xi$ of the carbon atoms and the interlayer spacing $d_{gg}$.

4.2. Range of slip lengths

For atomically smooth surfaces of 2-D nanomaterials such as graphene, the hydrodynamics slip length measured at the planar surface of the particle ranges from a few nanometres to tens of nanometres. As summarised in table 1, experiments performed on graphite (Maali et al. Reference Maali, Cohen-Bouhacina and Kellay2008; Ortiz-Young et al. Reference Ortiz-Young, Chiu, Kim, Voïtchovsky and Riedo2013), and ab initio calculations (Tocci et al. Reference Tocci, Joly and Michaelides2014), suggest $\lambda \approx 10\ \textrm {nm}$ for water and graphene. In the case of NMP or ethanol, the slip length of graphene is also relatively large (${>}10\ \textrm {nm}$) (Gravelle et al. Reference Gravelle, Kamal and Botto2020). Ionic liquid (IL) can also give large slip lengths, but in this case $\lambda$ has been found to depend strongly on the shear rate (Voeltzel et al. Reference Voeltzel, Fillot, Vergne and Joly2018).

Table 1. Literature values of the slip length for different solid material/solvent combinations. The materials and liquids are: diamond-like carbon (DIC); graphene oxide (GO); boron nitride (BN); molybdenum disulfide (MoS$_2$); ionic liquid (IL); n-methyl-2-pyrrolidone (NMP); octamethylcyclotetrasiloxane (OMCTS); cyclopentanone (CPO). Ab initio MD refers to a method allowing for the calculation of electronic behaviour from first principles by using a quantum mechanical method.

The slip length $\lambda$ can also be significant for atomically smooth and chemically homogeneous materials other than graphene. The interaction of BN and water gives $\lambda \sim 3\ \textrm {nm}$ (Tocci et al. Reference Tocci, Joly and Michaelides2014). While this value may seem small, it is still larger than the thickness of single-layer BN. Despite the large scatter in the literature data for $\lambda$, consequence of the dependence of MD simulation results on empirical force fields, table 1 suggests that relatively large slip lengths are not uncommon in 2-D nanomaterials, which is essentially a consequence of the fact than many 2-D nanomaterials have atomically smooth surfaces. To provide theoretical guidelines on a range of realistic values, characteristic of those in table 1, we will evaluate the rotational dynamics of nanoplatelets for $\lambda = 0$, $2$, $20$ and 200 nm.

4.3. Calculation of $k_{e}$ and $D_{r}$

Using a boundary integral method which will be described shortly, we have calculated the effective aspect ratio $k_{e}$ and the rotational diffusion coefficient $D_{r}$ for typical nanoplatelets by solving the incompressible Stokes equations

(4.1a,b)\begin{equation} \boldsymbol{\nabla} \boldsymbol{\cdot} \boldsymbol{\sigma}=0,\quad \boldsymbol{\nabla}\boldsymbol{\cdot} \boldsymbol{u}=0, \end{equation}

where $\sigma _{ij}=-\delta _{ij}p + \eta (\partial u_i/\partial x_j +\partial u_j/\partial x_i)$ is the hydrodynamic stress with $p$ the pressure and $\eta$ the viscosity; $\boldsymbol {u}$ is the velocity field. For the reference surface, we use a cuboid with rounded edges (§ 4.1). At the reference surface of the platelet we prescribed the Navier slip boundary condition

(4.2)\begin{equation} \boldsymbol{u}_{{sl}}= \frac{\lambda}{\eta} \boldsymbol{n} \times \boldsymbol{f}\times\boldsymbol{n}, \end{equation}

where $\boldsymbol {f}=\boldsymbol {\sigma }\boldsymbol {\cdot } \boldsymbol {n}$ is the hydrodynamic surface traction and $\boldsymbol {n}$ is the unit normal vector (Lauga, Brenner & Stone Reference Lauga, Brenner and Stone2008). In the present case, the macroscopic time scale is given by the rotational time period $P_{R} = \dot \gamma ^{-1} ( |k_{e}|^{-1} + |k_{e}|)$, which is controlled by the particle aspect ratio and by the shear rate. For realistic parameters, $P_{R}$ is always much larger than the time scale of the microscopic motion of the fluid molecules, which is typically of the order of picoseconds. A boundary integral equation is used to calculate the distribution of $\boldsymbol {f}$ corresponding to each given inclination angle $\phi$ (Pozrikidis Reference Pozrikidis1992). Denoting the reference surface by $S$, the boundary integral equation for a point $\boldsymbol {x}\in S$ is (Luo & Pozrikidis Reference Luo and Pozrikidis2008)

(4.3)\begin{equation} \int_{S} \boldsymbol{n} \boldsymbol{\cdot}\boldsymbol{K}(x',y') \boldsymbol{\cdot} \boldsymbol{u}_{{sl}} \,\mathrm{d} S -\frac{1}{\eta}\int_{S} \boldsymbol{G}(x',y') \boldsymbol{\cdot} \boldsymbol{f}\, \mathrm{d} S = \frac{\boldsymbol{u}_{{sl}}(\boldsymbol{x})}{2}-\boldsymbol{u}_{\infty}(\boldsymbol{x}), \end{equation}

where the integral is over the surface parameterised by the coordinate point $\boldsymbol {x}_1=(x_1,y_1)$. In (4.3), $x'=x_1-x$, $y'=y_1-y$; $\boldsymbol {G}$ and $\boldsymbol {K}$ are Green's functions corresponding to the 2-D ‘Stokeslet’ and ‘Stresslet’, respectively; $\boldsymbol {u}_{\infty }$ is the undisturbed simple shear flow field. The traction is used to calculate the hydrodynamic torque $T(\phi )= \boldsymbol {\hat {e}}_{z}\boldsymbol {\cdot } \int [{\boldsymbol {x}}\times \boldsymbol {f}]\,\mathrm {d}S$ acting on the fixed platelet. In order to achieve high accuracy with the boundary integral algorithm, a non-uniform grid has been implemented with the surface of the particle being divided into two regions, see figure 10(a). Namely, the edge region, made of the edges of arclength $S_{E}$ plus the neighbouring regions of lengths $S_{E}/4$, and the planar region. The total number of grid points, $N$, is distributed between the two regions, with the number of points of the edge region being $N/2$ up to a maximum of $32$ points for each end. Following a convergence analysis (figure 10b,c), we set $N=288$ for platelets with $a/b\leq 100$ and $N=384$ for platelets with $a/b> 100$; the value of $k_{e}$ is then calculated from (2.4).

Figure 10. (a) Example of the discretisation of a $n=1$ surface with $N=96$. (b,c) Effective aspect ratio $|k_e|$ versus the number of discretisation points $N$ for platelets with $\lambda =20\ \textrm {nm}$, $a/b=10$ or $a/b=100$, and either $n=1$ (b) or $n=10$ (c).

The (4.3) is evaluated numerically by using the following method. First, the traction and slip velocity are discretised as $N$ piecewise constant functions $\{\boldsymbol{f}[1]\ldots \boldsymbol {f}[N]\}$ and $\{\boldsymbol {u}^{{sl}}[1]\cdots \boldsymbol {u}^{{sl}}[N]\}$, respectively, each associated with a surface element $S_i=\int _{s_i}^{s_{i+1}}\,\mathrm {d}S$. With this discretisation, for each point $s_j$ for $j =\{1 \cdots N\}$, the discretised form of (4.3) becomes (Kamal et al. Reference Kamal, Gravelle and Botto2020)

(4.4)\begin{align} \sum_{i=1}^{N}\left[\boldsymbol{u}_{{sl}}[i]\boldsymbol{\cdot}\int_{s_i}^{s_{i+1}} \boldsymbol{K}(s,s_j)\boldsymbol{\cdot} \boldsymbol{n}(s)\,\mathrm{d} S(s) - \frac{\boldsymbol{f}[i]}{\eta}\boldsymbol{\cdot} \int_{s_i}^{s_{i+1}} \boldsymbol{G}(s,s_j)\,\mathrm{d} S(s) \right] =\frac{\boldsymbol{u}_{sl}(s_j)}{2} -\boldsymbol{u}_{\infty}(s_j). \end{align}

For $i\neq j$, each subintegral is evaluated using a Gauss–Legendre quadrature method. If $i=j$, the integrands are singular. The singular integrand containing the $\boldsymbol {G}$ tensor is evaluated by using a specific quadrature method for logarithmic singularities (Pozrikidis Reference Pozrikidis2002), and the singular integrand containing the $\boldsymbol {K}$ tensor is evaluated analytically by Taylor expansion about the singular point (Kamal et al. Reference Kamal, Gravelle and Botto2020). Equation (4.2) provides a closed relation between $\boldsymbol {u}_{{sl}}$ and $\boldsymbol {f}$, and thus (4.4) can be arranged into a closed system of $N$ linear equations for each component of the traction $f_{i}[j]$. This system is solved for $f_{i}[j]$ by using Gaussian elimination. To test the accuracy of our implementation, we compared the predicted value of $k_{e}$ with the corresponding exact solution for circular cylinder in shear flow (Kamal et al. Reference Kamal, Gravelle and Botto2020). This test confirms the expected second-order spatial convergence.

The rotational diffusion coefficient of a particle is given by $D_{r}=k_{B} T/ F_r$, where $k_{B}$ is the Boltzmann's constant, $T$ is the absolute temperature and $F_r$ is the rotational drag coefficient. The rotational drag coefficient is also calculated via the boundary integral method, by calculating the hydrodynamic torque exerted on the fluid by a platelet rotating with velocity $\boldsymbol {u}= \boldsymbol {\hat {e}}_{z} \times \boldsymbol {x}$ in a fluid otherwise at rest. In agreement with the result of Sherwood (Reference Sherwood2012) for a disk, we find for our model platelet that $F_r$ is independent of slip as $k \to 0$. In this limit we found $F_r \approx c_1 \eta a^3$, where $c_1 = 6.29 \pm 0.02$ (to be compared with $c_1 = 32/3$ for a 3-D disk of zero thickness (Leal & Hinch Reference Leal and Hinch1971; Sherwood Reference Sherwood2012), and to $c_1 = 2 {\rm \pi}$ for an infinite plate of zero thickness (Sherwood & Meeten Reference Sherwood and Meeten1991)). The second-order accuracy in the computation of $D_{r}$ was assessed by comparing against the exact solution for $D_{r}$ for a cylinder with a Navier slip boundary condition, as done by Luo & Pozrikidis (Reference Luo and Pozrikidis2008) for a sphere.

Values of $k_{e}$ for our model platelet with $\xi =0.25\ \textrm {nm}$, slip length $\lambda = 0$, $2$, $20$ and $200\ \textrm {nm}$, and $n = 1$ or $10$ are shown in figure 11. We limit our calculations to aspect ratios $k \ge 0.001$, because using $k<0.001$ leads to large numerical error owing to severe resolution requirements. For $\lambda =0$, the effective ratio of the platelet follows the power law relationship $k_{e} \sim g_{n} k^{3/4}$, with $g_{n}$ a prefactor that depends on $n$. This power law relationship is the same as for disks with ‘blunt edges’ (Singh, Koch & Stroock Reference Singh, Koch and Stroock2013). We find $g_{1} \approx 0.91$ for $n=1$ and $g_{10} \approx 1.0$ for $n=10$. As $\lambda$ increases, however, the relation between $k_{e}$ and $k$ is not necessary a power law, as shown for $\lambda =2,\;20$ and 200 nm for $n=1$, and $\lambda =20$ and 200 nm for $n=10$. In the limit $\lambda /a \gg 1$, the value of $k_{e}$ can be analytically approximated as $k_{e} \propto \text {i} \sqrt {k}$ (Kamal et al. Reference Kamal, Gravelle and Botto2020), as confirmed in figure 11 for platelet length $2 a \lessapprox 20\ \textrm {nm}$, $n=1$ and $\lambda =20\ \textrm {nm}$ and $200\ \textrm {nm}$ (we remark, however, that most applications of nanoplatelets typically involve $\lambda < a$).

Figure 11. Absolute value of the effective aspect ratio $k_{e}$ versus the platelet length $2 a$ for $\xi =0.25\ \textrm {nm}$ and for $n=1$ (a) and $n=10$ (b). The black lines correspond to the purely imaginary values of $k_{e}$ and the red lines to real values of $k_{e}$.

As the slip length increases, the value of $k_{e}^2$ decreases and changes sign at some critical value $\lambda _c$ (figure 12). An explanation for this is the following. Calculating $k_{e}$ requires the evaluation of the total hydrodynamic torque on a fixed platelet for two angles: $\phi =0$ and $\phi ={\rm \pi} /2$. The total torque for $\phi =0$ is more sensitive to $\lambda$ than the torque for $\phi ={\rm \pi} /2$, because slip affects primarily the flow moving in the direction tangential to the surface of the particle. Thus, predicting $\lambda _c$ essentially requires predicting the angle for which $T(0)$ changes sign. When $\lambda =0$, in an asymptotic expansion in powers of $b$, the $O(b)$ contribution to $T(0)$ turns out to be zero, because of an exact cancellation between torque contributions due to normal and tangential stresses. This cancellation results in a ‘clockwise’ torque that is second-order in the thickness, $T\propto b^2$; the second-order tangential traction determining this torque is $f_x^{(2)} \propto b$ (Singh et al. Reference Singh, Koch, Subramanian and Stroock2014). The effect of slip over the flat surface of the platelet is to reduce the tangential traction $f_x$, without changing significantly the normal traction $f_y$ (figure 13). An analysis for $\lambda /a \ll 1$ shows that the effect of slip is to reduce $f_x$ by an amount of $O(\lambda )$, resulting in a second-order traction $f_x^{(2,{sl})} \propto \lambda$ (Kamal et al. Reference Kamal, Gravelle and Botto2020), where ‘$sl$’ indicates that this traction is due to slip. The total torque $T(0)$ changes sign when $f_x^{(2,{sl})} > f_x^{(2)}$. This condition occurs at a threshold value $\lambda _c \approx h_n b$, where $h_n$ is a numerical prefactor that depends on the number of layers $n$ (Kamal et al. Reference Kamal, Gravelle and Botto2020). For $n=1$, our numerical procedure gives $h_1 \approx 0.8$, and $b \approx 0.25\ \textrm {nm}$ so $\lambda _c \approx 0.2\ \textrm {nm}$. For $n=10$, $h_{10}\approx 2.2$ and $b \approx 1.75\ \textrm {nm}$ so $\lambda _c \approx 3.85\ \textrm {nm}$ (figure 12). For moderate-to-large values of $\lambda /b$, slip can have a marked effect on the traction at and near the edges of the platelet (the position of the edge is marked by the dotted vertical line in figure 13). This traction can significantly affect $k_e$ for larger values of $\lambda$ so that $k_e$ is in general a non-trivial function of $\lambda$. Comparison of the traction and $T(0)$ with MD simulations reveals that the approximation of a uniform surface slip over the surface of the platelet still gives accurate predictions of $k_e$ even when $\lambda /b$ is large (Kamal et al. Reference Kamal, Gravelle and Botto2020). In the following, since $k_e$ is a non-trivial function of $\lambda$, $b$ and $a$ (figure 12), we will use the numerically computed values of $k_{e}$.

Figure 12. Effective aspect ratio $k_{e}^2$ versus the slip length $\lambda$ for different platelet lengths $2a$. The half-thickness of the platelet is $b\approx 0.25\ \textrm {nm}$ and $b\approx 1.75\ \textrm {nm}$ for $n=1$ and $n=10$, respectively. Red dashed lines mark the positions where $k_{e}^2$ changes sign.

Figure 13. The streamwise (a) and normal (b) hydrodynamic traction distribution along the surface of a platelet fixed at an orientation $\phi =0$ for $n=1$ and $a/b=10$. The vertical dotted line marks the position of the edge of the platelet.

4.4. Regimes of rotation

Here we analyse the effect of the Péclet number, comparing the rotational dynamics of nanoplatelets with and without surface slip. We first compare the cases $\lambda =0$ (no-slip) and $\lambda =2\ \textrm {nm}$ (slip) for fixed $a/b=100$ and $n=1$. The corresponding effective aspect ratios are $k_{e}=0.028$ and $k_{e}=0.063$i for $\lambda =0$ and $\lambda =2\ \textrm {nm}$, respectively. In figure 14, we plot the mean $\mu$ and variance $\sigma ^2$ of the rotation angle for these two cases as a function of $Pe$. In agreement with figure 7, the results show that slip has no effect on particle orientation at low $Pe$, but has a strong influence on both $\mu$ and $\sigma ^2$ at large $Pe$. To describe the qualitative differences in the trends of $\mu$ and $\sigma ^2$ between the slip and no-slip cases, we subdivide the data into three different regions (I, II and III, see figure 14). Region III corresponds to ${Pe} > {Pe}_{c}$, where $Pe_{c}$ is the critical value introduced in § 3.3 above which no full rotational cycles occur for $\lambda =2\ \textrm {nm}$. Region III is a ‘large slip effects’ regime. In this regime the platelet with slip performs small fluctuations around $\phi _{c}$. In this region the curves approach smoothly the values expected for ${Pe} \to \infty$ (§ 3.2): for $\lambda =2\ \textrm {nm}$, $\mu$ is approximately constant ($\mu \approx |k_{e}|$), and $\sigma ^2 \to 1/(2 {Pe} |k_{e}|)$ as ${Pe} \to \infty$. This latter asymptotic solution for $\sigma ^2$ can be obtained by inserting into (2.1) the linearisation of the angular velocity for angles close to $\phi _{c}$, $\varOmega =2 k_{e} (\phi -\phi _{c})$ and assuming that $p$ is Gaussian. For $\lambda =0$, $\mu$ vanishes as ${Pe} \to \infty$ and $\sigma ^2$ is finite, as predicted by Jeffery's theory. Regions I and II correspond to ${Pe}<{Pe}_{c}$. In region I, slip and no-slip curves for $\mu$ or $\sigma ^2$ are practically indistinguishable, as apparently in this region the effect of the hydrodynamic stresses is subdominant with respect to Brownian stresses in setting the orientational dynamics. In region II, differences in the curves due to slip exists, but are not as marked as in region III. Region II corresponds to a ‘moderate slip effects’ regime, in which full rotational cycles occur even for platelets with an imaginary $k_{e}$, but with a significantly reduced frequency as compared with region I. This reduction in frequency indicates that in this region hydrodynamics mitigates the randomising effect of Brownian motion.

Figure 14. Comparison of mean and variance of the fluctuating rotation angle for a nanoplatelet with slip ($k_{e}=0.063\text {i}$) and without slip ($k_{e}=0.028$). The red dashed line marks the asymptotic solution for $\sigma ^2$ for purely imaginary values of $k_{e}$. The nanoplatelet has half-width $b=0.25\ \textrm {nm}$ (i.e. $n=1$) and half-length $a=25\ \textrm {nm}$. Region I: ${Pe} < {Pe}_{I}$, negligible slip effects for both $\lambda = 0$ and $\lambda = 2$  nm. Region II: ${Pe}_{I} \leq {Pe}<{Pe}_{c}$, moderate slip effects region characterised by changes in the frequency of rotation due to slip. Region III: ${Pe}\geq {Pe}_{c}$, large slip effects region where slip completely suppresses the rotation of the nanoplatelet.

In regions II and III the effect of slip is to increase the mean and reduce the variance of the orientation distribution. The effect of slip on the variance is particularly evident: for ${Pe} = 10\,000$, a small slip of $\lambda = 2\ \textrm {nm}$ leads to a reduction by approximately one order of magnitude of the variance.

As explained in § 3.2, the value ${Pe}_{I}$ separating the ‘negligible slip effects’ regime and the ‘moderate slip effects’ regime must satisfy the condition ${Pe}_{I}\ll {Pe}_{m}=\min (|k_{e}^{-3}(\lambda =0)|,k_{e}^{-3}(\lambda )|)$. Our numerical simulations (shown in both figures 7 and 14) indicate that the correlation ${Pe}_{I}\simeq 0.1 {Pe}_{m}$ provides a good approximation to the data for practical purposes.

In figure 15, we extend the analyses of figure 14 to assess the extent of the regions of large slip effects, moderate slip effects and negligible slip effects by evaluating ${Pe}_{c}$ and ${Pe}_{I}$ in the ($Pe$, $2a$) space. For our comparison, we use the same platelets and corresponding values of $k_{e}$ used in figure 11. The difference to the analysis presented in § 3 is that now ${Pe}_{c}$ and ${Pe}_{m}$ can be evaluated as a function of the known variables $\lambda$ and $2a$, rather than $k_{e}$.

Figure 15. Effect of slip on the rotational dynamics of a nanoplatelet with $\text {length} = \text {depth} = 2a$ and $\text {width} = 2b$ in the ($Pe$, $2a$) space. Upper black curve marks the threshold Péclet number ${Pe}_{c}$ for which $\sigma ^2=2 k_{e}$; above this threshold the hydrodynamic slip has a dramatic effect on the particle dynamics. Lower black curve marks the threshold Péclet number ${Pe}_{I}$; above this threshold, slip effects become important. The dashed, dashed–dot and dash–double-dot lines correspond to $\dot {\gamma }\eta = 10^7\ \text {Pa}$, $\dot {\gamma }\eta = 2\times 10^4\ \text {Pa}$, $\dot {\gamma }\eta = 10\ \text {Pa}$, respectively.

We add to figure 15 red dash–dot lines corresponding to the $Pe$ number for three shear stress values: $\dot {\gamma }\eta = 10^7\ \textrm {Pa}$, $\dot {\gamma }\eta = 2\times 10^4\ \textrm {Pa}$, $\dot {\gamma }\eta = 10\ \textrm {Pa}$ (the Péclet number scales proportionally to the stress according to ${Pe}\approx c_1 \dot {\gamma }\eta a^3/(k_B T)$, where ‘$c_1 = 6.29 \pm 0.02$’ was calculated from a boundary integral simulation of the hydrodynamic torque on a particle rotating with assigned velocity in a still fluid, see § 4.3). These characteristic shear stresses are, respectively, typical of: exfoliation processes in a low viscosity fluid (Paton et al. Reference Paton2014); mixing in very viscous fluids (Huang & Terentjev Reference Huang and Terentjev2012) and dispersion by microfluidisation (Karagiannidis et al. Reference Karagiannidis2017; Paton et al. Reference Paton, Anderson, Pollard and Sainsbury2017); fast lubrication processes (Jonsson & Bhushan Reference Jonsson and Bhushan1995), high-speed blade coating (Willenbacher, Hanciogullari & Wagner Reference Willenbacher, Hanciogullari and Wagner1997) and MD simulations (Voeltzel et al. Reference Voeltzel, Fillot, Vergne and Joly2018; Gravelle et al. Reference Gravelle, Kamal and Botto2020; Gravelle, Kamal & Botto Reference Gravelle, Kamal and Botto2021). The effects of slip for each of these characteristic shear stresses can be assessed by identifying, for a given value of $a$, the region where the corresponding $Pe$ lies. For example, the lowest shear stress (red dashed line in figure 15) is within the region where slip effects are negligible for all the values of $\lambda$, $a$ and $n$ we considered. For the intermediate shear stress (dotted–dashed line in figure 15) slip has a noticeable impact on the rotational dynamics: thicker platelets ($n=10$) are likely to align if the slip length is sufficiently large so that $k_{e}$ is purely imaginary. Otherwise for the thinner platelets, such as for $n=1$, then there is a range of $a$ and $\lambda$ for which the average frequency for the platelets to complete a full rotational cycle is smaller than for a no-slip platelet. Finally, the highest shear stress (double-dotted–dashed line in figure 15) produces a value of $Pe$ well within the ‘large slip effects’ region where the platelet fluctuates around $\phi _{c}$ if $\lambda > \lambda _c$.

The length of the platelet also affects the rotational dynamics. The local power law exponent characterising the dependence of ${Pe}_{c}$ and ${Pe}_{I}$ on $a$ is smaller than $3$ for comparatively small values of $a$ ($a \ll 100\ \textrm {nm}$ approximately in figure 15), and approaches $3$ as $a$ increases. A consequence of the fact that the ‘moderate slip effects region’ is not bounded by parallel lines in a log–log plot is that for given particle thickness and shear stress value longer platelets are more likely to be affected by slip. Taking the case $n=1$, $\lambda =200\ \textrm {nm}$, and $\eta \dot {\gamma } = 2\times 10^4\ \textrm {Pa}$ as an example, the rotation of the platelet is not affected by slip for $2a<40\ \textrm {nm}$, whereas platelets with $2a>40\ \textrm {nm}$ enter the ‘moderate slip effects region’ where slip effects start becoming significant.

4.5. Effective viscosity of a dilute suspension of nanoplatelets with slip

As an illustration of a macroscopic effect due to slip, in the current section we analyse the effects of stable alignment on the orientational contribution to the effective viscosity $\eta _{{eff}}$ of a dilute suspension of nanoplatelets for finite values of $\lambda$. The effective shear viscosity of a dilute suspension of particles can be calculated as

(4.5)\begin{equation} \eta_{{eff}}=\eta+\frac{\sigma'_{xy}}{\dot{\gamma}}, \end{equation}

where $\sigma '_{xy}$ is the particle stress (Jeffery Reference Jeffery1922; Giesekus Reference Giesekus1962; Brenner Reference Brenner1974; Leal & Hinch Reference Leal and Hinch1971). For a suspension in which the motion of the particles is restricted to the flow plane, to first order in the particle concentration, the particle stress can be written as

(4.6)\begin{equation} \sigma'_{xy}=c\eta \dot{\gamma}\left(\frac{A}{2}\left\langle1-\cos{4\phi}\right\rangle+B +\frac{C}{2 {Pe}} \left\langle \sin{2\phi} \right\rangle \right), \end{equation}

where $A,\;B$ and $C$ are dimensionless coefficients (Rallison Reference Rallison1978) and $c$ is the solid fraction. The first two terms on the right-hand side of (4.6) are due to the hydrodynamic stress acting on each platelet. The third term is the contribution of Brownian fluctuations to the particle stress.

As seen in § 3.2, slip has a strong influence on the time-averaged orientation of the particle, with consequences on the statistical quantities $\left \langle 1- \cos {4\phi } \right \rangle$ and $\left \langle \sin {2\phi } \right \rangle$ in (4.6). A map quantifying how $\left \langle 1- \cos {4\phi } \right \rangle$ and $\left \langle \sin {2\phi } \right \rangle$ vary in the ($Pe$, $k_{e}$) space is shown in figure 16. For $Pe$ $\ll |k_{e}|^{-3}$, the quantities $\left \langle 1- \cos {4\phi } \right \rangle$ and $\left \langle \sin {2\phi } \right \rangle$ are practically independent of whether $k_{e}$ is real or purely imaginary. The term $\left \langle 1- \cos {4\phi } \right \rangle$ becomes significantly larger than $\left \langle \sin {2\phi } \right \rangle$ for small $Pe$. Thus, the effective viscosity depends increasingly on $\left \langle 1- \cos {4\phi } \right \rangle$ as $Pe$ decreases, in agreement with the theoretical results of Hinch & Leal (Reference Hinch and Leal1972) and Leahy et al. (Reference Leahy, Koch and Cohen2015) (for elongated particles with $\lambda =0$) and the experimental results of Del Giudice & Shen (Reference Del Giudice and Shen2017) for graphene oxide.

Figure 16. Colour maps of $\left \langle 1-\cos {(4\phi )}\right \rangle$ and $\left \langle \sin {(2\phi )}\right \rangle$ in the ($Pe$, $k_{e}$) space for (a,c) an imaginary value of $k_{e}$ and (b,d) a real value of $k_{e}$. The black line marks the equation ${Pe}=|k_{e}|^{-3}$.

For $Pe \gg |k_{e}|^{-3}$, hydrodynamic stresses become dominant. Figure 16 indeed reveals a strong dependency of $\left \langle 1- \cos {4\phi } \right \rangle$ and $\left \langle \sin {2\phi } \right \rangle$ on $k_{e}$ in this region. Since the prefactor of $\left \langle \sin {2\phi } \right \rangle$ tends to zero for large $Pe$, here we focus on analysing the term $\left \langle 1- \cos {4\phi } \right \rangle$. The leading-order values of $\left \langle 1- \cos {4\phi } \right \rangle$ in this region can be estimated directly by using the probability distributions calculated for $Pe$ $\to \infty$ in § 3.2:

(4.7) \begin{equation} \left\langle1-\cos{4\phi}\right\rangle= \left\{ \begin{array}{@{}ll} 4 k_e/(k_e +1)^2 & \text{if }k_{e} \in \mathbb{R}, \\ 1-\cos{4\phi_{c}} & \text{if }k_{e} \in i\mathbb{R}. \end{array} \right. \end{equation}

For $|k_{e}| \to 0$, $\left \langle 1-\cos {4\phi }\right \rangle$ tends to $4 k_e$ and $8 |k_{e}|^2$ for real or purely imaginary $k_{e}$, respectively. For $\lambda /a = 0$ and $\lambda /a \to \infty$, $\left \langle 1-\cos {4\phi }\right \rangle$ can be estimated in terms of the geometric aspect ratio $k$. For $\lambda =0$, $k_{e} \propto k$, and so $\left \langle 1-\cos {4\phi }\right \rangle \propto k$. For $\lambda /a \to \infty$ we have $k_{e} \propto \text {i} \sqrt {k}$, giving again $\left \langle 1-\cos {4\phi }\right \rangle \propto k$. Thus, quite interestingly, the limits $\lambda /a \to 0$ and $\lambda /a \to \infty$ yield a similar scaling relationship of $\left \langle 1-\cos {4\phi }\right \rangle$ with $k$.

For $\lambda =\lambda _c$, we have $k_{e}=0$ and therefore $\left \langle 1-\cos {4\phi }\right \rangle =0$. When $\left \langle 1-\cos {4\phi }\right \rangle$ vanishes, the particle stress reduces to $\sigma '_{xy}=c\eta \dot {\gamma }B$. If the variation of $B$ on $\lambda$ is sufficiently small in the neighbourhood of $\lambda _c$ (compared with $A\left \langle 1-\cos {4\phi }\right \rangle$), then the value $\lambda =\lambda _c$ will correspond to a local minimum in the $\sigma '_{xy}$ versus $\lambda$ curve, because $A\left \langle 1-\cos {4\phi }\right \rangle \ge 0$ for all values of $\lambda$. Thus, the decrease in the value of $\sigma '_{xy}$ with $\lambda$ may not be always monotonic for large $Pe$.

5. Discussion

We have shown that the classical Bretherton's equation of motion can describe the rotational dynamics of thin rigid particles with surface slip, as long as an effective aspect ratio $k_{e}$ is introduced to account for the hydrodynamic slip length $\lambda$ in addition to the particle's shape. For a fixed geometric aspect ratio $b/a$ and a small slip length $\lambda <b$ (with $b$ the platelet's half-thickness), $k_{e}$ is real, and an increase of $\lambda$ simply leads to a decrease of $k_{e}$. However, when $\lambda$ is sufficiently large in comparison with $b$, $k_{e}$ becomes a purely imaginary number. In the limit of infinite Péclet numbers, a platelet with an imaginary $k_{e}$ does not perform the periodic orbits predicted by the classical theory of Jeffery. Instead, the platelet's orientation fluctuates in time around a small angle $\phi _{c}$ with respect to the flow direction. We have identified a critical Péclet number ${Pe}_{c}$ above which such stabilisation occurs, and below which rotations due to Brownian fluctuations appear. The numerical results suggest ${Pe}_{c} \sim 0.45 |k_{e}|^{-3.4}$ for $|k_{e}| \ll 0.1$, and ${Pe}_{c} \sim 0.02 |k_{e}|^{-4.8}$ for $|k_{e}|> 0.1$, with a smooth transition between the two scalings. An alternative critical Péclet number ${Pe}_{m}$, based on balancing the orders of magnitude of the convective and diffusive terms in (2.1), predicts the hydrodynamic stresses acting on the platelet to be dominant with respect to the Brownian stresses when ${Pe}_{m} = \min (|k_{e}^{-3}(\lambda =0)|,k_{e}^{-3}(\lambda )|)$.

We find a negligible difference between the rotational behaviour of slip and no-slip platelets if ${Pe}$ is smaller than ${Pe}_{I} \approx 0.1{Pe}_{{m}}$. In this regime, Brownian stresses dominate over hydrodynamic stresses in the regime where slip effects are important. For $Pe$ smaller than ${Pe}_{c}$, Brownian fluctuations force the platelet with slip to complete full rotations, but with a period larger than the one expected in the no-slip case. Finally, we examined the effect of stable alignment on the orientational contribution to the effective viscosity by analysing how the term $\left \langle 1- \cos {4\phi } \right \rangle$ changes due to slip. At large $Pe$, we found that this term does not decrease monotonically with $\lambda$ but instead attains a minimum value at $\lambda =\lambda _c$.

Some assumptions in our mathematical model require a discussion in view of the results. One aspect regards the rigidity of the particle, as thin particles are intuitively expected to be flexible. Assuming that the bending rigidity follows $B \sim E b^3$ with $E$ the Young's modulus (Poot & van der Zant Reference Poot and van der Zant2008), where $E \sim 10^{11}\ \textrm {Pa}$ for pure graphene (Lindahl et al. Reference Lindahl, Midtvedt, Svensson, Nerushev, Lindvall, Isacsson and Campbell2012), for platelets aligned almost parallel to the direction of the shear flow, buckling should occur when $\eta \dot {\gamma } (a/b)^2 /E \ge 1$ (Lingard & Whitmore Reference Lingard and Whitmore1974; Kamal et al. Reference Kamal, Gravelle and Botto2020). Using the typical shear stress $\dot {\gamma }\eta \sim 10^4\ \textrm {Pa}$ for which slip effects become important (§ 4.4), one finds that the platelet behaves as rigid as long as the geometric aspect ratio $k \ge 10^{-7/2}\simeq 0.0003$. Therefore, the assumption of rigidity in our work may not be very restrictive for particles whose aspect ratio is not too extreme. Our results have been obtained for a 2-D geometry, which strictly speaking corresponds to a nanoplatelet with an infinite extent in the direction of the vorticity vector. Real nanoplatelets have, of course, a finite lateral size, with a shape that can vary considerably, from irregular ‘pentagon’ plate-shape (Del Giudice et al. Reference Del Giudice, Cunning, Ruoff and Shen2018) to 2-D ‘ribbon’-like (Hao et al. Reference Hao, Qian, Zhang and Hou2008)). However, it can be shown by comparing the hydrodynamic torque exerted on a 2-D plate-like particle with that on a 3-D disk that these different geometries lead to values of the effective aspect ratio $k_{e}$ that are very similar to each other (Kamal et al. Reference Kamal, Gravelle and Botto2020). Therefore, the rotational dynamics is expected to be qualitatively similar for different plate-like objects, as long as the platelet's extent in the vorticity direction is not too small. Finally, a 3-D object also possesses extra degrees of freedom, i.e. its rotation is not confined to a plane. However, as discussed by Leahy et al. (Reference Leahy, Koch and Cohen2015) and Hinch & Leal (Reference Hinch and Leal1972), the qualitative features of the rotational dynamics of a particle whose motion is confined to a plane are expected to be similar to those of a particle whose motion is in the full 3-D space. Therefore, we expect our model to have relevance to realistic systems. Additional insights would require the solution of full 3-D trajectories. MD results we have carried out with plate-like slip molecules rotating in the full 3-D space reveal that the 2-D approximation used here give rotational statistics that are comparable in trend and magnitude – with reasonably good approximation – to the ones computed from MD, giving us confidence in the value of the 2-D approximation employed here (Gravelle, Kamal & Botto Reference Gravelle, Kamal and Botto2021).

To illustrate how our theoretical results could be applied in practice, in § 4 we have considered geometric and slip parameters relevant to graphene and other 2-D nanomaterials. By calculating the values of ${Pe}_{m}$ and ${Pe}_{c}$ for typical values of $k$, $b$ and $\lambda$, we have found that stresses $\dot {\gamma }\eta$ of the order of $10^{4}\ \textrm {Pa}$ are required for hydrodynamic slip to affect quantitatively the rotational dynamics.

The use of fluid more viscous than water would allow for such large shear stress values. For instance, some ionic liquids have a viscosity approximately three orders of magnitude larger than that of water, while also showing slip length as large as 100 nm on graphene (Voeltzel et al. Reference Voeltzel, Fillot, Vergne and Joly2018). The use of very viscous fluids would have the further benefit of reducing the possibility of hydrodynamic instabilities and turbulences. Large viscous stresses may be accompanied by heating, but one could account for this effect by evaluating the viscosity and the slip length at the corresponding temperature.

An experimental verification of our results is not without challenges. For example, surface chemical modification of graphene nanoparticles is generally needed to obtain good dispersion of graphene in water, and when functionalisation is applied to the planar region of the platelet, the slip length is expected to be reduced (Wei, Peng & Xu Reference Wei, Peng and Xu2014). Particles for which the functionalisation is limited to the particle edges may not have this problem (Park et al. Reference Park, Kim, Sung, Kim and Park2017; Aliyeva et al. Reference Aliyeva, Alosmanov, Buniyatzadeh, Azizov and Maharramov2019), and could be more suitable to confirm the theory. Detecting rotation may not be trivial. For example, techniques that measure the average orientation of the particles may not enable one to clearly distinguish between a suspension in which very slender particles rotate but are oriented in a time-average sense, from a suspension in which the particle orientation angle is constant, since in both cases $\phi$ is expected to be small. Because the occurrence of a stable orientation depends only on the ratio $\lambda /b$, and not on the particle length, a solution could be to use plate-like particles that are thin ($b \sim 1\ \textrm {nm}$) but not too long. For a given shear stress the ideal particle to evidence slip effects would be the shortest particle whose value of $a$ is in the ‘moderate slip effects’ region, since this choice would give the largest variation in $k_{e}$ due to slip, and thus the largest variation in $\phi$. With this choice of particle length, the different rotational behaviours of slip and no-slip particles should be particularly evident.

In experiments, different statistical measures of particle orientation could be adopted, and some of them may be more sensitive to slip than others. To illustrate this dependence and the impact slip might have on orientational statistics that are evaluated in practice, we discuss two quantities that are usually measured in rheo-optical studies (Frattini & Fuller Reference Frattini and Fuller1986; Fuller Reference Fuller1995; Vermant et al. Reference Vermant, Yang and Fuller2001; Reddy et al. Reference Reddy, Perez-Juste, Pastoriza-Santos, Lang, Dhont, Liz-Marzan and Vermant2011): the average orientation angle $\chi$ and the degree of alignment $\Delta n'' /(\Delta n''_{max})$, defined as

(5.1a,b)\begin{equation} \chi=\frac{1}{2}\arctan\left(\frac{\left\langle\sin^2{\theta}\sin{2\phi}\right\rangle}{\left\langle\sin^2{\theta}\cos{2\phi}\right\rangle}\right), \quad \frac{\Delta n'' }{\Delta n''_{max}}=\sqrt{\left\langle\sin^2{\theta}\sin{2\phi}\right\rangle^2+\left\langle\sin^2{\theta}\cos{2\phi}\right\rangle^2}. \end{equation}

Figure 17 compares our theoretical prediction for $\chi$ and $\Delta n'' /(\Delta n''_{max})$ (in the case $\theta = {\rm \pi}/2$) to the experimental data of Reddy et al. (Reference Reddy, Natale, Prud'homme and Vermant2018). The experiments are for dilute polydisperse suspensions of functionalised graphene platelets, with average half-length $a=250\ \textrm {nm}$ and half-thickness $b(n=5)=0.86\ \textrm {nm}$, and slightly oblate gold nanospheroids, with $b/a=0.55$ and lengths of either $170\ \textrm {nm}$ or $290\ \textrm {nm}$. The functionalised graphene sheets were dispersed in mineral oil (viscosity $\eta = 100\ \textrm {mPa}\,\textrm {s}$) and the oblate gold spheroids were dispersed in a glycerol/water mixture at 99.5 % glycerol ($\eta = 1.2\ \textrm {Pa}\,\textrm {s}$). The theoretical values (lines) are for particles having the same average aspect ratio as in the experiments, and considering the limiting cases of $\lambda =0$ and $\lambda =20\ \textrm {nm}$ (figure 17). For large $Pe$, the functionalised graphene platelets where found experimentally to converge to a finite angle $\chi = 5^{\circ }$, instead of approaching $0$ (as one would expect for a platelet with a no-slip boundary in the limit ${Pe} \to \infty$). The plot shows that our theory raises the $\chi -{Pe}$ curve closer to $\chi = 5^{\circ }$ at large $Pe$, although probably a value $\lambda =20\ \textrm {nm}$ is larger than the one occurring in the experiment. Importantly, the plots show that for the slightly oblate spheroids, slight changes in $k_{e}$ due to slip makes only a slight change to $\chi$ (figure 17), but a more marked change in $\Delta n'' /(\Delta n''_{max})$, suggesting that the second quantity is a more sensitive measure of slip effects.

Figure 17. Comparison of our theory (for $\theta ={\rm \pi} /2$) with experimental results of Reddy et al. (Reference Reddy, Natale, Prud'homme and Vermant2018) for $\chi$ and $\Delta n'' /\Delta n''_{max}$ (symbols), as defined in (5.1a,b). The different lines correspond to the following model parameters: functionalised graphene plateletes (FGS) with $k_{e}\approx 0.016$ (no-slip) and $k_{e}=0.035\text {i}$ (slip with $\lambda =20\ \text {nm}$); slightly oblate nanospheroids (NP) with $b/a=0.55$, for $k_{e} \approx 0.55$ (no-slip) and $k_{e}\approx 0.46$ (slip with $\lambda /b=0.21$).

We suggest that experiments similar to those of Reddy et al. (Reference Reddy, Natale, Prud'homme and Vermant2018) could be done in which both $\chi$ and $\Delta n'' /(\Delta n''_{max})$ are compared for a range of slip lengths and values of particle thickness. Rather than plotting $\chi$ or $\Delta n'' /(\Delta n''_{max})$, one could also plot $\left \langle \sin^2 {\theta }\sin { 2 \phi }\right \rangle /2$ and $(2-2\left \langle \sin^2 {\theta }\cos { 2 \phi } \right \rangle - \left \langle \sin^2 {\theta }\sin { 2 \phi } \right \rangle ^2 )/4$. For small angles these quantities approximate the mean $\left \langle \phi \right \rangle$ and variance $\left \langle \phi ^2 \right \rangle$, and should therefore have a marked variation in the neighbourhood of $\lambda /b=1$ when plotted against $\lambda /b$. Another approach to evidence the effect of slip would be to conduct steady-state viscosity measurements (White et al. Reference White, Hawkins, Miyamoto, Takahara and Sue2015; Del Giudice & Shen Reference Del Giudice and Shen2017), by focusing on the high-$Pe$ regime where the effect of slip is maximum.

Our theory is not limited to nanometrically thin particles. Slip lengths of order $1\ \mathrm {\mu }\textrm {m}$ could be achieved for example, through the creation of a wetted/gas layer over the particle's surface (Lee, Charrault & Neto Reference Lee, Charrault and Neto2014), depletion layers (Tuinier & Taniguchi Reference Tuinier and Taniguchi2004; Fan, Dhont & Tuinier Reference Fan, Dhont and Tuinier2007) or surface nanobubbles (Neto et al. Reference Neto, Evans, Bonaccurso, Butt and Craig2005; Yang et al. Reference Yang, Kooij, Poelsema, Lohse and Zandvliet2008). This raises the possibility of conducting experiments with relatively large plate-like particles whose dynamics can be accessed optically.

Applications of the theory developed in the current paper may include: (i) the development of rheological models for 2-D nanomaterials in suspension (e.g. graphene liquid crystals); (ii) the design of experimental methods to measure the slip properties of plate-like particles or macromolecules, which could yield insights into molecular interactions at the solid–liquid interface; (iii) the use of hydrodynamic slip to favour the flow of 2-D nanomaterials in narrow channel (e.g. to avoid clogging). Regarding (i), because we have shown that slip plate-like particles have a smaller propensity to rotate in shear flow than no-slip particles, and thus interactions with other particles are more limited than in the no-slip case, theories developed for the dilute case could also apply to more concentrated systems, as pointed out previously for ring-shaped particles (Singh et al. Reference Singh, Koch and Stroock2013; Borker et al. Reference Borker, Stroock and Koch2018). Finally, alignment is an essential ingredient to impart superior properties to nanocomposite materials. Surface modifications producing substantial slip could be used to align particles produced from the exfoliation of 2-D layered materials, which can be easily produced on mass scales by liquid-phase exfoliation (Botto Reference Botto2019; Gravelle et al. Reference Gravelle, Kamal and Botto2020; Salussolia et al. Reference Salussolia, Barbieri, Pugno and Botto2020) and have therefore potential for applications. Because the ideal infinite Péclet number regime may not be achievable in practice, our results provide theoretical guidelines for deciding in which cases slip will have a dominant effect on the alignment of plate-like nanoparticles in the presence of Brownian motion.

Funding

This work was supported by the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation program through project FlexNanoFlow (Grant agreement no. 715475).

Declaration of interests

The authors report no conflict of interest.

References

REFERENCES

Aliyeva, S., Alosmanov, R., Buniyatzadeh, I., Azizov, A. & Maharramov, A. 2019 Recent developments in edge-selective functionalization of surface of graphite and derivatives–a review. Soft Matt. 17, 448466.CrossRefGoogle Scholar
Allison, S.A. 1999 Low Reynolds number transport properties of axisymmetric particles employing stick and slip boundary conditions. Macromolecules 32, 53045312.CrossRefGoogle Scholar
Anczurowski, E. & Mason, S.G. 1967 a The kinetics of flowing dispersions: II. Equilibrium orientations of rods and discs (theoretical). J. Colloid Interface Sci. 23, 522532.CrossRefGoogle Scholar
Anczurowski, E. & Mason, S.G. 1967 b The kinetics of flowing dispersions: III. Equilibrium orientations of rods and discs (experimental). J. Colloid Interface Sci. 23, 533546.CrossRefGoogle Scholar
Bocquet, L. & Barrat, J.L. 2007 Flow boundary conditions from nano-to micro-scales. Soft Matt. 3, 685693.CrossRefGoogle ScholarPubMed
Borker, N.S., Stroock, A.D. & Koch, D.L. 2018 Controlling rotation and migration of rings in a simple shear flow through geometric modifications. J. Fluid Mech. 840, 379407.CrossRefGoogle Scholar
Botto, L. 2019 Toward nanomechanical models of liquid-phase exfoliation of layered 2D nanomaterials: analysis of a $\pi -$peel model. Front. Mater. 6, 302.CrossRefGoogle Scholar
Brenner, H. 1974 Rheology of a dilute suspension of axisymmetric Brownian particles. Intl J. Multiphase Flow 1, 195341.CrossRefGoogle Scholar
Bretherton, F.P. 1962 The motion of rigid particles in a shear flow at low Reynolds number. J. Fluid Mech. 14, 284304.CrossRefGoogle Scholar
Brown, A.B.D, Clarke, S.M., Convert, P. & Rennie, A.R. 2000 Orientational order in concentrated dispersions of plate-like kaolinite particles under shear. J. Rheol. 44, 221233.CrossRefGoogle Scholar
Burgers, J.M. 1938 On the motion of small particles of elongated form suspended in a viscous liquid. Kon. Ned. Akad. Wet. Verhand.(Eerste Sectie) 16, 113184.Google Scholar
Chang, Y.C. & Keh, H.J. 2012 Creeping-flow rotation of a slip spheroid about its axis of revolution. Theor. Comput. Fluid Dyn. 26 (1–4), 173183.CrossRefGoogle Scholar
Chung, D.D.L. 2002 Review graphite. J. Mater. Sci. 37, 14751489.CrossRefGoogle Scholar
Cox, R.G. 1971 The motion of long slender bodies in a viscous fluid. Part 2. Shear flow. J. Fluid Mech. 45, 625657.CrossRefGoogle Scholar
Del Giudice, F., Cunning, B.V., Ruoff, R.S. & Shen, A.Q. 2018 Filling the gap between transient and steady shear rheology of aqueous graphene oxide dispersions. Rheol. Acta 57, 293306.CrossRefGoogle Scholar
Del Giudice, F. & Shen, A.Q. 2017 Shear rheology of graphene oxide dispersions. Curr. Opin. Chem. Engng 16, 2330.CrossRefGoogle Scholar
Doi, M. & Edwards, S.F. 1988 The Theory of Polymer Dynamics, International Series of Monographs on Physics, vol. 73. Oxford University Press.Google Scholar
Falk, K., Sedlmeier, F., Joly, L., Netz, R.R. & Bocquet, L. 2012 Ultralow liquid/solid friction in carbon nanotubes: comprehensive theory for alcohols, alkanes, OMCTS, and water. Langmuir 28, 1426114272.CrossRefGoogle ScholarPubMed
Fan, T.H., Dhont, J.K.G. & Tuinier, R. 2007 Motion of a sphere through a polymer solution. Phys. Rev. E 75, 011803.CrossRefGoogle ScholarPubMed
Férec, J., Heniche, M., Heuzey, M.C., Ausias, G. & Carreau, P.J. 2008 Numerical solution of the Fokker–Planck equation for fiber suspensions: application to the Folgar–Tucker–Lipscomb model. J. Non-Newtonian Fluid Mech. 155, 2029.CrossRefGoogle Scholar
Frattini, P.L. & Fuller, G.G. 1986 Rheo-optical studies of the effect of weak Brownian rotations in sheared suspensions. J. Fluid Mech. 168, 119150.CrossRefGoogle Scholar
Fuller, G.G. 1995 Optical Rheometry of Complex Fluids. Oxford University Press on Demand.Google Scholar
Gardiner, C. 2004 Handbook of Stochastic Methods: For Physics, Chemistry & The Natural Sciences. Synergetics, vol. 13. Springer.CrossRefGoogle Scholar
Giesekus, H. 1962 Elasto-viskose Flüssigkeiten, für die in stationären Schichtströmungen sämtliche Normalspannungskomponenten verschieden groß sind. Rheol. Acta 2, 5062.CrossRefGoogle Scholar
Gravelle, S., Joly, L., Ybert, C. & Bocquet, L. 2014 Large permeabilities of hourglass nanopores: from hydrodynamics to single file transport. J. Chem. Phys. 141, 1826.CrossRefGoogle ScholarPubMed
Gravelle, S., Kamal, C. & Botto, L. 2020 Liquid exfoliation of multilayer graphene in sheared solvents: A molecular dynamics investigation. J. Chem. Phys. 152 (10), 104701.CrossRefGoogle ScholarPubMed
Gravelle, S., Kamal, C. & Botto, L. 2021 Violations of Jeffery's theory in the dynamics of nanographene in shear flow. Phys. Rev. Fluids 6, 034303.CrossRefGoogle Scholar
Hao, R., Qian, W., Zhang, L. & Hou, Y. 2008 Aqueous dispersions of tcnq-anion-stabilized graphene sheets. Chem. Commun. 48, 65766578.CrossRefGoogle Scholar
Herzhaft, B. & Guazzelli, É 1999 Experimental study of the sedimentation of dilute and semi-dilute suspensions of fibres. J. Fluid Mech. 384, 133158.CrossRefGoogle Scholar
Hinch, E.J. & Leal, L.G. 1972 The effect of Brownian motion on the rheological properties of a suspension of non-spherical particles. J. Fluid Mech. 52, 683712.CrossRefGoogle Scholar
Hinch, E.J. & Leal, L.G. 1973 Time-dependent shear flows of a suspension of particles with weak Brownian rotations. J. Fluid Mech. 57, 753767.CrossRefGoogle Scholar
Huang, Y.Y. & Terentjev, E.M. 2012 Dispersion of carbon nanotubes: mixing, sonication, stabilization, and composite properties. Polymers 4, 275295.CrossRefGoogle Scholar
Ivanov, Y., Van de Ven, T.G.M. & Mason, S.G. 1982 Damped oscillations in the viscosity of suspensions of rigid rods. I. Monomodal suspensions. J. Rheol. 26, 213230.CrossRefGoogle Scholar
Jeffery, G.B. 1922 The motion of ellipsoidal particles immersed in a viscous fluid. Proc. R. Soc. Lond. A 102, 161179.Google Scholar
Jogun, S.M. & Zukoski, C.F. 1999 Rheology and microstructure of dense suspensions of plate-shaped colloidal particles. J. Rheol. 43 (4), 847871.CrossRefGoogle Scholar
Jonsson, U. & Bhushan, B. 1995 Measurement of rheological properties of ultrathin lubricant films at very high shear rates and near-ambient pressure. J. Appl. Phys. 78, 3107.CrossRefGoogle Scholar
Kamal, C., Gravelle, S. & Botto, L. 2020 Hydrodynamic slip can align thin nanoplatelets in shear flow. Nat. Commun. 11, 2425.CrossRefGoogle ScholarPubMed
Kannam, S.K., Todd, B.D., Hansen, J.S. & Daivis, P.J. 2013 How fast does water flow in carbon nanotubes? J. Chem. Phys. 138, 094701.CrossRefGoogle ScholarPubMed
Karagiannidis, P.G., et al. 2017 Microfluidization of graphite and formulation of graphene-based conductive inks. ACS Nano 11, 27422755.CrossRefGoogle ScholarPubMed
Keh, H.J. & Chang, Y.C. 2008 Slow motion of a slip spheroid along its axis of revolution. Intl J. Multiphase Flow 34 (8), 713722.CrossRefGoogle Scholar
Keh, H.J. & Huang, C.H. 2004 Slow motion of axisymmetric slip particles along their axes of revolution. Intl J. Engng Sci. 42, 16211644.CrossRefGoogle Scholar
Keller, S.R. & Skalak, R. 1982 Motion of a tank-treading ellipsoidal particle in a shear flow. J. Fluid Mech. 120, 2747.CrossRefGoogle Scholar
Kim, S. & Karrila, S.J. 2013 Microhydrodynamics: Principles and Selected Applications. Courier Corporation.Google Scholar
Koltonow, A.R., Luo, C., Luo, J. & Huang, J. 2017 Graphene oxide sheets in solvents: to crumple or not to crumple? ACS Omega 2, 80058009.CrossRefGoogle ScholarPubMed
Kumar, A., Sharma, K. & Dixit, A.R. 2019 A review of the mechanical and thermal properties of graphene and its hybrid polymer nanocomposites for structural applications. J. Mater. Sci. 54, 59926026.CrossRefGoogle Scholar
Lauga, E., Brenner, M.P. & Stone, H.A. 2008 Microfluidics: the no-slip boundary condition. In Springer Handbook of Experimental Fluid Mechnaics (ed. C. Tropea, J.F. Foss & A. Yarin). p. 1219. Springer.CrossRefGoogle Scholar
Leahy, B.D., Cheng, X., Ong, D.C., Liddell-Watson, C. & Cohen, I. 2013 Enhancing rotational diffusion using oscillatory shear. Phys. Rev. Lett. 110, 228301.CrossRefGoogle ScholarPubMed
Leahy, B.D., Koch, D.L. & Cohen, I. 2015 The effect of shear flow on the rotational diffusion of a single axisymmetric particle. J. Fluid Mech. 772, 4279.CrossRefGoogle Scholar
Leahy, B.D., Koch, D.L. & Cohen, I. 2017 Controlling the alignment of rodlike colloidal particles with time-dependent shear flows. J. Rheol. 61, 979996.CrossRefGoogle Scholar
Leal, L.G. & Hinch, E.J. 1971 The effect of weak Brownian rotations on particles in shear flow. J. Fluid Mech. 46, 685703.CrossRefGoogle Scholar
Lee, T., Charrault, E. & Neto, C. 2014 Interfacial slip on rough, patterned and soft surfaces: A review of experiments and simulations. Adv. Colloid Interface Sci. 210, 2138.CrossRefGoogle ScholarPubMed
Lindahl, N., Midtvedt, D., Svensson, J., Nerushev, O.A., Lindvall, N., Isacsson, A. & Campbell, E.E.B. 2012 Determination of the bending rigidity of graphene via electrostatic actuation of buckled membranes. Nano Lett. 12, 35263531.CrossRefGoogle ScholarPubMed
Lingard, P.S. & Whitmore, R.L. 1974 The deformation of disc-shaped particles by a shearing fluid with application to the red blood cell. J. Colloid Interface Sci. 49, 119127.CrossRefGoogle Scholar
Loyalka, S.K. & Griffin, J.L. 1994 Rotation of non-spherical axi-symmetric particles in the slip regime. J. Aerosol Sci. 25, 509525.CrossRefGoogle Scholar
Luan, B. & Zhou, R. 2016 Wettability and friction of water on a MoS2 nanosheet. Appl. Phys. Lett. 108, 131601.CrossRefGoogle Scholar
Luo, H. & Pozrikidis, C. 2008 Effect of surface slip on Stokes flow past a spherical particle in infinite fluid and near a plane wall. J. Engng Maths 62, 121.CrossRefGoogle Scholar
Maali, A., Cohen-Bouhacina, T. & Kellay, H. 2008 Measurement of the slip length of water flow on graphite surface. Appl. Phys. Lett. 92, 053101.CrossRefGoogle Scholar
Mueller, S., Llewellin, E.W. & Mader, H.M. 2009 The rheology of suspensions of solid particles. Proc. R. Soc. A 466, 12011228.CrossRefGoogle Scholar
Neto, C., Evans, D.R., Bonaccurso, E., Butt, H.-J. & Craig, V.S.J. 2005 Boundary slip in Newtonian liquids: a review of experimental studies. Rep. Prog. Phys. 68, 2859.CrossRefGoogle Scholar
Ortiz-Young, D., Chiu, H.-C., Kim, S., Voïtchovsky, K. & Riedo, E. 2013 The interplay between apparent viscosity and wettability in nanoconfined water. Nat. Commun. 4, 2482.CrossRefGoogle ScholarPubMed
Park, J., Kim, Y.S., Sung, S.J., Kim, T. & Park, C.R. 2017 Highly dispersible edge-selectively oxidized graphene with improved electrical performance. Nanoscale 9, 16991708.CrossRefGoogle ScholarPubMed
Paton, K.R., Anderson, J., Pollard, A.J. & Sainsbury, T. 2017 Production of few-layer graphene by microfluidization. Mater. Res. Express 4, 025604.CrossRefGoogle Scholar
Paton, K.R., et al. 2014 Scalable production of large quantities of defect-free few-layer graphene by shear exfoliation in liquids. Nat. Mater. 13 (6), 624630.CrossRefGoogle ScholarPubMed
Peterlin, A. 1938 Über die Viskosität von verdÜnnten Lösungen und Suspensionen in Abhängigkeit von der Teilchenform. Z. Phys. 111, 232263.CrossRefGoogle Scholar
Poot, N. & van der Zant, H.S.J. 2008 Nanomechanical properties of few-layer graphene membranes. Appl. Phys. Lett. 92, 063111.CrossRefGoogle Scholar
Poulin, P., Jalili, R., Neri, W., Nallet, F., Divoux, T., Colin, A., Aboutalebi, S.H., Wallace, G. & Zakri, C. 2016 Superflexibility of graphene oxide. Proc. Natl Acad. Sci. 113, 1108811093.CrossRefGoogle ScholarPubMed
Pozrikidis, C. 1992 Boundary Integral and Singularity Methods for Linearized Viscous Flow. Cambridge University Press.CrossRefGoogle Scholar
Pozrikidis, C. 2002 A Practical Guide to Boundary Element Methods with the Software Library BEMLIB. CRC Press.CrossRefGoogle Scholar
Press, W.H., Teukolsky, S.A., Vetterling, W.T. & Flannery, B.P. 2007 Numerical Recipes 3rd Edition: The Art of Scientific Computing. Cambridge University Press.Google Scholar
Radisavljevic, B., Radenovic, A., Brivio, J., Giacometti, V. & Kis, A. 2011 Single-layer MoS2 transistors. Nat. Nanotechnol. 6, 147150.CrossRefGoogle ScholarPubMed
Rallison, J.M. 1978 The effects of Brownian rotations in a dilute suspension of rigid particles of arbitrary shape. J. Fluid Mech. 84, 237263.CrossRefGoogle Scholar
Reddy, N.K., Natale, G., Prud'homme, R.K. & Vermant, J. 2018 Rheo-optical analysis of functionalized graphene suspensions. Langmuir 34, 78447851.CrossRefGoogle ScholarPubMed
Reddy, N.K., Perez-Juste, J., Pastoriza-Santos, I., Lang, P.R., Dhont, J.K.G., Liz-Marzan, L.M. & Vermant, J. 2011 Flow dichroism as a reliable method to measure the hydrodynamic aspect ratio of gold nanoparticles. ACS Nano 5, 49354944.CrossRefGoogle ScholarPubMed
Sadron, C. 1953 Flow Properties of Disperse Systems. North-Holland.Google Scholar
Salussolia, G., Barbieri, E., Pugno, N.M. & Botto, L. 2020 Micromechanics of liquid-phase exfoliation of a layered 2d material: a hydrodynamic peeling model. J. Mech. Phys. Solids 134, 103764.CrossRefGoogle Scholar
Scheraga, H.A. 1955 Non-Newtonian viscosity of solutions of ellipsoidal particles. J. Chem. Phys. 23, 1526.CrossRefGoogle Scholar
Sellier, A. 2012 Stokes flow about a slip arbitrary-shaped particle. CMES 87, 157176.Google Scholar
Sellier, A. 2013 Arbitrary Stokes flow about a fixed or freely-suspended slip particle. Comp. Model. Engng Comp. 96, 159176.Google Scholar
Sherwood, J.D. & Meeten, G.H. 1991 The use of the vane to measure the shear modulus of linear elastic solids. J. Non-Newtonian Fluid Mech. 41, 101118.CrossRefGoogle Scholar
Sherwood, J.D. 2012 Resistance coefficients for Stokes flow around a disk with a Navier slip condition. Phys. Fluids 24, 093103.CrossRefGoogle Scholar
Silmore, K.S., Strano, M. & Swan, J.W. 2021 Buckling, crumpling, and tumbling of semiflexible sheets in simple shear flow. Soft Matt.CrossRefGoogle ScholarPubMed
Singh, V., Koch, D.L. & Stroock, A.D. 2013 Rigid ring-shaped particles that align in simple shear flow. J. Fluid Mech. 722, 121158.CrossRefGoogle Scholar
Singh, V., Koch, D.L., Subramanian, G. & Stroock, A.D. 2014 Rotational motion of a thin axisymmetric disk in a low Reynolds number linear flow. Phys. Fluids 26, 033303.CrossRefGoogle Scholar
Stover, C.A., Koch, D.L. & Cohen, C. 1992 Observations of fibre orientation in simple shear flow of semi-dilute suspensions. J. Fluid Mech. 238, 277296.CrossRefGoogle Scholar
Tawari, A.L., Koch, D.L. & Cohen, C. 2001 Electrical double-layer effects on the Brownian diffusivity and aggregation rate of laponite clay particles. J. Colloid Interface Sci. 240, 5466.CrossRefGoogle ScholarPubMed
Tocci, G., Joly, L. & Michaelides, A. 2014 Friction of water on graphene and hexagonal boron nitride from ab initio methods: very different slippage despite very similar interface structures. Nano Lett. 14, 68726877.CrossRefGoogle ScholarPubMed
Tuinier, R. & Taniguchi, T. 2004 Polymer depletion-induced slip near an interface. J. Phys-Condens. Mater. 17, L9.CrossRefGoogle Scholar
Vermant, J., Yang, H. & Fuller, G.G. 2001 Rheooptical determination of aspect ratio and polydispersity of nonspherical particles. AICHE J. 47, 790798.CrossRefGoogle Scholar
Voeltzel, N., Fillot, N., Vergne, P. & Joly, L. 2018 Orders of magnitude changes in the friction of an Ionic liquid on carbonaceous surfaces. J. Phys. Chem. C 122, 21452154.CrossRefGoogle Scholar
Wei, N., Peng, X. & Xu, Z. 2014 Breakdown of fast water transport in graphene oxides. Phys. Rev. E 89 (1), 012113.CrossRefGoogle ScholarPubMed
White, K.L., Hawkins, S., Miyamoto, M., Takahara, A. & Sue, H.-J. 2015 Effects of aspect ratio and concentration on rheology of epoxy suspensions containing model plate-like nanoparticles. Phys. Fluids 27, 123306.CrossRefGoogle Scholar
Wick, P., et al. 2014 Classification framework for graphene-based materials. Angew. Chem. Int. Ed. 53, 77147718.CrossRefGoogle ScholarPubMed
Willenbacher, N., Hanciogullari, H. & Wagner, H.G. 1997 High shear rheology of paper coating colors–more than just viscosity. Chem. Engng Technol. 20, 557563.CrossRefGoogle Scholar
Willis, D.G. 1977 A kinematic model of preferred orientation. Geol. Soc. Am. Bull. 88, 883894.2.0.CO;2>CrossRefGoogle Scholar
Xu, Y. & Green, M.J. 2014 Brownian dynamics simulations of nanosheet solutions under shear. J. Chem. Phys. 141, 024905.CrossRefGoogle ScholarPubMed
Yang, S., Kooij, E.S., Poelsema, B., Lohse, D. & Zandvliet, H.J.W. 2008 Correlation between geometry and nanobubble distribution on HOPG surface. Europhys. Lett. 81, 64006.CrossRefGoogle Scholar
Youngren, G.K. & Acrivos, A. 1975 Rotational friction coefficients for ellipsoids and chemical molecules with the slip boundary condition. J. Chem. Phys. 63, 38463848.CrossRefGoogle Scholar
Zhang, J., Xu, X. & Qian, T. 2015 Anisotropic particle in viscous shear flow: Navier slip, reciprocal symmetry, and Jeffery orbit. Phys. Rev. E 91, 033016.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. Sketch illustrating the rotational dynamics of a platelet having (a) zero slip length, as described by Jeffery's theory and (b) slip length $\lambda$ significantly greater than the platelet's half-thickness $b$. The behaviours described in panels (a,b) are characteristic of the infinite Péclet number limit. In this paper we enquire about the effect of Brownian fluctuations on the rotational dynamics.

Figure 1

Figure 2. (a) Rotation angle $\phi$ for a particle with real effective aspect ratio $k_{e} = 0.16$ and ${Pe} = \infty$. (b) Associated angular velocity $\dot {\gamma }\varOmega$; notice that $\varOmega <0 \ \forall \phi$. (c) Rotation angle $\phi$ for a particle with imaginary effective aspect ratio $k_{e} = 0.28 \text {i}$ and ${Pe} = \infty$, showing stabilisation at $\phi _{c} \approx 0.27$. (d) Corresponding non-dimensional angular velocity, with the two equilibrium points at $\phi = \pm \phi _{c}$. The ‘${\bigodot}$’ symbol marks the stable angle $\phi _{c}$; the angle denoted by ‘${\bigotimes}$’ is unstable.

Figure 2

Figure 3. Time evolution of the rotation angle for different $Pe$, as calculated from (3.1), comparing (ac) real $k_{e}$ ($k_{e}=0.11$) to (df) purely imaginary $k_{e}$ ($k_{e}=0.21\textrm {i}$). The red dashed line is the value of $\phi _{c}$ obtained for ${Pe} \to \infty$.

Figure 3

Figure 4. Spectral coefficients $a_n$ and $b_n$ for $k_e=0.05\text {i}$ and $Pe=10^{6}$ and $t \to \infty$. The spectral coefficients are truncated at $n=1000$, since $a_n, b_n \to 0$ above this threshold.

Figure 4

Figure 5. Mean (a), variance (b) and skewness (c) of the angular probability distribution versus time, for a nanoplatelet with $k_{e}=0.1\text {i}$ and different values of $Pe$.

Figure 5

Figure 6. Steady-state orientation distribution function $p(\phi )$ for a real $(k_e = 0.1)$ or a purely imaginary $(k_e = 0.1\textrm {i})$ value of $k_{e}$ for different $Pe$ numbers. The red dotted line marks the orientation distribution function for ${Pe} \to \infty$ and $k_{e}=0.1$ (a) and the location of $\phi _{c}$ for $k_{e}=0.1 i$ (b). The values [$\mu$, $\sigma ^2$, $\left \langle (\phi -\mu )^3\right \rangle$] in increasing order of $Pe$ are $[0.17,0.39,-0.36]$, $[0.069,0.21,-0.40]$ and $[0.019,0.14,-0.30]$, respectively, for $k_{e}=0.1$; and $[0.19,0.38,-0.37]$, $[0.094,0.16,-0.53]$ and $[0.075,0.033,-1.1]$, respectively, for $k_{e}=0.1\text {i}$.

Figure 6

Figure 7. Colour maps of the mean $\mu$, variance $\sigma ^2$ and skewness $\left \langle (\phi - \mu )^3\right \rangle$ of the steady-state orientation distribution in the ($Pe$, $k_{e}$) space. The black straight line is ${Pe}=|k_{e}|^{-3}$. For $Pe$ above this line, the difference in the moments between real and purely imaginary values of $k_{e}$ becomes noticeable.

Figure 7

Figure 8. Threshold value of the Péclet number ${Pe}_{c}$ beyond which no platelet rotation is expected, as a function of $|k_{e}|$ (with $k_{e}$ purely imaginary). Black disks are numerical values, lines are fits for $|k_{e}| \gtrapprox 0.1$ (dot–dashed blue line) and $|k_{e}| \lessapprox 0.1$ (dashed red line), respectively.

Figure 8

Figure 9. (ac) Water density profiles as extracted from MD simulations for a number of layer $n=1$ (a), $2$ (b) and $3$ (c), respectively. The centres of the carbon atoms are located along the grey lines. Details of the molecules dynamics simulations are given in Kamal et al. (2020). The colour field is the water density, from white (low density) to blue (high density), and the red dashed lines show the reference surfaces. (d) Sketch of the reference surface for $n=3$, showing the slip length $\lambda$, the effective radius $\xi$ of the carbon atoms and the interlayer spacing $d_{gg}$.

Figure 9

Table 1. Literature values of the slip length for different solid material/solvent combinations. The materials and liquids are: diamond-like carbon (DIC); graphene oxide (GO); boron nitride (BN); molybdenum disulfide (MoS$_2$); ionic liquid (IL); n-methyl-2-pyrrolidone (NMP); octamethylcyclotetrasiloxane (OMCTS); cyclopentanone (CPO). Ab initio MD refers to a method allowing for the calculation of electronic behaviour from first principles by using a quantum mechanical method.

Figure 10

Figure 10. (a) Example of the discretisation of a $n=1$ surface with $N=96$. (b,c) Effective aspect ratio $|k_e|$ versus the number of discretisation points $N$ for platelets with $\lambda =20\ \textrm {nm}$, $a/b=10$ or $a/b=100$, and either $n=1$ (b) or $n=10$ (c).

Figure 11

Figure 11. Absolute value of the effective aspect ratio $k_{e}$ versus the platelet length $2 a$ for $\xi =0.25\ \textrm {nm}$ and for $n=1$ (a) and $n=10$ (b). The black lines correspond to the purely imaginary values of $k_{e}$ and the red lines to real values of $k_{e}$.

Figure 12

Figure 12. Effective aspect ratio $k_{e}^2$ versus the slip length $\lambda$ for different platelet lengths $2a$. The half-thickness of the platelet is $b\approx 0.25\ \textrm {nm}$ and $b\approx 1.75\ \textrm {nm}$ for $n=1$ and $n=10$, respectively. Red dashed lines mark the positions where $k_{e}^2$ changes sign.

Figure 13

Figure 13. The streamwise (a) and normal (b) hydrodynamic traction distribution along the surface of a platelet fixed at an orientation $\phi =0$ for $n=1$ and $a/b=10$. The vertical dotted line marks the position of the edge of the platelet.

Figure 14

Figure 14. Comparison of mean and variance of the fluctuating rotation angle for a nanoplatelet with slip ($k_{e}=0.063\text {i}$) and without slip ($k_{e}=0.028$). The red dashed line marks the asymptotic solution for $\sigma ^2$ for purely imaginary values of $k_{e}$. The nanoplatelet has half-width $b=0.25\ \textrm {nm}$ (i.e. $n=1$) and half-length $a=25\ \textrm {nm}$. Region I: ${Pe} < {Pe}_{I}$, negligible slip effects for both $\lambda = 0$ and $\lambda = 2$  nm. Region II: ${Pe}_{I} \leq {Pe}<{Pe}_{c}$, moderate slip effects region characterised by changes in the frequency of rotation due to slip. Region III: ${Pe}\geq {Pe}_{c}$, large slip effects region where slip completely suppresses the rotation of the nanoplatelet.

Figure 15

Figure 15. Effect of slip on the rotational dynamics of a nanoplatelet with $\text {length} = \text {depth} = 2a$ and $\text {width} = 2b$ in the ($Pe$, $2a$) space. Upper black curve marks the threshold Péclet number ${Pe}_{c}$ for which $\sigma ^2=2 k_{e}$; above this threshold the hydrodynamic slip has a dramatic effect on the particle dynamics. Lower black curve marks the threshold Péclet number ${Pe}_{I}$; above this threshold, slip effects become important. The dashed, dashed–dot and dash–double-dot lines correspond to $\dot {\gamma }\eta = 10^7\ \text {Pa}$, $\dot {\gamma }\eta = 2\times 10^4\ \text {Pa}$, $\dot {\gamma }\eta = 10\ \text {Pa}$, respectively.

Figure 16

Figure 16. Colour maps of $\left \langle 1-\cos {(4\phi )}\right \rangle$ and $\left \langle \sin {(2\phi )}\right \rangle$ in the ($Pe$, $k_{e}$) space for (a,c) an imaginary value of $k_{e}$ and (b,d) a real value of $k_{e}$. The black line marks the equation ${Pe}=|k_{e}|^{-3}$.

Figure 17

Figure 17. Comparison of our theory (for $\theta ={\rm \pi} /2$) with experimental results of Reddy et al. (2018) for $\chi$ and $\Delta n'' /\Delta n''_{max}$ (symbols), as defined in (5.1a,b). The different lines correspond to the following model parameters: functionalised graphene plateletes (FGS) with $k_{e}\approx 0.016$ (no-slip) and $k_{e}=0.035\text {i}$ (slip with $\lambda =20\ \text {nm}$); slightly oblate nanospheroids (NP) with $b/a=0.55$, for $k_{e} \approx 0.55$ (no-slip) and $k_{e}\approx 0.46$ (slip with $\lambda /b=0.21$).