You are on page 1of 18

Orientational dynamics and rheology of active suspensions in weakly viscoelastic flows

Akash Choudhary1 ,∗ Sankalp Nambiar2 , and Holger Stark1†


1
Institute of Theoretical Physics, Technische Universität Berlin, Hardenbergstr. 36, 10623 Berlin, Germany
2
Nordita, KTH Royal Institute of Technology and Stockholm University, Stockholm 10691, Sweden

Abstract. Microswimmer suspensions are constantly in non-equilibrium and exhibit macroscale


properties that are in stark contrast to passive suspensions. Motivated by ubiquitous microbial sys-
tems suspended in biological fluids, we analyse the rheological response of a suspension of elongated
microswimmers to a steady shear flow in a weakly viscoelastic fluid. At the individual level, we
find that the viscoelastic stresses generated by activity substantially modify the Jeffery orbits well-
known from Newtonian fluids. The orientational dynamics depends on the swimmer type and, in
particular, microswimmers can resist flow-induced rotation and align at an angle with the flow. To
arXiv:2303.15241v1 [cond-mat.soft] 27 Mar 2023

analyze its impact on the bulk rheological response, we study a dilute ensemble of microswimmers in
the presence of stochastic noise due to tumbling and rotational diffusion. Compared to Newtonian
fluids, activity and its elastic response in polymeric fluids alter the orientational distribution and
substantially amplify the swimmer-induced viscosity. This suggests that pusher suspensions reach
the regime of superfluidity at lower volume fractions.

INTRODUCTION even down to the ‘superfluidic’ limit [21]. The mecha-


nism, first outlined in Ref. [22] and elaborated further in
Systems of particulate matter suspended in fluids Refs. [23, 24], is as follows. The orientations of elongated
are prevalent in numerous natural and industrial pro- swimmers in shear flow follow periodic Jeffery orbits [25]
cesses. Active suspensions are particulate systems that and are also subject to thermal noise. This competition
are driven out of equilibrium by converting chemical yields a mean orientation that points in the extensional
energy or fuel into mechanical work to achieve self- quadrant of applied shear flow. With such an orientation
propulsion [1, 2]. Motile microorganisms are the pro- extensile microswimmers (pushers) support the applied
totypical example of active motion that is generated shear flow and thereby reduce the effective shear vis-
via metachronal actuation of hairlike cilia (Paramecium, cosity, whereas contractile swimmers like C. reinhardtii
Volvox ), by whipping cell-attached flagellar appendages (pullers) resist it and thus increase viscosity [26].
(spermatozoa, algae), or by rotating a bundle of helical Since almost all biological fluids are non-Newtonian,
flagella (bacteria) [3]. Microswimmers navigate through there has been a recent interest towards developing theo-
their environment by sensing or interacting with gradi- retical frameworks that capture the individual and collec-
ents in hydrodynamic, chemical, thermal, and light fields; tive dynamics of microswimmers in complex fluids [27–
a strategy known as taxis [4–10]. A particular example 29]. For example, experiments have shown reduced tum-
is rheotaxis, where microswimmers experience hydrody- bling and increased persistence lengths of bacteria in
namic gradients and swim against the flow, which is rele- polymeric fluids [30–32]. Viscoelasticity can also initi-
vant for biofilm formation and reproduction [2, 8, 11, 12]. ate spatiotemporal order in active suspensions. For ex-
Pathogenic microswimmers when infiltrating human ample, a recent study found that DNA polymers trig-
and animal bodies have to pass through mucus linings ger oscillatory vortices in confined suspension of E.coli
that lubricate and protect our respiratory tracks, eyes, [33]. However, determining the effective shear viscosity
urogenital and gastrointestinal systems [13]. The pres- of an active suspension in a viscoelastic fluid has been an
ence of mucin fibres (typically 3-10 nm in length [14]) uncharted territory because of its complexity: the elas-
and DNA makes these mucus linings viscoelastic and tic relaxation of non-Newtonian fluids and their shear-
shear thinning [15]. The linings are often subjected to thinning/thickening property.
shearing motion, for instance, during blinking, cough- To gain better insights into biological and artificial
ing, reproduction, and continuos mucociliary clearance motility in biological fluids, this communication inves-
in respiratory systems. In such microbiological flows, the tigates the role of non-linear polymeric stresses in non-
non-Newtonian fluid properties can alter the swimmer’s Newtonian fluids. Specifically, this non-linearity allows
rheotactic behaviour [16, 17]. us to directly couple a background shear flow to the ac-
In this article we study the bulk rheology of mi- tive disturbance flow generated by a microswimmer. For
croswimmers in viscoelastic flows. For Newtonian flu- spherical microswimmers in a Poiseuille flow, we already
ids, several rheological experiments on suspensions of showed that due to such a coupling they experience a
extensile swimmers like E.coli have shown that activ- swimming lift force that depends on the swimmer type
ity can drastically reduce the effective viscosity [18–20], [17]. Here, we will show that this nonlinear coupling in-
fluences the Jeffery orbit of an elongated swimmer in a
shear flow, which thereby also fundamentally alters the
∗ a.choudhary@campus.tu-berlin.de bulk rheological response. Such activity-induced changes
† holger.stark@tu-berlin.de in the orientational dynamics cannot be observed in New-
2

(a) (b) y tal stress tensor. It follows the second-order fluid (SOF)
model: T = −P I + 2 E + Wi S [34]. Here, E denotes the
V 1 = ˙ yex
<latexit sha1_base64="bhJ+f53R3xTLGr2yCQCcOIjdlp4=">AAACJHicbVDLSsQwFE19O76qLt0EB8HV0IqoIMKgG5cKzgOmdUgz6RhM0pLciqX0Y9z4K25c+MCFG7/FzGMxPg6EHM65N7n3RKngBjzv05manpmdm19YrCwtr6yuuesbTZNkmrIGTUSi2xExTHDFGsBBsHaqGZGRYK3o9mzgt+6YNjxRV5CnLJSkr3jMKQErdd3jIEpEz+TSXkWzvC4CrmLIS3yCg14CRdAnUpIS53iykJXd4r7sulWv5g2B/xJ/TKpojIuu+2bfpJlkCqggxnR8L4WwIBo4FaysBJlhKaG3pM86lioimQmL4ZIl3rFKD8eJtkcBHqqTHQWRZjCerZQEbsxvbyD+53UyiI/Cgqs0A6bo6KM4ExgSPEgM97hmFERuCaGa21kxvSGaULC5VmwI/u+V/5LmXs0/qO1f7lfrp+M4FtAW2ka7yEeHqI7O0QVqIIoe0BN6Qa/Oo/PsvDsfo9IpZ9yziX7A+foGpZ+mrg==</latexit>
<latexit sha1_base64="EMzZGwJtY7oVIh0YKPvwwmAZe6E=">AAAB9XicbVDLSgMxFL1TX7W+qi7dBIvgqsyIqMuiG5cV7APasWQymTY0kwxJRilD/8ONC0Xc+i/u/Bsz7Sy09UDI4Zx7yckJEs60cd1vp7Syura+Ud6sbG3v7O5V9w/aWqaK0BaRXKpugDXlTNCWYYbTbqIojgNOO8H4Jvc7j1RpJsW9mSTUj/FQsIgRbKz00A8kD/UktleWTAfVmlt3Z0DLxCtIDQo0B9WvfihJGlNhCMda9zw3MX6GlWGE02mln2qaYDLGQ9qzVOCYaj+bpZ6iE6uEKJLKHmHQTP29keFY59HsZIzNSC96ufif10tNdOVnTCSpoYLMH4pSjoxEeQUoZIoSwyeWYKKYzYrICCtMjC2qYkvwFr+8TNpnde+ifn53XmtcF3WU4QiO4RQ8uIQG3EITWkBAwTO8wpvz5Lw4787HfLTkFDuH8AfO5w9KEpMK</latexit>

p ∆
rate-of-strain tensor and S = 4E ⋅ E + 2δ E is the poly-
" meric stress tensor, which is quadratic in E and contains
Us !
<latexit sha1_base64="KjPx5yUrPVMO98KQD3iO9bcr4K0=">AAAB6nicbVBNS8NAEJ3Ur1q/qh69LBbBU0mkqMeiF48VTVtoQ9lsJ+3SzSbsboRS+hO8eFDEq7/Im//GbZuDtj4YeLw3w8y8MBVcG9f9dgpr6xubW8Xt0s7u3v5B+fCoqZNMMfRZIhLVDqlGwSX6hhuB7VQhjUOBrXB0O/NbT6g0T+SjGacYxHQgecQZNVZ68Hu6V664VXcOskq8nFQgR6NX/ur2E5bFKA0TVOuO56YmmFBlOBM4LXUzjSllIzrAjqWSxqiDyfzUKTmzSp9EibIlDZmrvycmNNZ6HIe2M6ZmqJe9mfif18lMdB1MuEwzg5ItFkWZICYhs79JnytkRowtoUxxeythQ6ooMzadkg3BW355lTQvqt5ltXZfq9Rv8jiKcAKncA4eXEEd7qABPjAYwDO8wpsjnBfn3flYtBacfOYY/sD5/AE9VI3I</latexit>

z the lower-convected time derivative. The SOF model


x not only allows us to capture the elastic effects perti-
nent to Boger fluids, but also to obtain analytical re-
sults for small Wi. Since we consider a steady shear rate
Fig. 1. Microswimmers in viscoelastic shear flow. in this work, we disregard the partial time derivative of
Schematics showing (a) the dilute suspension of microswim- E. In above governing equations, length, velocity, and
mers in an external shear flow V ∞ of a viscoelastic fluid, (b) pressure are already non-dimensionalized by the swim-
coordinate system moving with an individual swimmer that mer length (l = 2a), γ̇l, and µf γ̇, respectively, where µf
is modelled as an active prolate spheroid of major axis a and is the fluid shear viscosity. Also, the stress tensor is writ-
minor axis b. Here p and Us denote the orientation and swim- ten in dimensionless units using characteristic numbers
ming speed, respectively. (Wi and δ). In particular, Wi = trelax γ̇ is the shear based
Weissenberg number that quantifies the importance of
elasticity in the medium. It compares the shear rate γ̇
tonian fluids because the viscous stress tensor does not or inverse shearing time to the polymer relaxation time
permit such a nonlinear coupling. trelax = (Ψ1 + Ψ2 )/µf , where Ψ1 and Ψ2 are the normal
The current work uses the model of a second-order fluid stress coefficients. Furthermore, δ = −Ψ1 /2(Ψ1 + Ψ2 ) is
with a single elastic relaxation time that captures the the viscometric parameter that typically varies between
dynamics of polymeric fluids in the dilute limit (Boger −0.7 to −0.5. In ‘Methods: Hydrodynamic model’ we ex-
fluids) [34, 35]. For weak elasticity, quantified by the plain in detail how we solve the governing equations using
Weissenberg number, we perform a perturbative analy- a systematic perturbation expansion in Wi to address the
sis. Thereby we show that the inherent non-linearity in limit of weak viscoelasticity.
the polymeric stress tensor together with activity signifi- Active spheroid in viscoelastic shear flow. To
cantly alters the Jeffrey orbits known from the Newtonian determine the dynamics of an active spheroid, we first
fluid and of passive rods in a viscoelastic fluid [36, 37]. note that a swimmer disturbs the flow field both pas-
To evaluate the influence of thermal noise, we combine sively (due to its rigid body) and actively (due to self-
our results with the orientational Smoluchowski equation propulsion). The disturbance fields are implemented us-
and find that the altered deterministic dynamics also sub- ing hydrodynamic multipoles [39, 40]. Flagellated mi-
stantially affects the orientational distribution, known as croswimmers like E.coli and Chlamydomonas generate
the suspension microstructure [38]. It strongly differs a force-dipole flow field and higher order disturbances:
between extensile (pushers) and contractile (pullers) mi- source dipole, rotlet dipole, and force quadrupole. We
croswimmers. The orientational distribution allows us included all four of them and found that only the force-
to directly determine the effective viscosity of the active dipole flow field affects the swimmer dynamics in leading
suspension from an orientational average over the swim- order in Wi. The force-dipole field around the active
mer stresslets. Our analysis shows that fluid elasticity spheroid is σ rr̂2 [3(r̂ ⋅ p)2 − 1] with r̂ = r/r. Here, σ is
reduces the effective viscosity of active suspensions for a non-dimensional parameter equal to the ratio of the
both pushers and pullers compared to Newtonian fluids. force-dipole strength (σ ∗ ) to the stresslet imposed by the
The reduction increases with activity and allows to reach shear flow (8πµf γ̇l3 ), where σ > 0 represents pushers and
the superfluidic limit for smaller swimmer densities. σ < 0 pullers. For the current work, we focus on swim-
mers of 5µm size and shear rates of order 1−10s−1 , which
corresponds to σ of typical wild-type E.coli being roughly
RESULTS 0.04 − 0.1 [41, 42].
In Newtonian fluids, the equation of motion for the
Setup and second-order fluid. We consider a di- orientation p (θ, φ) gives the Jefferey orbits. To formu-
lute suspension of microswimmers in a shear flow of a late this equation for viscoelastic shear, we note that the
viscoelastic fluid, for instance, consisting of polymers polymeric stress tensor S is quadratic in the rate-of-strain
dissolved in a Newtonian fluid as shown in Fig. 1(a). tensor and vorticity [34]. Thus, similar to Ref. [43] we
Figure 1(b) depicts the coordinate system moving with determine all terms that by symmetry contribute to the
the microswimmer that is modelled as an active prolate rate of change ṗ up to first order in Wi. Neglecting small
spheroid and swims with speed Us . The uniformly dis- terms, we arrive at
tributed polymers are much smaller than the microswim-
mers and hence modeled within a continuum description. ṗ = (I − pp) ⋅ (E∞ ⋅ p) [Λ + Wi σα1 + Wi β1 E∞ ∶ pp]
The inertia-less hydrodynamics is governed by the mass + Ω∞ × p [1 + Wi σα2 + Wi β2 E∞ ∶ pp] + O(Wi2 ) (1)
and momentum conservation as ∇ ⋅ V = 0 and ∇ ⋅ T = 0,
respectively, where V is the velocity field and T is the to- in non-dimensional form, where Ω is the angular veloc-
3

(a) 𝑦 Log-rolling (b) 𝑦 Log-rolling (c) 𝑦 Alignment


(Passive) (Weak Pusher) (Pusher) 0.01
0
- 0.01
𝑥 𝑥 𝑥 -2 -1 0
Align.
𝑧 𝑧 𝑧
𝜙!"
(d) 𝑦 Log-rolling (e) 𝑦 (f) 𝑦 Alignment Align.
Shear plane rotation SPR
(Puller) (Strong Puller) (V. Strong Puller)
LR

𝑥 𝑥 𝑥
𝑧 𝑧 𝑧
𝜎
Fig. 2. Orientation dynamics of passive (σ = 0) and active
Fig. 3. State diagram showing the different states of the
particles in weakly viscoelastic shear flow. The blue curves
orientational dynamics for three Wi values. With increas-
show time traces of the orientation vector p on the unit sphere
ing dipole strength σ from left to right, we observe align-
starting at the red dot. (a) σ = 0, (b) σ = 0.06, (c) σ = 0.2,
ment (Align.), shear-plane rotation (SPR), log rolling (LR)
(d) σ = −0.2, (e) σ = −0.8, (f) σ = −2.4 . Other parameters:
and again alignment. In the alignment state, the alignment
Wi = 0.2, λ = 5, θ0 = π/5, φ0 = π/2, α1 = 4.62, α2 = −0.33,
angle φeq is plotted versus σ as solid lines. The dots on the
β1 = 0.68, and β2 = 1.46, δ = −0.6.
horizontal axis indicate the critical values σc at which align-
ment (Align.) occurs either to the left of the dashed-dotted
lines (σ < 0) or to the right (σ > 0). The dashed lines sepa-
ity and the superscript ∞ denotes the quantities that rate SPR and LR from each other at σ = −1.93, −0.49, −0.19
belong to the prescribed shear flow. The shape factor for increasing Wi. The inset shows the variation of Lyapunov
2
Λ = −1+λ contains the aspect ratio λ = a/b, the ratio exponent with σ. Other parameters are chosen as in Fig. 2.
1+λ2
of major to minor axis. The shape factor approaches
+1 and −1 for needle and disk-like particles, respectively.
Since microswimmers are usually elongated, we focus on similar to the ones of passive spheroids, albeit with a
prolate spheroids of Λ > 0.9, which corresponds to λ > 4; larger aspect ratio (see Fig. 2b) since they exhibit more
it also helps in simplifying the calculations (see ‘Meth- skewed trajectories. This can directly be inferred from
ods: Hydrodynamic model’). The terms with coefficients Eq. (1), where activity (σ) in the term with Wi σα1 mod-
αi and βi represent the active and passive viscoelastic ifies the shape factor. Since α1 > 0, a pusher (σ > 0) effec-
contributions, respectively. These coefficients are evalu- tively increases Λ, which corresponds to a higher aspect
ated explicitly as outlined in ‘Methods: Orientation dy- ratio. The opposite occurs for weak pullers (σ < 0) as
namics’ using the Lorentz reciprocal theorem, where we they behave like a passive spheroid with reduced aspect
also derive Eq. (1). For our relevant parameters, we find ratio. Here the trajectories are more circular as shown in
α1 ≫ ∣α2 ∣. Since the coefficients vary only weakly with Fig. 2d.
λ and δ, they are treated as constants in the following As the dipole strength of a pusher increases beyond a
discussion of results (see also Supplementary Note 2A). critical value σc , the orientation drifts to the shear plane
In the absence of polymers (Wi = 0), Eq. (1) reduces (θ = π/2) and aligns at an angle φeq with the flow di-
to the well-known Jeffery equation [25] that has an in- rection (see Fig. 2c). We quantify this transition in Fig.
finite number of neutrally stable solutions, i.e., the mi- 3 and show that σc (dots close to 0) decreases with in-
croswimmer’s orientation traces periodic orbits that de- creasing Wi while φeq increases with both σ and Wi.
pend on initial conditions. During orbital time period This dynamical behaviour can again be discerned from
T = 2π(λ + λ−1 )/γ̇, they spent the majority of their time Eq. (1), which essentially balances the effect of rotational
(∝ λ) aligned near the flow-vorticity plane. When poly- (Ω∞ ) and elongational (E∞ ) flow on p. Now, activity
mers are present, the passive viscoelastic effect breaks together with viscoelasticity allows to control this bal-
the degeneracy of Jeffery orbits and the microswimmer ance. In particular, for our case of α1 ≫ ∣α2 ∣, we expect
slowly drifts towards an alignment with the vorticity axis, the elongational flow to dominate the dynamics for in-
known as the “log-rolling” state [37, 44, 45]. Figure 2a creasing activity. Indeed, when the effective shape factor
shows the curve traced by the orientation of a spheroid Λ+Wi σα1 exceeds one at a critical value σc , the dynamics
on a unit sphere while drifting towards the vorticity axis. of p transitions from the orbital to the alignment state as
This dynamics can be deduced from Eq. (1) for passive favored by the elongational flow. In ‘Methods: Dynami-
spheroids (σ = 0) when the terms with coefficients β1 cal analysis 1’, we show for a modified dynamical system
and β2 are present. In Supplementary Note 2C we also that the relevant eigenvalue of the dynamical matrix be-
compare our results of passive spheroids with Brunn [37]. comes real at σc , as expected for such a transition, and
Now, we illustrate the influence of activity. For weak the corresponding eigenvector gives the alignment angle
pushers (0 < σ ≲ 0.1), the orbits towards log rolling look φeq . As σ increases, φeq grows from zero and approaches
4

π/4 for large σ, since in this limit, the elongational flow where ψ(p) is normalized to one. The first term de-
(E∞ ) with its principle axis along φ = π/4 completely scribes the orientational drift using ∇p as the gradient
dominates the dynamics of p. operator on the unit sphere. To compare the strength of
For pullers, Fig. 2e shows that increasing the activity flow-induced reorientation to the mean time τ between
induces a transition from log rolling towards rotation in two tumbling events, we introduce the flow Péclet num-
the shear plane. This orbit is also observed for passive ber Pef = γ̇τ . Tumbling away from the swimming di-
oblate particles in viscoelastic flows [37, 44, 45] and acts rection p is handled by the third term and rotational
as an attracting limit cycle. We have performed a stabil- diffusion by the second term. For bacteria, the latter is
ity analysis for the shear-plane rotation in ‘Methods: Dy- typically small compared to tumbling since τ ∼ 1s [46]
namical analysis 2’ and calculated the Lyapunov stability and Dr ≲ 0.1s−1 , as an estimate from the Stokes-Einstein
exponent L. It determines the exponential time variation relation shows. In restricting ourselves to Eq. (2), we
of the disturbed limit cycle. In the inset of Fig. 3, we assume a spatially uniform system valid when hydrody-
show the Lyapunov exponent as a function of σ. For weak namic and steric interactions with bounding surfaces can
and moderate pullers, L > 0 shows that shear-plane ro- be neglected [29, 40]. In particular, this means that the
tation is an unstable limit cycle, where the system drifts mean length of persistent swimming, Us τ , where Us is the
towards stable log rolling. As the dipole strength of the swimming speed, is much smaller than the spatial extent
puller becomes more negative, L turns negative mean- of the system. Finally, we consider weak fluid elastic-
ing that shear-plane rotation is stable. Upon analysing ity with relaxation time trelax ≪ τ, Dr−1 , so that the fluid
the orbital dynamics in the shear-plane rotation, we find relaxes faster than the time scales given by stochastic
that increasing ∣σ∣ gradually slows down the swimmer’s reorientations [51]. Thus, we do not need to take into
rotation near the y-axis, along which the flow gradient is account any memory in the rotational noise.
applied. Eventually, a transition to permanent alignment
with φeq = π/2 occurs, which can be similarly analyzed as In the following, we want to emulate a rheological ex-
the alignment transition of pushers. However, here it oc- periment and vary the shear rate γ̇ via Pef , while the
curs at the effective shape factor Λ + Wi σc α1 = −1. With fluid and swimmer properties are kept constant. There-
further increasing ∣σ∣ the active spheroid tilts against the fore, in Eq. (1) for the rotational drift velocity ṗ, we
flow, in contrast to pushers, and asymptotes at 3π/4, rewrite Wi as Pef De, where the Deborah number De =
which is the direction of the second principal axis of the trelax /τ compares the fluid relaxation time to τ . Further-
elongational flow (E∞ ). This behavior is illustrated in more, the second relevant parameter, Wiσ, does not ex-
Fig. 2f and Fig. 3. Finally, Fig. 3 also shows that as Wi plicitly depend on γ̇ but rather quantifies the strength
increases, the regime of shear-plane rotation shrinks. of activity relative to fluid elasticity. To vary activ-
ity independent of other parameters, we replace Wiσ by
Impact of noise on orientational dynamics. The Pea De/8π. Here, Pea is the signed activity Peclet num-
deterministic orientational dynamics is disturbed by two ber that is positive for pushers and negative for pullers.
types of stochastic reorientations of the swimming direc- Using Us ∼ σ ∗ /(µf l2 ) for the swimming speed with the
tion, which we now address with the help of the Smolu- force dipole moment σ ∗ [41], we can write it in the familiar
chowski equation. First, a bacterium tumbles, which is form Pea ∼ Us τ /l. Thus, its magnitude compares the per-
triggered when the rotation of one of its flagella reverses sistence length to the body length l. A wild-type E.coli
so that it leaves the flagellar bundle [46–48]. For a wild typically has Pea ≲ 5 [46, 52]. We numerically solve Eq.
type E.coli, tumbling occurs roughly every 1s, where it (2) for arbitrary P ef by expanding Ψ in spherical har-
attains a new random orientation. Although tumbling monics and taking into account the first 100 harmonics.
is biased in the forward direction with a mean tumbling The numerical solution is also verified analytically in the
angle of roughly 68.5○ , we model it to be unbiased for limit of Pef ≪ 1 (see further details in ‘Methods: Kinetic
computational ease because it only affects our results model’).
marginally as suggested by Nambiar et al. [49]. Second,
thermal rotational diffusion continuously reorients a mi- Figure 4 shows the orientational probability distribu-
croswimmer but its effect is small compared to tumbling. tion of passive and active particles for various cases. It
We now evaluate the orientational distribution of an quantifies the orientational ‘microstructure’ of a suspen-
ensemble of non-interacting microswimmers in steady sion of non-interacting and orientable particles subject
state, caused by the three mechanisms discussed so far: to shear flow [49, 53, 54]. We begin with discussing the
deterministic motion in background shear flow, tum- case of weak shear rate (Pef < 1) in Figs. 4a-d. In the ab-
bling, and rotational diffusion. The steady-state proba- sence of shear flow, the microstructure is solely governed
bility distribution function ψ(p) for the orientation vec- by rotational noise and is therefore isotropic. For weak
tor of non-interacting microswimmers is governed by the shear rates the extensional part (E∞ ) of the shear flow
Smoluchowski equation [50] primarily distorts the microstructure, which peaks near
the extensional axis as noted for Newtonian fluids in Ref.
1 [55]. We quantify this in ‘Methods: Kinetic model 1’ by
Pef ∇p ⋅ (ṗψ) − τ Dr ∇2p ψ + (ψ − ) = 0, (2) solving Eq. (2) via a perturbation expansion in Pef and

5

obtain the first-order correction: 𝑦 𝑦


(a) (b) 0.128
3 E∞ ∶ pp Passive Pusher
ψ(1) = (Λ + α1 DePea /8π) (3) 0.113
4π 1 + 6τ Dr
0.098
In the absence of activity, there is no deviation from the 𝑥 𝑥
𝑧 𝑧 0.083
Newtonian microstructure at this order.
In Figs. 4a-d we use a weak shear rate of Pef = 0.5 0.068
and ∣Pea ∣ = 20, which is an activity close to mutated (c) 𝑦 (d)
strains of tumbling E. coli (pusher) [56] and the algae C. 0.14
Puller Pusher
reinhardtii (puller). The observed distributions follow 0.12 Passive

PDF
the trend of Eq. (3). While pushers (Pea > 0) enhance 0.10 Puller
the alignment along the extensional axis of the applied 0.08
𝑥 Newtonian
shear flow, pullers (Pea < 0) weaken it. We note that the 𝑧 0.06
0 𝜋 π
peak of the orientational distributions in the shear plane 4 𝜙
2
π
(plot d) exhibits a small deviation from φ = π/4. This
is due to the rotational flow (Ω∞ ), which contributes to 𝑦 𝑦
(e) 𝜙 (f) 0.45
ψ(p) in second order in Pef [54]. Passive Pusher
As the shear rate increases, the deterministic dynamics 𝑡 0.35
becomes more visible. We illustrate this in Figs. 4e-h for
0.25
Pef = 5. Compared to Fig. 4a, the peak in ψ(p) for pas-
sive particles shifts towards the flow axis and becomes 𝑥 𝑥
𝑧 𝑧 0.15
more anisotropic along the flow-vorticity plane, as the
0.05
deterministic velocity ṗ is smallest in this plane. Eventu-
(g) 𝑦 (h)
ally, ṗ becomes zero in the deterministic log-rolling state 𝜙 0.30
0.25
observed in viscoelastic fluids (Fig. 2a). However, this 1.2 0.2
0.20
Puller 𝑡 0.15
1.0 0.10
state is not completely reflected in the distribution be- 00 𝜃
π
𝜋π
0.8

PDF
2

cause the slow relaxation towards the log-rolling axis is Pusher


0.6
always interrupted by the dominating stochastic reorien- Passive
𝑥 0.4 Newtonian
tations [51, 57, 58]. Note that due to the fluid elasticity, 𝑧 0.2
Puller

the peak of ψ(p) is slightly closer to the flow axis com- 0 𝜋 π

pared to the Newtonian case as Fig. 4h shows. Further- 4 𝜙


2
π

more, the inset therein also illustrates a broader distri-


bution in the flow-vorticity plane (φ = 0). These findings (i) 𝑦 (j) 𝑦
𝜙 0.22
Pusher Puller
are consistent with earlier studies on passive fibre suspen- No Tumble
sions at moderate shear rates [51, 57]. Further increas-
No Tumble 𝑡 0.18

ing Pef spreads the distribution even more in the flow- 𝑥 0.14
vorticity plane and, ultimately, for Pef ≳ 100 the peak
develops along the vorticity axis, which is entirely rem- 𝑥 0.10
𝑧
𝑧
iniscent of log rolling. However, this is a regime which 0.06
we cannot strictly capture as our analysis requires De,
Wi < 1, which means Pef < 10.
Figure 4f demonstrates that the activity of pushers in Fig. 4. Orientational probability distribution ψ(p) for
conjunction with the higher shear rate at Pef = 5 makes different parameters: (a-d) weak shear flow at Pef = 0.5, (e-
the distribution more anisotropic and, according to Fig. j) strong shear flow at Pef = 5. (a,e) Pea = 0, (b,f) Pea =
4h, strongly focused at an angle φ = 18○ . This is close to 20, (c,g) Pea = −20 and always with Dr τ = 0.1, (d,h) ψ(p)
the alignment angle φeq = 21○ of the deterministic align- in the shear plane; the inset in (h) shows ψ(p) in the flow-
ment state with the parameters Wi = 0.5, σ = 0.16 in vorticity plane. Non-tumbling particles: (i) Pea = 100, (j)
Fig. 3. Conversely, pullers reduce the anisotropy in the Pea = −100. The insets in (e,g,j) show φ(t) for the shear-plane
rotation state corresponding to the respective distributions.
orientational distribution as shown in Fig. 4g. Similar to
Other parameters are chosen as in Fig. 2 with De = 0.1. To
the passive case, we can directly relate this observation locate these distributions in Fig. 3, we give the corresponding
to the deterministic velocity ṗ with parameters Wi = 0.5, Weissenberg and activity numbers: (a-d) Wi = 0.05, σ = ±1.6;
σ = −0.16 that indicate again the log-rolling state; the (e-h) Wi = 0.5, σ = ±0.16; (i,j) Wi = 0.5, σ = ±0.8.
resulting orbit is similar to that depicted in Fig. 2d.
Compared to the passive case, the orbit is more circular,
which we already related to a reduced effective aspect ra-
tio. Therefore, the orientation is less constrained to the [59]. Consequently, the particle is less susceptible to the
flow-vorticity plane (see also the inset of Fig. 4g) and, as effect of stochastic reorientations and thus, its distribu-
a result, the distribution is broader and less anisotropic tion has a peak closer to the flow axis. Hence, unlike
6

active Newtonian suspensions [24], the microstructure of by Einstein [63] for spherical particles, and then later
microswimmers in a viscoelastic fluid is clearly sensitive generalized to elongated particles by Hinch and Leal
to their hydrodynamic signature being either pushers or [38, 53]. For the second-order fluid, we follow Férec
pullers. et al. [64] and approximate it using the Geisekus form,
We also examined the orientational distributions of mi- ∇

croswimmers that do not tumble and only experience ro- ΣF 3


p = −µf nl Aγ̇ ⟨pp⟩/2, where A = π/6 ln(2λ) is a shape
tational thermal noise. Therefore, in this case we elimi- factor to access the friction of a long slender body

nate the last term on the right-hand side of Eq. (2) and and ⟨pp⟩ is the upper-convected time derivative of
redefine our parameters Pef , Pea , and De replacing τ by the second moment of ψ(p). The expression for ΠF
Dr−1 . For passive particles and activity Pea ∼ O(10), the was originally derived for dumbbells suspended in a
microstructure is qualitatively similar to the distributions Newtonian fluid [65]. As in Ref. [64] we use it here as
in Figs. 4a-h. However, microswimmers that do not tum- an approximation for the stress response of particles in
ble can have larger persistence lengths [60]. In particular, a weakly viscoelastic fluid, as there are no closed form
one can genetically modify E. Coli such that tumbling expressions available. The consequences of the second-
does not occur [41, 56, 61]. Therefore, in Figs. 3i and j order fluid come in through the orientational dynamics
we show the orientational distributions for Pea = ±100. ṗ in Eq. (1) and further terms in the upper-convected
Compared to Fig. 4f, increasing the activity of a pusher derivative (see ‘Methods: Rheology’). The expressions
moves the peak in the distribution function (at φ = 34○ ) for the three contributions to the particle stress tensor
even closer to the alignment angle (φeq = 36○ ) of the de- are pertinent to studies on slender rods and fibres (i.e.,
terministic alignment state. For pullers we observe an Λ → 1), and henceforth, we will be focusing on spheroids
even more drastic change upon increasing the activity. with larger aspect ratios (λ = 10). Using Eq. (4) in the
The peak in the orientational distribution shifts to the definition of the particle-induced contribution to shear
quadrant where the compressional part of the shear flow viscosity, ϕµp = Σp xy /γ̇, together with our characteristic
occurs (see Fig. 4j). Interestingly, this is not due to parameters, we obtain
the deterministic alignment of pullers as the parameters
(Wi = 0.5, σ = −0.8) belong to the shear-plane rotation µp ∇ 8 Pea I
state (see Fig. 3). As already explained in the discus- = −4A⟨pp⟩xy + (6τ Dr A − ) [⟨pp⟩ − ] .
µf Pef 8π 3 xy
sion of Fig. 3, the rotational velocity ṗ slows down near
(5)
φ = π/2 when the effective shape factor Λ + Wi σc α1 be-
We use numerical solutions of Eq. (2) for the orienta-
comes negative. Therefore, the active particle resembles
tional distribution ψ(p) to evaluate µp for varying Pef
a passive oblate spheroid [62]. Indeed, the inset of Fig. 4j
and Pea . We also match the numerical values of µp with
shows how the puller orientation spends more time near
an expression, which we obtain using the analytic form of
φ = π/2. This again shows that even in the presence of
ψ(p) from Eq. (3) in the limit of small Pef (the derivation
significant noise the microstructure is determined by the
is provided in ‘Methods: Rheology’). In Supplementary
deterministic dynamics.
Note 3, we also show that our results agree with Saintil-
Shear viscosity of active suspensions. Finally, we
lan [24] in the Newtonian limit.
evaluate the effective shear viscosity of a dilute active
Figure 5a shows the particle-induced contribution to
suspension from the total stress tensor, Σ = Σf + Σp ,
shear viscosity normalized by the bare fluid viscosity µf ,
where subscripts f and p refer to fluid and particle con-
tributions, respectively. The deviatoric component of which is plotted versus shear strength Pef . We begin
Σ provides the shear viscosity, µ = µf + ϕµp , where with discussing suspensions in a Newtonian fluid (dashed
ϕµp = Σp xy /γ̇ is due to the suspended microswimmers lines) [24]. When a suspension of passive rods (black) is
and ϕ = na3 is proportional to their volume fraction with sheared weakly (Pef ≪ 1), the orientational microstruc-
n being the particle density. Note that the shear vis- ture is governed by Eq. (3) with De = 0 and resembles
cosity µf of the second-order fluid does not depend on Fig. 4a. Here, the rods are aligned along the exten-
the shear rate, while µp is dependent on it. To evaluate sional axis of the applied shear flow and resist shearing
µp , we need to average over the stresslets Π(p) gener- (µp > 0), which enhances viscosity. For a suspension of
ated by the suspended particles. Using the orientational pushers (brown dashed) in the weak-shear regime, the
distribution, we obtain microstructure still resembles Fig. 4a, but now the ex-
tensile force dipoles support the elongational part of the
Σp = n⟨Π⟩ = n ∫ Π(p)ψ(p)dp, (4) shear flow, which results in µp < 0 so that the total shear
S viscosity is smaller than µf . Conversely, pullers (green
where Π(p) is the sum of three contribu- dashed) with their contractile force dipoles additionally
tions [24], which give the following stress ten- resist shear flow and thereby enhance viscosity. Note
that, since the microstructure of active suspensions in a
sors: ΣA
p = nσ ∗ [⟨pp⟩ − I/3] /8π due to activity,
T
Newtonian fluid is unchanged by the hydrodynamic sig-
Σp = 3nkB T [⟨pp⟩ − I/3] due to thermal reorientations, nature of microswimmers, the magnitude of the activity
and ΣF p due to the resistance of passive particles to contribution to µp is identical for pushers and pullers.
shear. For Newtonian fluids, the latter was first derived As the shear rate increases, the magnitude of µp for both
7

passive and active rods reduces, which resembles a typical (a) Pea = 20 (b) Pea = 20
<latexit sha1_base64="bTZz+WwKwB716ACCbRZGxBW+yj0=">AAACCnicbVBNS8NAEN34WetX1aOXaBG8WJJS1ItQ9OKxgv2ApoTNdtIu3WzC7kQsoWcv/hUvHhTx6i/w5r8xaXuorQ8GHu/NMDPPiwTXaFk/xtLyyuraem4jv7m1vbNb2Ntv6DBWDOosFKFqeVSD4BLqyFFAK1JAA09A0xvcZH7zAZTmobzHYQSdgPYk9zmjmEpu4chBeMSkBiOXXpUtx8nPCmdlK+8WilbJGsNcJPaUFMkUNbfw7XRDFgcgkQmqddu2IuwkVCFnAkZ5J9YQUTagPWinVNIAdCcZvzIyT1Kla/qhSkuiOVZnJxIaaD0MvLQzoNjX814m/ue1Y/QvOwmXUYwg2WSRHwsTQzPLxexyBQzFMCWUKZ7earI+VZRhml4Wgj3/8iJplEv2ealyVylWr6dx5MghOSanxCYXpEpuSY3UCSNP5IW8kXfj2Xg1PozPSeuSMZ05IH9gfP0CPWuZUQ==</latexit>
<latexit sha1_base64="Xw+5DnJqIYknP7knMTGDYnLT/qQ=">AAAB+HicbVBNS8NAEN3Ur1o/GvXoJVgETyUpRb0IRS8eK9gPaEPYbKft0s0HuxOxhv4SLx4U8epP8ea/cdPmoK0PBh7vzTAzz48FV2jb30ZhbX1jc6u4XdrZ3dsvmweHbRUlkkGLRSKSXZ8qEDyEFnIU0I0l0MAX0PEnN5nfeQCpeBTe4zQGN6CjkA85o6glzyz3ER4xbcLMo1c1u+SZFbtqz2GtEicnFZKj6Zlf/UHEkgBCZIIq1XPsGN2USuRMwKzUTxTElE3oCHqahjQA5abzw2fWqVYG1jCSukK05urviZQGSk0DX3cGFMdq2cvE/7xegsNLN+VhnCCEbLFomAgLIytLwRpwCQzFVBPKJNe3WmxMJWWos8pCcJZfXiXtWtU5r9bv6pXGdR5HkRyTE3JGHHJBGuSWNEmLMJKQZ/JK3own48V4Nz4WrQUjnzkif2B8/gDCopKB</latexit>

shear-thinning/thickening behaviour. For passive parti- Pea = 20


Newt. Puller
cles the microstructure is similar to Fig. 4e. The align- 0
Passive
ment close to the flow axis explains why the resistance of 𝜇! Flow
0.4 - 0.5
particles to the applied shear flow is reduced. For active 0.2 𝜇" Thermal

particles, the activity-induced flow becomes less and less- 0.20 Pea =-20 Active
Newt. Pusher
-1
important with increasing shear rate as quantified by the- 0.4 Peaa =20
Pe 20
<latexit sha1_base64="bTZz+WwKwB716ACCbRZGxBW+yj0=">AAACCnicbVBNS8NAEN34WetX1aOXaBG8WJJS1ItQ9OKxgv2ApoTNdtIu3WzC7kQsoWcv/hUvHhTx6i/w5r8xaXuorQ8GHu/NMDPPiwTXaFk/xtLyyuraem4jv7m1vbNb2Ntv6DBWDOosFKFqeVSD4BLqyFFAK1JAA09A0xvcZH7zAZTmobzHYQSdgPYk9zmjmEpu4chBeMSkBiOXXpUtx8nPCmdlK+8WilbJGsNcJPaUFMkUNbfw7XRDFgcgkQmqddu2IuwkVCFnAkZ5J9YQUTagPWinVNIAdCcZvzIyT1Kla/qhSkuiOVZnJxIaaD0MvLQzoNjX814m/ue1Y/QvOwmXUYwg2WSRHwsTQzPLxexyBQzFMCWUKZ7earI+VZRhml4Wgj3/8iJplEv2ealyVylWr6dx5MghOSanxCYXpEpuSY3UCSNP5IW8kXfj2Xg1PozPSeuSMZ05IH9gfP0CPWuZUQ==</latexit>

- 0.6 Pea = 20
inverse proportionality to Pef in Eq. (5). 0.1 1 10
0.1 0.5 1 5 10
Pe Pe ff
Pe
<latexit sha1_base64="9PH/Jr7+HnYtOooQHn0fOkV09xg=">AAAB8nicbVBNS8NAEN3Ur1q/qh69LBbBU0mkqMeiF48V7Ae0oWy2k3bpZhN2J2IJ/RlePCji1V/jzX/jts1BWx8MPN6bYWZekEhh0HW/ncLa+sbmVnG7tLO7t39QPjxqmTjVHJo8lrHuBMyAFAqaKFBCJ9HAokBCOxjfzvz2I2gjYvWAkwT8iA2VCAVnaKVuD+EJswZM+2G/XHGr7hx0lXg5qZAcjX75qzeIeRqBQi6ZMV3PTdDPmEbBJUxLvdRAwviYDaFrqWIRGD+bnzylZ1YZ0DDWthTSufp7ImORMZMosJ0Rw5FZ9mbif143xfDaz4RKUgTFF4vCVFKM6ex/OhAaOMqJJYxrYW+lfMQ042hTKtkQvOWXV0nroupdVmv3tUr9Jo+jSE7IKTknHrkidXJHGqRJOInJM3klbw46L86787FoLTj5zDH5A+fzB6rQkYQ=</latexit> <latexit sha1_base64="9PH/Jr7+HnYtOooQHn0fOkV09xg=">AAAB8nicbVBNS8NAEN3Ur1q/qh69LBbBU0mkqMeiF48V7Ae0oWy2k3bpZhN2J2IJ/RlePCji1V/jzX/jts1BWx8MPN6bYWZekEhh0HW/ncLa+sbmVnG7tLO7t39QPjxqmTjVHJo8lrHuBMyAFAqaKFBCJ9HAokBCOxjfzvz2I2gjYvWAkwT8iA2VCAVnaKVuD+EJswZM+2G/XHGr7hx0lXg5qZAcjX75qzeIeRqBQi6ZMV3PTdDPmEbBJUxLvdRAwviYDaFrqWIRGD+bnzylZ1YZ0DDWthTSufp7ImORMZMosJ0Rw5FZ9mbif143xfDaz4RKUgTFF4vCVFKM6ex/OhAaOMqJJYxrYW+lfMQ042hTKtkQvOWXV0nroupdVmv3tUr9Jo+jSE7IKTknHrkidXJHGqRJOInJM3klbw46L86787FoLTj5zDH5A+fzB6rQkYQ=</latexit>

Now, we discuss the results for the second-order fluid Pef |Pea | = 20 f
<latexit sha1_base64="iZiCXOrFhYK1E3r07PDdat8RYEQ=">AAACDnicbVDJSgNBEO2JW4xb1KOXxhDwFGYkqBch6MVjBLNAJoSeTk3S2LPQXSOGSb7Ai7/ixYMiXj1782/sLKAmPih4vFdFVT0vlkKjbX9ZmaXlldW17HpuY3Nreye/u1fXUaI41HgkI9X0mAYpQqihQAnNWAELPAkN7/Zy7DfuQGkRhTc4iKEdsF4ofMEZGqmTLw5dhHtMqzDqsOH5sU1dN/cj+dSVkjqdfMEu2RPQReLMSIHMUO3kP91uxJMAQuSSad1y7BjbKVMouIRRzk00xIzfsh60DA1ZALqdTt4Z0aJRutSPlKkQ6UT9PZGyQOtB4JnOgGFfz3tj8T+vlaB/1k5FGCcIIZ8u8hNJMaLjbGhXKOAoB4YwroS5lfI+U4yjSTBnQnDmX14k9eOSc1IqX5cLlYtZHFlyQA7JEXHIKamQK1IlNcLJA3kiL+TVerSerTfrfdqasWYz++QPrI9v8oObZQ==</latexit>

(SOF) in Fig. 5a (solid lines). For the suspension of (c) No Tumble (d) No Tumble: |Pea | = 100
<latexit sha1_base64="6eDU1rJeW3aHk2ve6FuP/JMes/g=">AAAB+nicbVBNS8NAEN3Ur1q/Uj16CRbBU0lE1ItQ9OKxgv2ANoTNdtIu3XywO1FL2p/ixYMiXv0l3vw3btsctPXBwOO9GWbm+YngCm372yisrK6tbxQ3S1vbO7t7Znm/qeJUMmiwWMSy7VMFgkfQQI4C2okEGvoCWv7wZuq3HkAqHkf3OErADWk/4gFnFLXkmeVxF+EJszpMPDq+cmzbMyt21Z7BWiZOTiokR90zv7q9mKUhRMgEVarj2Am6GZXImYBJqZsqSCgb0j50NI1oCMrNZqdPrGOt9KwglroitGbq74mMhkqNQl93hhQHatGbiv95nRSDSzfjUZIiRGy+KEiFhbE1zcHqcQkMxUgTyiTXt1psQCVlqNMq6RCcxZeXSfO06pxXz+7OKrXrPI4iOSRH5IQ45ILUyC2pkwZh5JE8k1fyZoyNF+Pd+Ji3Fox85oD8gfH5A8swk7I=</latexit>

Pef ⌧ 1
passive rods (black), we do not find an appreciable de- Newt.
VE Puller
viation from the Newtonian case until Pef ≈ 10, similar Newt. Puller
VE Pusher
𝜇! Newt. Pusher
to Refs. [51, 57]. However, for active suspensions the 0.4
𝜇" 0.2
modifications are substantial and qualitatively different Newt.
0 Pea =-20
VE Puller
for pushers (brown) and pullers (green). For pushers, VE Pusher
- 0.2 Puller
- 0.4 Pusher
Pea =20
the reduction in viscosity is further enhanced compared - 0.6
to Newtonian fluids, whereas the viscosity enhancement Pea 0.1 0.5 1 5 Pe
10
f
<latexit sha1_base64="TrZq5lDQotV2Bm76pCEL/gElgXQ=">AAAB8nicbVBNS8NAEN3Ur1q/qh69LBbBU0mkqMeiF48V7Ae0oWy2k3bpZhN2J2IJ/RlePCji1V/jzX/jts1BWx8MPN6bYWZekEhh0HW/ncLa+sbmVnG7tLO7t39QPjxqmTjVHJo8lrHuBMyAFAqaKFBCJ9HAokBCOxjfzvz2I2gjYvWAkwT8iA2VCAVnaKVuD+EJswZM+6xfrrhVdw66SrycVEiORr/81RvEPI1AIZfMmK7nJuhnTKPgEqalXmogYXzMhtC1VLEIjJ/NT57SM6sMaBhrWwrpXP09kbHImEkU2M6I4cgsezPxP6+bYnjtZ0IlKYLii0VhKinGdPY/HQgNHOXEEsa1sLdSPmKacbQplWwI3vLLq6R1UfUuq7X7WqV+k8dRJCfklJwTj1yROrkjDdIknMTkmbySNwedF+fd+Vi0Fpx85pj8gfP5A6M8kX8=</latexit>

<latexit sha1_base64="9PH/Jr7+HnYtOooQHn0fOkV09xg=">AAAB8nicbVBNS8NAEN3Ur1q/qh69LBbBU0mkqMeiF48V7Ae0oWy2k3bpZhN2J2IJ/RlePCji1V/jzX/jts1BWx8MPN6bYWZekEhh0HW/ncLa+sbmVnG7tLO7t39QPjxqmTjVHJo8lrHuBMyAFAqaKFBCJ9HAokBCOxjfzvz2I2gjYvWAkwT8iA2VCAVnaKVuD+EJswZM+2G/XHGr7hx0lXg5qZAcjX75qzeIeRqBQi6ZMV3PTdDPmEbBJUxLvdRAwviYDaFrqWIRGD+bnzylZ1YZ0DDWthTSufp7ImORMZMosJ0Rw5FZ9mbif143xfDaz4RKUgTFF4vCVFKM6ex/OhAaOMqJJYxrYW+lfMQ042hTKtkQvOWXV0nroupdVmv3tUr9Jo+jSE7IKTknHrkidXJHGqRJOInJM3klbw46L86787FoLTj5zDH5A+fzB6rQkYQ=</latexit>

of pullers is reduced. We understand this using the ori- Pef


entational distributions in Fig. 4. Pushers are even more Fig. 5. Particle-induced contribution to viscosity
aligned in the extensional quadrant of the applied shear µp /µf in a second-order fluid. (a) Plotted vs. shear rate
flow (Fig. 4b,d,f), which explains their further increased Pef for pusher (Pea > 0), puller (Pea < 0), and a passive
support of shear flow. In contrast, pullers are less aligned rod (Pea = 0) with solid lines. Dashed lines correspond to
compared to the Newtonian case (Fig. 4c,d,g) and, there- the Newtonian case. (b) Contributions to particle-induced
fore, oppose shear flow less. Figure 5b for pushers clearly viscosity for Pea = 20. (c) µp /µf vs. activity Pea in the
A limit Pef → 0 for tumbling and non-tumbling active rods.
shows that the active stress (Σ ) dominates and hence is
(d) µp /µf vs. shear rate Pef for non-tumbling active rods
primarily responsible for the observed viscous response. with large persistence length, ∣Pea ∣ = 100. Dotted black lines
In the limit of shear rate much smaller than tumbling near Pef = 0.1 in (a,d) correspond to the analytical results for
rate, Pef ≪1 , we can calculate µp /µf analytically using Pef ≪ 1. Other parameters: De = 0.1, λ = 10, α1 = 5.92, α2 =
Eq. (3) (see ‘Methods: Rheology’) and obtain: −0.91 β1 = 0.63, β2 = 1.18.

µp 4A(5 − 3Λ) 48Aτ Dr − Pea /π De Pea α1


= + (Λ + ) 5c, dashed lines). Non-tumbling microswimmers can ex-
µf 15 5 (1 + 6τ Dr ) 8π
De Pea Aα1 hibit high persistence lengths. Thus, in Fig. 5d we show
− . (6) for ∣Pea ∣ = 100 that their contribution to viscosity is neg-
10π ative over a wide range of shear rates for both pushers
We discuss the result here since the influence of passive and pullers in a second-order fluid. Hence, we find that
and active rods on the shear viscosity is the strongest in the elasticity of a fluid always results in a reduced to-
this limit. In Fig. 5c we plot µp /µf vs. Pea for tumbling tal viscosity as compared to suspensions in a Newtonian
microswimmers (solid lines). As Eq. (6) shows, elasticity fluid.
in the fluid adds a quadratic dependence in Pea , while in
a Newtonian fluid the dependence is only linear [24]. For
all Pea , except a small region close to Pea = 0, the values DISCUSSION
of µp /µf remain below those obtained in the Newtonian
limit. Thus, fluid elasticity reduces the shear viscosity In this article we show how activity influences the dy-
for both pusher and puller suspensions. Note that, as namics of a sheared microswimmer suspension in a vis-
we increase the activity of pullers (Pea ≲ −50), they too coelastic fluid at an individual level and in the bulk. At
can reduce the shear viscosity (µp /µf < 0) like pushers. the individual level, the orientational dynamics of pas-
This is a consequence of the orientational distribution sive rods is well known from Jeffery’s orbits in Newtonian
shown in Fig. 4j; pullers with large activity align prefer- shear [25] and ‘log-rolling’ orbits in viscoelastic shear flow
entially in the compressional quadrant of the shear flow [36, 37]. Our analytical result [Eq. (1)], derived for elon-
and thereby also support the shearing fluid similar to gated active spheroids in a second-order fluid, demon-
pushers aligned in the extensional quadrant. strates how the active flow field of a swimmer modifies the
As described earlier, we can treat the case of non- orientational dynamics. Extensile swimmers like E.coli
tumbling microswimmers by replacing τ by Dr−1 in the (pushers) drift to the shear plane and align at an angle
definitions of Pef , Pea , and De. With these parame- to the flow direction. For contractile swimmers like C.
ters, Eq. (6) can be formulated in the limit τ → ∞ to reinhardtii (pullers), activity effectively lowers their as-
show that the viscosity contribution of non-tumbling ac- pect ratio. With increasing dipole strength, pullers show
tive rods also follows a quadratic dependence in Pea (Fig. log rolling, transition to shear-plane rotation, and ulti-
8

mately align at an angle against the flow direction. The METHODS


latter occurs for strong pullers at dipole strengths rele-
vant to artificial swimmers with large propulsion speed Hydrodynamic model
[66, 67]. The hydrodynamics around the spheroidal particle is
To demonstrate the impact of the individual swim- governed by the mass and momentum conservation as
mer dynamics on the bulk rheological behaviour, we em- ∇ ⋅ V = 0, ∇ ⋅ T = 0. Here T is the total stress tensor de-
ploy the Smoluchowski equation to evaluate the orien- ∆
tational probability distribution of an ensemble of ac- fined as T = −P I + 2 E + Wi S [34], where S = 4E ⋅ E + 2δ E.
tive spheroids called the suspension microstructure [38]. The lower-convected derivative is defined as
Accounting for tumbling and rotary diffusion, we find ∆ ∂E
that the microstructure substantially depends on the E=[ + (V ⋅ ∇) E + E ⋅ ∇V + ∇V T ⋅ E] . (7)
hydrodynamic signature of the swimmer unlike suspen- ∂t
sions in Newtonian fluids [22, 24, 49]. Pushers align Since we consider a steady shear flow and a steadily active
more strongly in the extensional quadrant of the ap- swimmer, we can disregard the temporal derivative in the
plied shear flow, while the alignment of pullers is sig- above stress tensor. In particle frame of reference, the
nificantly weaker. This activity-specific microstructure boundary condition on its surface is the no-slip condition
significantly modifies the effective shear viscosity com- V = Ωp × r, where the rotational velocity Ωp is currently
pared to Newtonian fluids [21]. In particular, the viscos- unknown and will be evaluated by using the torque-free
ity reduction of a dilute pusher suspension is more pro- condition. We split the velocity field into the disturbance
nounced under viscoelastic shear flow, while the viscosity field v and background flow field V ∞ = E∞ ⋅r+Ω∞ ×r i.e.,
enhancement of pullers is weaker. Thus, the activity of V = v + V ∞ . The equations governing the disturbance
a microswimmer contributes in two ways to the rheology flow field are
of swimmer suspensions in a viscoelastic fluid; directly
through its active stresses and indirectly by coupling to ∇ ⋅ v = 0, −∇p + ∇2 v = −Wi (∇ ⋅ s), where
the elasticity of the fluid and thereby influencing the ori- ∆ ∆
s = 4(e ⋅ e + w) + 2 δ( e + w). (8)
entational microstructure. In particular, in the weak-
shear limit the particle-induced contribution to viscosity Here e is the rate of strain tensor for the disturbance
scales quadratically with the activity Pea . In total, we ∆ ∆
flow (∇v + ∇v † )/2, whereas w, e and w are the different
find that the elasticity of a second-order fluid always re- constituents of the polymeric tensor (s) due to the distur-
duces the total viscosity of the microswimmer suspension. bance flow field. Specifically, w is the interaction tensor
Especially for pushers, this can help to reach the regime defined as E∞ ⋅ e + e ⋅ E∞ , arising from the interaction
of superfluidity at lower volume fractions compared to ∆
Newtonian fluids. between background flow and disturbance field; and w is
Biologically relevant fluids are viscoelastic. In this ar- the lower-convected derivative of e and E∞ with respect
ticle we presented a first systematic study of the individ- to V ∞ and v, respectively:
ual dynamics and the bulk rheology of microswimmers ∆
suspended in such fluids. Earlier investigations on ac- w = V ∞ ⋅ ∇e + e ⋅ ∇V ∞ † + ∇V ∞ ⋅ e (9)
tive suspensions in quiescent [68] and vortical viscoelastic + v ⋅ ∇E∞ + E∞ ⋅ ∇v † + ∇v ⋅ E∞ .
flows [69] assumed conventional Newtonian Jeffery orbits.
The boundary conditions of the disturbance field at par-
In the light of our results, these studies on the collec-
ticle surface and far away are
tive behaviour need to be revisited. Our investigations
makes several assumptions to address the complexity of v = ∆Ω × r − E∞ ⋅ r at r ∈ S and v → 0 as r → ∞,
microbial flows, which offers ample opportunities for fu- (10)
ture developments. Biological fluids can be more complex respectively. Here ∆Ω is the difference in particle angular
and exhibit shear-thinning/-thickening properties or sev- velocity and background vorticity (Ωp − Ω∞ ).
eral characteristic relaxation times, which requires more For small values of Wi, the disturbance field variables
evolved modeling using, for example, the FENE-P or are expanded as f = f (0) +Wi f (1) +⋯. Here, f is a generic
Giesekus model. Furthermore, mucus besides being sig- field variable that represents velocity (v), pressure (p),
nificantly shear-thinning has relaxation times larger than and angular velocity (Ωp ). We substitute this expansion
1 second, which is of the order of the bacterial mean in the Eq. (8) and obtain the O(1) Stokes problem as
free tumble time [15, 28, 29]. Thus, noise becomes non-
Markovian and memory needs to be incorporated in the ∇ ⋅ v (0) = 0, −∇p(0) + ∇2 v (0) = 0 (11)
stochastic description, for example, within a generalized
Langevin equation [70]. with the boundary condition at particle surface:

v (0) = ∆Ω(0) × r − E∞ ⋅ r at r ∈ S, (12)

and a decaying condition at infinity (v (0) → 0). Using


the finite multipole expansion around the spheroid [39,
9

43, 71], we obtain the disturbance velocity for a passive with a decaying condition at infinity (v (1) → 0). Conven-
spheroid at O(1) as tionally, a solution to Eqs. (17) and (18) (i.e., v (1) ) is
(0)
sought, which, on the implementation of torque-free con-
vi = QR R R
ij,k εjkl {[A pl pm + B (δlm − pl pm )] ∆Ωm (1)
dition, reveals the modification of Jeffery orbits (Ωp ).
+ CR εlmn pm E∞
no po }
We employ the reciprocal theorem [71, 72] to avoid solv-
ing for the first order flow field and directly obtain the
∞ rotation velocity as
ij,k + χQij,llk ) {[A pjklm + B pjklm + C pjklm ] Elm
+ (QS Q S A S B S C

− CR (εjlm pk pm + εklm pj pm ) ∆Ωl }. (13) t


p ⋅T =∫
Ω(1) s(0) ∶ ∇v t dV. (19)
Vf

Here Q represent the spheroidal multipoles i.e. integral Here, superscript t refers to the test/auxiliary problem
representation of fundamental singularities spread on a where the particle rotates at a unit velocity Ωt = ei ,
line √extending from one foci to another (−c to c, where where i corresponds to either of the three Cartesian co-
c = λ2 − 1/λ); QR , QS represent rotlet and stresslet ordinates. The test torque is:
around the spheroid, respectively. Here χ = 8(λ12 −1) rep-
resents the strength of higher order quadrupolar field 64πc3 R
T t = R ⋅ Ωt , where R = [A pp + BR (I − pp)] (20)
(QQij,llk ). Since we focus on elongated particles, we ex-
3
clude this component for computational ease (λ = {5, 10} We solve Eq. (19) for three test field torques (along three
correspond to χ = {0.005, 0.001}). Following Einarsson Cartesian coordinates) and obtain the following system
et al. [43], we simplify these multipoles as
of equations: R† ⋅ Ωp = {I1 , I2 , I3 }, where † represents
(1)

transpose and Ii is the solution of volume integral (19)


QR 0 1
ij,k = (δik rj − δij rk ) J3 + (δij pk − δik pj ) J3 (14a)
for the test field in ith direction. We further simplify the
QSij,k = δjk ri J30 − δjk pi J31
− 3ri rj rk J50 (14b) above relation by noting that R is a symmetric matrix
(1)
1
+ 3(pi rj rk + pj ri rk + pk ri rj )J5 and ṗ(1) = Ωp × p:
− 3(ri pj pk + rj pi pk + rk pi pj )J52 + 3pi pj pk J53 . ṗ(1) = (R−1 ⋅ {I1 , I2 , I3 }) × p (21)
Here I and J represent various integrals defined as Eq. (21) is evaluated over a discretized angular grid,
where each point requires solving the volume integral in
c ξb Eq. (19). The polymeric stress therein (s(0) ) is given by
Iab = ∫ dξ; Jab = c2 Iab − Iab+2 . (15)
−c ∣r − ξp∣a Eq. (8) and is entirely dependent on O(1) disturbance
flow fields i.e., passive disturbance in Eq. (13) and active
In Eq. (13), the fourth order orientation tensors (pA , pB , disturbance in Eq. (16). We find that out of all active
pC ) and other coefficients (A, B, C) are described in the fields, only force-dipole (∼ r−2 ) provides a modification
Supplementary Note 1. at the current order of approximation; the rest decays in
At O(1), we add the activity using the far-field descrip- odd powers of distance (∼ r−3 ), and due to antisymmetry,
tions, which consists of a force-dipole (FD), source-dipole give zero contribution to the volume integral at O(W i).
(SD), rotlet-dipole (RD), and force-quadrupole (FQ) [40]:
Orientation dynamics
(0) −r 3r (r ⋅ p)2 We evaluate ṗ(1) analytically by noting that it stems
vF D = σF D ( 3 + ) (16a)
r r5 from the leading order polymeric stress [specifically s(0)
in Eq. (19)]; its form in Eq. (8) suggests that the modifi-
(0) 3rr p
vSD = σSD ( − I) 3 (16b) cation of a passive particle will be quadratic in the flow
r2 2r
gradient tensor, which can be decomposed in symmetric
(0) 3p ⋅ r
vRD = σRD [ 5 (p × r)] (16c) E∞ and antisymmetric O∞ (rotation-rate tensor) compo-
r
nents. Following Einarsson et al. [43], this can be written
(0) p 3 5(p ⋅ r)3 r in the general form as:
vF Q = σF Q { 3
− 5 [3(p ⋅ r)r + (p ⋅ r)2 p − ]}
r r r2
(1,P ) (1) (2) (3)
(16d) ṗi jk Elm + Kijklm Ejk Olm + Kijklm Ojk Olm ,
= Kijklm E∞ ∞ ∞ ∞ ∞ ∞

(22)
At O(Wi), the governing equation is where superscript P denotes passive contribution. The
coefficients of the fifth-order tensor K are composed of
∇ ⋅ v (1) = 0, −∇p(1) + ∇2 v (1) = − (∇ ⋅ s(0) ), where all possible permutations of the orientation vector p with
∆(0) ∆ (0) E∞ and O∞ :
s(0) = 4(e(0) ⋅ e(0) + w(0) ) + 2 δ( e +w ), (17)
(i)
Kijklm = ∑ (βn[1] pn1 pn2 pn3 pn4 pn5 + βn[2] pn1 pn2 pn3 δn4 n5
The boundary condition at the particle surface being n=5!

v (1) = Up(1) + Ω(1) +βn[3] pn1 δn2 n3 δn4 n5 ) .


p ×r at r ∈ S, (18)
10

Here β represents the unique coefficients for all the 5! The eigensystem of the above exponential matrix deter-
terms. mines the orbital dynamics of spheroid. For the two-
When activity is added in the hydrodynamics, the form dimensional shear flow, the eigenvalues√ of the matrix
of polymeric stress tensor in Eq. (8) suggests that the in the exponent (ΛE∞ + O∞ ) are: 0, ± Λ2 − 1/2. The
gradients of disturbance velocity (directed with p) will non-zero eigenvalues are purely imaginary as 0 < Λ < 1,
multiply with gradients of passive disturbance (which where Λ = 1 for an infinitely slender particle. As a con-
originate from E∞ and O∞ ) to yield further modification sequence of this imaginary pair, we observe degenerate
to ṗ(1) . Thus, the contribution resulting from interaction infinite solutions of orbits in Newtonian fluids. For the
of active disturbances (p) with background flow (E∞ and case of pure elongational flow (O∞ ≡ 0), the eigenvalues
O∞ ) takes the following form at O(Wi): are real: 0, ±Λ/2, which reveals the absence of periodic
(1,A) (1) (2)
orbits. The normalized eigenvector corresponding to the
ṗi = σ [Lijk E∞
jk + Lijk Ojk ] ,

(23) positive eigenvalue
√ √ determines the equilibrium orienta-
tion: {1/ 2, 1/ 2, 0} i.e., 45° or 225° in the two exten-
where coefficients of the third-order tensor L are com- sion quadrants.
posed of the following combinations
We now use Eq. (1) for the second-order fluid and
(i)
Lijk = ∑ (αn[1] pn1 pn2 pn3 + αn[2] pn1 δn2 n3 ) . consider only the active viscoelastic component because
n=5! the passive effects do not contribute to the alignment
[this assumption is later verified by matching the results
Here α represent the unique coefficients for all the 3!
with numerical integration of complete Eq. (1)]. In this
terms. We simplify Eq. (22) and (23) by using the sym-
case, the equation of motion for the unconserved vector
metry arguments and noting that magnitude of p is al-
q is
ways unity. We obtain the following six irreducible terms:
ṗ(1) = E∞ ∶ pp [β1 (I − pp) ⋅ (E∞ ⋅ p) + β2 O∞ ⋅ p] q̇ = [E∞ (Λ + Wiσα1 ) + O∞ (1 + Wiσα2 )] ⋅ q, (29)
+ (I − pp) ⋅ [β3 (E∞ ⋅ E∞ ) ⋅ p + β4 (O∞ ⋅ E∞ ) ⋅ p]
which yields the solution as
+ σα1 (I − pp) ⋅ (E∞ ⋅ p) + σα2 O∞ ⋅ p. (24)
We calculate the coefficients (αi and βi ) by calculating q(t) = exp [ [E∞ (Λ + Wiσα1 ) + O∞ (1 + Wiσα2 )] t]q(0).
ṗ(1) numerically for six independent orientations (p) and (30)
solve the system of equations from Eq. (24) to extract We obtain the eigenvalues of the fundamental matrix as
the coefficients. We find that β3 and β4 are nearly zero √
1
(≲ 10−5 ) and thus neglect them. We obtain the analytical 0, ± Λ2 − 1 + 2σWi (α1 Λ − α2 ) + Wi2 σ 2 (α12 − α22 ).
approximation as 2
1−Λ
ṗ(1) = (I − pp) ⋅ (E∞ ⋅ p) [σα1 + β1 E∞ ∶ pp] These eigenvalues turn real when σ > Wi(α1 −α2 )
or σ <
+ O∞ ⋅ p [σα2 + β2 E∞ ∶ pp] , (25) −1−Λ
Wi(α1 +α2 )
. For instance, for a particle of aspect ra-
tio λ = 5, the negative and positive critical values are
We use the above equation in the main text as Eq. (1),
σcrit = −0.448/Wi; 0.016/Wi, which matches with Fig. 3
where we write O∞ ⋅p as Ω∞ ×p. In Supplementary Note
generated from the integration of full Eq. (1) (see Sup-
2A, we show the weak variation of these coefficients with
plementary Note 2B). In the alignment regimes of Fig.
particle aspect ratio λ and the viscometric coefficient δ.
3, we further find that the normalized eigenvector corre-
Dynamical analysis sponding to the positive eigenvalue matches the angle of
1. Eigenvalue analysis of alignment dynamics. Following alignment (φeq ) obtained numerically.
Bretherton [73], we explain the alignment of a particle In addition to validating our results of Fig. 3, Eq. (30)
in the shear plane by extracting a fundamental matrix provides a key insight that the contribution from active
solution of Eq. (1). For this, we first consider the non- disturbances is to effectively alter the strength of elon-
dimensional Jeffery equation in the expanded form as gation (prefactor of E∞ ) and the rotation rate (prefac-
ṗ = O∞ ⋅ p + Λ [E∞ ⋅ p − p(E∞ ∶ pp)] , (26) tor of O∞ ). The first prefactor is the more decisive as
α1 ≫ ∣α2 ∣. It suggests that a pusher (σ > 0) disturbs the
which can be represented in an unconserved form that local shear flow in order to increase the weight of elon-
facilitates a fundamental matrix solution gation. Once the activity exceeds the critical limit, the
local shear flow transforms into an effective elongation
q̇ = (ΛE∞ + O∞ ) ⋅ q. (27)
flow whose axis of elongation points in the direction of
q(t)
Here ∣q(t)∣ = p(t) and ∣q(t)∣ = exp[Λ (E∞ ∶ pp) t] repre- the normalized eigenvector associated with the positive
sents the exponential elongation of q. We study the an- eigenvalue. As pusher’s activity further increases, the lo-
gular dynamics of q, as its normalization yields back the cally elongated flow asymptotically approaches the pure
orientation vector p. The solution to Eq. (27) is elongation state where the impact of O∞ is negligible in
comparison (similar to our discussion of O∞ ≡ 0 case in
q(t) = exp [(ΛE∞ + O∞ )t]q(0). (28) previous paragraph).
11

2. Stability analysis of orbits. Here we find the stability Following [49, 54], we solve the above inhomogeneous
exponent of the complete non-linear Eq. (1) to analyze linear differential equation using the Green’s function
the onset of shear-plane rolling state as shown in Fig. G(p∣p′ ), which is governed by
2,3. For Newtonian fluids, θ = π/2 is one of the infinite
neutral Jeffery orbits. For SOF, the polymeric stress in − τ Dr ∇2p G(p∣p′ ) + G(p∣p′ ) = δ(p − p′ ), (37)
the fluid lifts this degeneracy. To quantify this, we write
down the angular dynamics as where δ represents the Dirac-delta function. Following
[75], we expand this G into spherical harmonic series as
θ̇ = f (θ, φ), φ̇ = g(θ, φ), (30a,b)
∞ n
and perform a Taylor expansion near the shear-plane: G(p∣p′ ) = ∑ ∑ Cn,m (p′ )Ynm (p). (38)
n=0 m=−n
θ(t) = π/2+(t). Here  represents the deviation from Jef-
fery orbit and and we determine its growth i.e., whether
Here Cn,m are the series coefficients and Ynm (p) repre-
it grows or shrinks with time. At O() we obtain
sent the spherical harmonics. Similarly, the Dirac-delta
d ∂f function can also be related to the spherical harmonics as
= ∣ , (31) δ(p − p′ ) = ∑∞ n m m m
n=0 ∑m=−n Yn (p)Yn (p ), where Yn repre-

dt ∂θ θ=π/2
sents the corresponding complex conjugate spherical har-
which can be simplified and integrated over an orbit to monic [75]. Eq. (37) is now expressible as
obtain
∞ n
2 m m
∑ ∑ Cn,m (p ){−τ Dr ∇p [Yn (p)] + Yn (p)} =
⎡ −2π ⎤ ′
⎢ 1 ∂f ⎥
 = 0 exp ⎢∫
⎢ ( ) dφ⎥⎥ . (32) n=0 m=−n
⎢ 0 g(θ, φ) ∂θ θ=π/2 ⎥ ∞ n
⎣ ⎦ m
∑ ∑ Yn (p)Ynm (p′ ). (39)
Here the integration is over −2π because the particle’s n=0 m=−n
rotation due to shear is in −φ direction. Eq. (32)
can be expressed in the form of Lyapunov exponent Using the properties of spherical harmonics [75], we sub-
( = 0 exp[LT ]) [74] as stitute ∇2p Ynm = −n(n+1)Ynm , take the inner product with
respect to YNM (p) on both sides of Eq. (39), and use the
−2π
1 1 ∂ θ̇ orthogonality property to obtain
L∼
T ∫0 [ ]
φ̇ ∂θ θ=π/2
dφ, (33)
Ynm (p′ )
Cn,m (p′ ) = , (40)
where T is the time period of a Jeffery orbit 2π(λ+λ−1 )/γ̇. 1 + τ Dr n(n + 1)
We show the solution to Eq. (33) in the inset of Fig. 3. √
where the normalization condition yields C0,0 = 1/ 4π.
Kinetic model
We finally use the Green’s function solution to find ψ(1) :
1. Near-equilibrium microstructure. First, we detail
the evaluation of microstructure near equilibrium (Pef =
3 E∞ ∶ pp
γ̇τ ≪ 1) i.e., when stochastic reorientation dominates the ψ(1) = (Λ + α1 DePea /8π) [ ]. (41)
shear-induced reorientation. In this limit, we expand the 4π 1 + 6τ Dr
orientation distribution as
In the limit of weak shear, this solution matches with the
ψ = ψ(0) + Pef ψ(1) + ⋯. (34) numerically obtained ψ for arbitrary Pef , whose evalua-
tion is discussed next.
We substitute this in Eq. (2) and collect the zeroth and
first order terms in Pef . At O(1), we get ψ(0) = 1/4π as 2. Numerical solution valid for arbitrary Pef . Now we
the isotropic microstructure. At O(Pef ), we have detail the numerical solution of Eq. (2), which uses the
n m
decomposition of ψ as ∑∞ n=0 ∑m=−n Cn,m (p)Yn (p). We
∇p ⋅{[ṗ(0) + De Pea ṗA /8π] ψ(0) }−τ Dr ∇2p ψ(1) +ψ(1) = 0.
(1) substitute this expansion in Eq. (2) and use the proper-
ties of spherical harmonics to obtain
(35)
Here we denote the equation of motion (Eq. 1) as having 1 ∞ n
(1) (1)
three parts: ṗ = ṗ(0) +DePef ṗP +DePea ṗA , where sub- − + ∑ ∑ Cn,m [1 + τ Dr n(n + 1)] Ynm +
4π n=0 m=−n
scripts represent passive and active components. Upon ∞ n
(1)
substituting this in Eq. (35), we find that the ṗP term Pef ∑ ∑ Cn,m H(Ynm ) = 0. (42)
2 n=0 m=−n
only contributes at O(Pef ), and is thus neglected. Upon
simplification, we find that the microstructure at O(Pef ) Here H(Ynm ) is a collection of expressions obtained by
is governed by simplifying ∇p ⋅ (ṗψ) . These are represented in terms of
3 angular momentum operators for computational conve-
−τ Dr ∇2p ψ(1) +ψ(1) = (Λ+α1 DePea /8π)E∞ ∶ pp. (36) nience [54] (detailed in Supplementary Note 3). We take

12

an inner product with Yij on both sides of Eq. (42) and The last term in the above expression is negligible and
use the orthogonality property to obtain is generated from the non-Newtonian component of ΣF p.
Thus, the major modifications due to fluid elasticity scale
1 j
− ∫ Y dp + Ci,j [1 + τ Dr i(i + 1)] + as Pe2a , as also depicted in Fig. 5d.
4π S i
∞ n
Pef ∑ ∑ Cn,m ∫ Yij H(Ynm )dp = 0. (43)
n=0 m=−n S
ACKNOWLEDGEMENTS
Since the n = 0 mode is already known from normaliza-
tion, the above equations can be recasted as the follow- Support from the Alexander von Humboldt Founda-
ing linear system of equations where Ci,j (for i ≥ 1) is tion is gratefully acknowledged. We also acknowledge
unknown: financial support from the Collaborative Research Cen-
Ci,j [1 + τ Dr i(i + 1)] + ter 910 funded by the Deutsche Forschungsgemeinschaft.
100 n
S.N acknowledges support from the Swedish Research
Pef ∑ ∑ Cn,m ∫ Yij H(Ynm )dp = Council (under grant no. 638-2013-9243) and Nordita,
n=1 m=−n S which is partially supported by NordForsk. A.C thanks
Jonas Einarsson for his advice on the effective equation
−Pef C0,0 ∫ Yij H(Y00 )dp., (44)
S of motion.
To find Ci,j coefficients, the above equations are solved
using the Clebsch-Gordon coefficient formulation, where
AUTHOR CONTRIBUTIONS
first 100 modes in n are used. We note that although the
numerical approach is valid for arbitrary Pef , there is
an indirect upper bound of weak viscoelasticity that we A.C. and H.S. planned the research. A.C performed
must adhere to, as we are using the SOF model. In non- the calculations. A.C., S.N., and H.S. analyzed the re-
dimensional terms this bound is: De Pef < 1. Thus, for sults and wrote the manuscript.
De = 0.1, we explore the results within an upper bound
of Pef = 10.
CONFLICTS OF INTEREST
Rheology. First we show the expanded version of the
flow-induced component of the particle stress
There are no conflicts to declare.
ΣF 3 ∇
p = −µf nl Aγ̇ a2 /2 (45)
where ai is a shorthand notation for the ensemble of
orientation moment of ith order: ⟨p⊗i ⟩. Expanding the
upper-convected derivative, we obtain

a2 = (Λ − 1) (E∞ ⋅ a2 + a2 ⋅ E∞ ) − 2Λa4 ∶ E∞
+ De[Pef β1 (E∞ ⋅ a4 ∶ E∞ − 2E∞ ∶ a6 ∶ E∞ + E∞ ∶ a4 ⋅ E∞ )
+ Pef β2 (O∞ ⋅ a4 ∶ E∞ − E∞ ∶ a4 ⋅ O∞ )
+ Pea α1 (E∞ ⋅ a2 − 2a4 ∶ E∞ + a2 ⋅ E∞ )
+ Pea α2 (O∞ ⋅ a2 − a2 ⋅ O∞ ) ]. (46)

This ΣF p matches with the Newtonian bulk stress for


De = 0 [53]. For near-equilibrium results (Pef ≪ 1), the
above expression is substituted in Eq. (5) and orientation
distribution from Eq. (41) is used to perform the ensem-
ble integral. Since the passive viscoelastic contribution
is O(Pe2f ), it does not contribute at the O(Pef ) calcu-
lations. The viscosity relates to the deviatoric particle-
induced stress as ϕµp = Σp xy /γ̇, where ϕ = na3 is the
volume fraction. This viscometeric relation can be sim-
plified to obtain the near-equilibrium viscosity ratio
µp 4A(5 − 3Λ) 48Aτ Dr − Pea /π De Pea α1
= + (Λ + )
µf 15 5 (1 + 6τ Dr ) 8π
De Pea Aα1
− . (47)
10π
1

Supplementary material for: “Orientational dynamics and rheology of active


suspensions in weakly viscoelastic flows”

I. COEFFICIENTS OF THE MULTIPOLE EXPANSION

The tensorial coefficients of the spheroidal multipoles are

pajklm = (pj pk − δjk /3)(pl pm − δlm /3), (S1a)

pbjklm = pj pl δkm + pj pm δkl + pk pl δjm + pk pm δjl − 4pj pk pl pm , (S1b)

pcjklm = −δjk δlm + δjl δkm + δjm δkl + pl pm δjk + pj pk δlm − pj pm δkl − pk pm δjl − pj pl δkm − pk pl δjm + pj pk pl pm . (S1c)

The other scalar coefficients are


√ √
R λ2 − 1 R λ2 − 1(λ2 + 1) R (λ2 − 1)3/2
A = , B = , C = , (S2a)
4(C − λ3 + λ) 4(C + λ3 − λ − 2Cλ2 ) 4(C + λ3 − λ − 2Cλ2 )

(λ2 − 1)3/2 (λ2 − 1)3/2 (Cλ + λ4 − 3λ2 + 2) (λ2 − 1)3/2


AS = , BS
= , C S
= .
4(C − 3λ3 + 3λ + 2Cλ2 ) 8(C + λ3 − λ − 2Cλ2 )(−3Cλ + λ4 + λ2 − 2) 2(3C + λ3 + 2λ5 − 7λ3 + 5λ)
(S2b)
√ √
Here C = λ2 − 1 coth−1 [λ/ λ2 − 1].

II. FURTHER RESULTS OF ORIENTATION DYNAMICS

A. Coefficients of equation of motion

Fig. S1 shows the coefficients of the analytical equation of motion. They vary weakly with aspect ratio λ and the
viscometric parameter δ, which typically lies between −0.7 and −0.5. Since the coefficients vary weakly with λ and
the integration time is very large for λ > 10, we consider aspect ratios between 5 and 10.

(a) (b)
1.5 𝛽"
𝛼!

1.0

0.5 𝛽!
𝛼"
0
5 6 7 8 9 10
𝜆 𝜆

Fig. S1. Coefficients of equation of motion (Eq. 1 in main text). αi represent active coefficients and βi are passive coefficients.

B. Alignment angle

Fig. S2 shows the comparison of angles of alignment obtained by integrating the analytical equation of motion
(Eq. 1 of main text) and those from the eigenvalue analysis. Note that the phase diagram has a weak dependency on
aspect ratio.
2

3
4

𝜙!" 2

4 4

-4 -2 0 2
𝜎

Fig. S2. Comparison of alignment angles obtained by integrating the full equation of motion (dots) with those found using
analytical approximation (line). Blue: λ = 5, red: λ = 10. Vertical dashed lines separate the regions of alignment. Other
parameters: δ = −0.6, Wi = 0.1 .

C. Passive spheroid in weakly viscoelastic shear flow

The non-dimensional equation of motion for the orbits is


ṗ = (I − pp) ⋅ (E∞ ⋅ p) [Λ + Wi σα1 + Wi β1 E∞ ∶ pp] + Ω∞ × p [1 + Wi σα2 + Wi β2 E∞ ∶ pp] . (S3)
This can be written in the trigonometric form as
Λ
θ̇ = sin 2θ sin 2φ + Wi β1 sin3 θ cos θ sin2 φ cos2 φ (S4a)
4

1 Wi
φ̇ = − [λ2 sin2 φ + cos2 φ] + [β1 cos(2φ) − β2 ] sin2 θ sin 2φ. (S4b)
λ2 +1 4
To quantify the log-rolling and compare it with Brunn [37], we use the following transformation [59]:

⎛ Cλ ⎞ tan φ
θ = arctan √ , τ = arctan ( ). (S5)
⎝ λ2 cos2 φ + sin2 φ ⎠ λ

Here C represents the orbit constant (C = 0 and C = ∞ being the log-rolling and shear-plane rotation orbits, respec-
tively) and τ is essentially the time coordinate that is scaled with T /2π, where T is the Jeffery orbit. Using the above
relation, we obtain the evolution equation of the orbit constant as
[C 3 (β1 λ2 − β2 λ2 + β1 + β2 ) sin2 τ cos2 τ ]
Ċ(t) = Wi ,
2 (λ2 C 2 cos2 τ + C 2 sin2 τ + 1)
λ ⎡ sin τ cos τ (2β1 λ2 C 2 cos2 τ + β1 + β2 ) 1 ⎤
⎢ ⎥
τ̇ (t) = + Wi ⎢ − (β 1 + β 2 ) sin 2τ ⎥. (S6)
1 + λ2 ⎢ 2C 2 (λ2 cos2 τ + sin2 τ ) + 2 4 ⎥
⎣ ⎦
We now compare our result with Brunn [37] who, for a rigid tri-dumbbell, showed that the particle slowly drifts
towards the log-rolling orbit. We show the match between orbit drift from Eq. (S6) and Eq. (4.17) from Brunn in
Fig. S3. Note that over one period the particle’s orbit declines in a wiggly manner. Interestingly, these wiggles have
a physical significance and can be characterized as orbit reduction with two distinct plateaus. Fig. S3b shows that
the long plateau (around t/T = 1/4) occurs first and is the state where particle is in the vicinity of flow-vorticity plane
(where it slows down and spends most of its time). Once the particle exits the alignment state, it flips quickly to the
other side (−x axis). This flip generates the short plateau around t/T = 1/2.
.

III. DETAILS OF KINETIC MODEL

Here we provide full expressions and further details involved in the computational steps of kinetic model. The
spectral decomposition of ψ is carried out as
∞ n
ψ(θ, φ) = ∑ ∑ Cn,m (p)Ynm (p). (S7)
n=0 m=−n
3

(a) 1.0 (b) 1.0


Our work
0.8
Brunn 0.9
C 0.6 C
C0 0.4 C0 0.8

0.2 0.7
0 2 4 6 8 10 0 0.2 0.4 0.6 0.8 1.0
t/T t/T

Fig. S3. Comparison with Brunn [37] for the variation of the orbit constant C/C0 , where C0 is the orbit constant for t = 0.
(a) Variation over 10 periods. (b) Variation over one period. Other parameters: λ = 5, T = 2π λ λ+1 , δ = −0.6, Wi = 0.1.
2

Substituting this in the Smoluchowski equation of main text and using the properties of spherical harmonics yields

n n
1 ∞ ∞
− + ∑ ∑ Cn,m [1 + τ Dr n(n + 1)] Ynm + Pef ∑ ∑ Cn,m H(Ynm ) = 0. (S8)
4π n=0 m=−n n=0 m=−n

Here H(Ynm ) is a term obtained by simplifying ∇p ⋅ (ṗψ) and is represented in terms of angular momentum operators
for computational convenience [54].

i Lz iΛ
H(Ynm ) = [− + (3iYnm sin2 (θ) sin(2φ) + Ly sin(2θ) sin(φ) + Lz cos(2θ) sin2 (φ) + Lz cos2 (φ))] +
2 2
i Pea
De{ [α1 (3iYnm sin2 (θ) sin(2φ) + Ly sin(2θ) sin(φ) + Lz cos(2θ) sin2 (φ) + Lz cos2 (φ)) − α2 Lz ] +
16π
i Pef
[β1 sin2 (θ)( − iYnm (3 + 5 cos(2θ)) − 10iYnm sin2 (θ) cos(4φ) + 2Ly sin(2θ) cos(φ) − 2Ly sin(2θ) cos(3φ)+
16
2Lz sin2 (θ) sin(4φ) + 2Lz cos(2θ) sin(2φ) + 2Lz sin(2φ)) + β2 sin2 (θ) (8iYnm cos(2φ) − 4Lz sin(2φ)) ]}. (S9)

Here, the angular momentum operators (L) are shorthand notation for the following gradients on harmonic Ynm :

∂Ynm ∂Y m ∂Ynm
Ly = Ly ∣Ynm ⟩ = −i cos φ + i cot θ sin φ n and Lz = Lz ∣Ynm ⟩ = −i . (S10)
∂θ ∂φ ∂φ

The trigonometric functions in the Eq. (S9) are converted into spherical harmonics. For example:


2π −1 1 √ √
sin θ cos θ sin φ = i (Y + Y21 ), cos2 θ = (2 4π/5Y20 + 4πY00 ) . (S11)
15 2 3
4

Using such transformations, Eq. (S9) yields


√ √ √ √
√ 2π 2π 2π −2 2π 2
H(Ynm ) = [−i πLz Y00 + Λ(− Ly Y2 −
−1 1
Ly Y2 + i m
Y2 (Lz − 3Yn ) + i Y (Lz + 3Ynm )
15 15 15 15 2

1 √ 2 π
+ i πLz Y00 + i Lz Y20 )] +
3 3 5
α1Ly Y −1 α1Ly Y 1 iLz Y00 (α1 − 3α2) iα1Lz Y2−2 iα1Lz Y20 iα1Lz Y22
De{Pea [− √ 2 − √ 2 + √ + √ + √ + √ +
4 30π 4 30π 24 π 4 30π 12 5π 4 30π
√ √
1 3 m −2 1 3
− i α1Yn Y2 + i α1Ynm Y22 ] +
4 10π 4 10π
√ √ √ √ √
1 2π 1 2π 1 π 1 π 1 π
Pef [ i β1 Ly Y2 − i
−1
β1 Ly Y21 − i β1 Ly Y4−3 − i β1 Ly Y4−1 + i β1 Ly Y41 +
7 15 7 15 3 35 21 5 21 5
√ √ √ √
1 π 1 π 1 π 1 2π
i 3
β1 Ly Y4 − Lz Y2 (β1 − 7β2 ) +
−2 2
Lz Y2 (β1 − 7β2 ) − β1 Y4−4 (Lz − 5Ynm )+
3 35 7 30 7 30 3 35
√ √ √ √
1 2π m 1 2π 1 2π 2 π 2√
4
β1 Y4 (Lz + 5Yn ) − β1 Lz Y4 +
−2 2
β1 Lz Y4 + β1 Ynm Y20 − πβ1 Ynm Y40 +
3 35 21 5 21 5 7 5 21
√ √
2π m −2 2π
− β2 Yn Y2 − β2 Ynm Y22 ]} (S12)
15 15

The angular momentum operators are further simplified using the following relations [75]

1√ 1√
Ly = Ly ∣Ynm ⟩ = n(n + 1) − m(m + 1)Ynm+1 − n(n + 1) − m(m − 1)Ynm−1 and Lz = Lz ∣Ynm ⟩ = mYnm . (S13)
2 2
We substitute above relations in Eq. (S12) to simplify H(Ynm ) and use that in Eq. (S8). Next, we take an inner
product with Yij on both sides of Eq. (S8) and use the orthogonality property to obtain
n
1 j

j m
− ∫ Yi dp + Ci,j [1 + τ Dr i(i + 1)] + Pef ∑ ∑ Cn,m ∫ Yi H(Yn )dp = 0. (S14)
4π S n=0 m=−n S

√ √
The first term is only non-zero for the zeroth mode (i = 0), which yields −1/ 4π. Since the C0,0 = 1/ 4π is already
known from normalization, the first term cancels the i = 0 mode of second term. Using these identities, Eq. (14) can
be recasted as the following linear system of equations in terms of Ci,j :
100 n
Ci,j [1 + τ Dr i(i + 1)] + Pef ∑ ∑ Cn,m ∫ Yij H(Ynm )dp = −Pef C0,0 ∫ Yij H(Y00 )dp, for i ≥ 1, (S15)
n=1 m=−n S S

where the right hand side is known. The two integral terms in the above set of equations contain the inner product of
three spherical harmonics. These are represented in terms of Clebsch-Gordon coefficients that exploit the orthogonal
relations between harmonics for computational ease [75]:
¿
π 2π Á (2a + 1)(2n + 1) a n N a n N
∫ ∫ YNM Yab Ynm sin θ dφdθ = Á
À [ ][ ]. (S16)
0 0 4π(2N + 1) 0 0 0 b m M

The bracket terms here are the Clebsch-Gordan coefficients inbuilt in Mathematica 13.
skip
Below we show the match between our results (obtained from the above kinetic model) in the limit De = 0 and
those obtained by Saintillan [24, Fig. 5a, τ̃ = 1] .

REFERENCES

[1] M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. Liverpool, J. Prost, M. Rao, and R. A. Simha, Rev. Mod. Phys. 85,
1143 (2013).
5

0.3
0.2

μp
μf
0.1
0.0
100 101 102
Pef

Fig. S4. Comparing the shear viscosity as a function of Pef from our analysis in the Newtonian fluid limit with that of
Saintillan [24]. for ∣Pea∣
8πA
= 1, Λ = 1, τ = 1, Dr = 0.1. Lines (blue is puller, red is pusher) indicate our viscosity ratio and dots
correspond to Saintillan [24], whose viscosity is defined here with respect to the particle volume fraction ϕ = na3 . Saintillan
[24] used ϕ = nl3 A.

[2] A. Zöttl and H. Stark, Journal of Physics: Condensed Matter 28, 253001 (2016).
[3] E. Lauga, The fluid dynamics of cell motility, Vol. 62 (Cambridge University Press, 2020).
[4] H. C. Berg and D. A. Brown, nature 239, 500 (1972).
[5] M. Demir and H. Salman, Biophysical journal 103, 1683 (2012).
[6] H. Stark, Accounts of Chemical Research 51, 2681 (2018).
[7] A. J. Mathijssen, N. Figueroa-Morales, G. Junot, É. Clément, A. Lindner, and A. Zöttl, Nat. Comm. 10, 1 (2019).
[8] G. Jing, A. Zöttl, É. Clément, and A. Lindner, Science advances 6, eabb2012 (2020).
[9] V. S. Doan, P. Saingam, T. Yan, and S. Shin, ACS nano 14, 14219 (2020).
[10] A. Ramamonjy, J. Dervaux, and P. Brunet, Physical Review Letters 128, 258101 (2022).
[11] S. S. Suarez and A. Pacey, Hum. Reprod. Update 12, 23 (2006).
[12] A. Zöttl and H. Stark, Phys. Rev. Lett. 108, 218104 (2012).
[13] J. Sznitman and P. E. Arratia, Complex Fluids in Biological Systems (Springer, 2015) pp. 245–281.
[14] R. Shogren, T. A. Gerken, and N. Jentoft, Biochemistry 28, 5525 (1989).
[15] S. K. Lai, Y.-Y. Wang, D. Wirtz, and J. Hanes, Advanced drug delivery reviews 61, 86 (2009).
[16] A. J. Mathijssen, T. N. Shendruk, J. M. Yeomans, and A. Doostmohammadi, Phys. Rev. Lett. 116, 028104 (2016).
[17] A. Choudhary and H. Stark, Soft Matter 18, 48 (2022).
[18] A. Sokolov and I. S. Aranson, Physical review letters 103, 148101 (2009).
[19] J. Gachelin, G. Mino, H. Berthet, A. Lindner, A. Rousselet, and É. Clément, Physical review letters 110, 268103 (2013).
[20] A. G. McDonnell, T. C. Gopesh, J. Lo, M. O’Bryan, L. Y. Yeo, J. R. Friend, and R. Prabhakar, Soft Matter 11, 4658
(2015).
[21] H. M. López, J. Gachelin, C. Douarche, H. Auradou, and E. Clément, Phys. Rev. Lett. 115, 028301 (2015).
[22] Y. Hatwalne, S. Ramaswamy, M. Rao, and R. A. Simha, Phys. Rev. Lett. 92, 118101 (2004).
[23] B. M. Haines, A. Sokolov, I. S. Aranson, L. Berlyand, and D. A. Karpeev, Physical Review E 80, 041922 (2009).
[24] D. Saintillan, Exp. Mech. 50, 1275 (2010).
[25] G. B. Jeffery, Proceedings of the Royal Society of London. 102, 161 (1922).
[26] S. Rafaı̈, L. Jibuti, and P. Peyla, Phys. Rev. Lett. 104, 098102 (2010).
[27] G. Li and A. M. Ardekani, Phys. Rev. Lett. 117, 118001 (2016).
[28] G. Li, E. Lauga, and A. M. Ardekani, Journal of Non-Newtonian Fluid Mechanics 297, 104655 (2021).
[29] S. E. Spagnolie and P. T. Underhill, arXiv preprint arXiv:2208.03537 (2022).
[30] A. Patteson, A. Gopinath, M. Goulian, and P. Arratia, Sci. rep. 5, 15761 (2015).
[31] A. Zöttl and J. M. Yeomans, Nature Physics 15, 554 (2019).
[32] S. Kamdar, S. Shin, P. Leishangthem, L. F. Francis, X. Xu, and X. Cheng, Nature 603, 819 (2022).
[33] S. Liu, S. Shankar, M. C. Marchetti, and Y. Wu, Nature 590, 80 (2021).
[34] R. B. Bird, R. C. Armstrong, and O. Hassager, Dynamics of polymeric liquids. Vol. 1: Fluid mechanics (1987).
[35] D. F. James, Annual Review of Fluid Mechanics 41, 129 (2009).
[36] L. Leal, J. Fluid. Mech. 69, 305 (1975).
[37] P. Brunn, J. Fluid Mech. 82, 529 (1977).
[38] E. Hinch and L. Leal, J. Fluid. Mech. 71, 481 (1975).
[39] A. T. Chwang and T. Y.-T. Wu, Journal of Fluid mechanics 67, 787 (1975).
[40] S. E. Spagnolie and E. Lauga, J. Fluid Mech. 700, 105–147 (2012).
[41] A. P. Berke, L. Turner, H. C. Berg, and E. Lauga, Phys. Rev. Lett. 101, 038102 (2008).
[42] K. Drescher, R. E. Goldstein, N. Michel, M. Polin, and I. Tuval, Phys. Rev. Lett. 105, 168101 (2010).
[43] J. Einarsson, F. Candelier, F. Lundell, J. Angilella, and B. Mehlig, Physics of Fluids 27, 063301 (2015).
[44] F. Gauthier, H. Goldsmith, and S. Mason, Rheo.l Acta 10, 344 (1971).
[45] G. D’Avino and P. L. Maffettone, Journal of Non-Newtonian Fluid Mechanics 215, 80 (2015).
[46] H. Berg, Random Walks in Biology. Princeton Univ. Press, Princeton, NJ (1983).
[47] E. Lauga, Ann. Rev. Fluid Mech. 48, 105 (2016).
6

[48] T. C. Adhyapak and H. Stark, Soft Matter 12, 5621 (2016).


[49] S. Nambiar, P. Nott, and G. Subramanian, Journal of Fluid Mechanics 812, 41 (2017).
[50] M. Doi, S. F. Edwards, and S. F. Edwards, The theory of polymer dynamics, Vol. 73 (Oxford University Press, 1988).
[51] C. Cohen, B. Chung, and W. Stasiak, Rheo. Acta 26, 217 (1987).
[52] S. Chattopadhyay, R. Moldovan, C. Yeung, and X. Wu, Proc. Natl. Acad. Sci. 103, 13712 (2006).
[53] E. Hinch and L. Leal, J. Fluid. Mech. 76, 187 (1976).
[54] S. B. Chen and D. L. Koch, Physics of Fluids 8, 2792 (1996).
[55] E. Hinch and L. Leal, Journal of Fluid Mechanics 52, 683 (1972).
[56] A. J. Wolfe, M. P. Conley, and H. C. Berg, Proceedings of the National Academy of Sciences 85, 6711 (1988).
[57] B. Chung and C. Cohen, J. Nonnewton. Fluid Mech. 25, 289 (1987).
[58] Y. Iso, D. L. Koch, and C. Cohen, J. Nonnewton. Fluid Mech. 62, 115 (1996).
[59] L. Leal and E. Hinch, Journal of Fluid Mechanics 46, 685 (1971).
[60] J. Elgeti, R. G. Winkler, and G. Gompper, Reports on progress in physics 78, 056601 (2015).
[61] G. Junot, N. Figueroa-Morales, T. Darnige, A. Lindner, R. Soto, H. Auradou, and E. Clément, EPL (Europhysics Letters)
126, 44003 (2019).
[62] A. Brown, S. Clarke, P. Convert, and A. Rennie, Journal of rheology 44, 221 (2000).
[63] A. Einstein, Ann. Phys. 19, 289 (1906).
[64] J. Férec, E. Bertevas, B. Khoo, G. Ausias, and N. Phan-Thien, Journal of Non-Newtonian Fluid Mechanics 239, 62
(2017).
[65] R. B. Bird, C. F. Curtiss, R. C. Armstrong, and O. Hassager, Dynamics of polymeric liquids, volume 2: Kinetic theory
(Wiley, 1987).
[66] L. Ren, N. Nama, J. M. McNeill, F. Soto, Z. Yan, W. Liu, W. Wang, J. Wang, and T. E. Mallouk, Sci. Adv. 5, 3084
(2019).
[67] A. Aghakhani, O. Yasa, P. Wrede, and M. Sitti, Proc. Natl. Acad. Sci. 117, 3469 (2020).
[68] Y. Bozorgi and P. T. Underhill, Physical Review E 84, 061901 (2011).
[69] A. M. Ardekani and E. Gore, Phys. Rev. E 85, 056309 (2012).
[70] N. Narinder, C. Bechinger, and J. R. Gomez-Solano, Phys. Rev. Lett. 121, 078003 (2018).
[71] S. A. Abtahi and G. J. Elfring, Physics of Fluids 31, 103106 (2019).
[72] G. J. Elfring and E. Lauga, in Complex fluids in biological systems (Springer, 2015) pp. 283–317.
[73] F. P. Bretherton, J. Fluid Mech. 14, 284 (1962).
[74] S. H. Strogatz, Nonlinear dynamics and chaos: with applications to physics, biology, chemistry, and engineering (CRC
Press, 2018).
[75] G. B. Arfken and H. J. Weber, Mathematical methods for physicists (American Association of Physics Teachers, 1999).

You might also like