This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Spheres and fibres in turbulent flows at various Reynolds numbers

Ianto Cannon1    Stefano Olivieri1,2    Marco E. Rosti1 [email protected] 1Complex Fluids and Flows Unit, Okinawa Institute of Science and Technology Graduate University, 1919-1 Tancha, Onna-son, Okinawa 904-0495, Japan
2 Department of Aerospace Engineering, Universidad Carlos III de Madrid, Avda. de la Universidad, 30. 28911 Leganés, Spain
Abstract

We perform fully coupled numerical simulations using immersed boundary methods of finite-size spheres and fibres suspended in a turbulent flow for a range of Taylor Reynolds numbers 12.8<Reλ<44212.8<Re_{\lambda}<442 and solid mass fractions 0M10\leq M\leq 1. Both spheres and fibres reduce the turbulence intensity with respect to the single-phase flow at all Reynolds numbers, with fibres causing a more significant reduction than the spheres. The particles’ effect on the anomalous dissipation tends to vanish as ReRe\to\infty. A scale-by-scale analysis shows that both particle shapes provide a “spectral shortcut” to the flow, but the shortcut extends further into the dissipative range in the case of fibres. Multifractal spectra of the near-particle dissipation show that spheres enhance dissipation in two-dimensional sheets, and fibres enhance the dissipation in structures with a dimension greater than one and less than two. In addition, we show that spheres suppress vortical flow structures, whereas fibres produce structures which completely overcome the turbulent vortex stretching behaviour in their vicinity.

Refer to caption Refer to caption

Figure 1: Images of the simulated domain. Left column: flows with fixed spheres. Middle column: single phase flows. Right column: flows with fixed fibres. Flows on the top row have ReABC=894Re_{ABC}=894. Flows on the bottom row have ReABC=55.9Re_{ABC}=55.9. At the boundary of each domain, we show the dissipation on a logarithmic scale, where ϵs\epsilon_{s} is the mean dissipation of the single-phase flow at ReABC=894Re_{ABC}=894.

I Introduction

Particle-laden turbulent flows abound in our environment; see, for example, volcanic ash clouds, sandstorms, and ocean microplastics. In these cases, the particles are seen in a range of shapes, and the intensity of the carrier turbulent flow can significantly vary. In this context, the present study delves into the interaction between particles and turbulent flows, focusing on two distinct particle shapes, spheres and fibres, and exploring turbulent flows across a large range of Reynolds numbers.

The interaction of particles with turbulent flows has garnered significant research attention since Richardson and Proctor [1] used balloons to track eddies in the Earth’s atmosphere in 1927. Light spherical particles as tracers are now a ubiquitous experimental tool [2], and even fibre-shaped tracers have been employed [3]. For heavier particles, more complex motion is seen, such as inertial clustering [4], preferential sampling [5], and enhancement of the mean flow [6]. Numerical simulations in this field have been a valuable tool, as results can be readily processed to study length-scale dependent statistics such as radial distribution functions [7], and histograms of Voronoï area [4]. Most of such particle-laden turbulent studies make the one-way coupling assumption, whereby the particles are affected by the fluid motion, but the fluid does not feel any effect of the particles. When the total mass fraction or volume fraction of particles in the system increases, the one-way coupling assumption becomes invalid, and we must model the back-reaction effect of the particles on the fluid. This opens up a zoo of new phenomena, including drag reduction [8], cluster-induced turbulence [9], and new scalings in the energy spectrum [10]. A recent review by Brandt and Coletti [11] has pointed out a gap in the current understanding of particles in turbulence, which lies between small heavy particles and large weakly buoyant particles. In addition, most real-world particle-laden flows involve particles with a high degree of anisotropy, while research in particle-laden turbulence has overwhelmingly focused on spherical particles.

This article addresses the gaps by studying a range of particle mass fractions MM, ranging from single-phase (M=0M=0) to fixed particles (M=1M=1). We make simulations using a fully coupled approach to elucidate how particles modulate turbulence. Furthermore, we vary the turbulence intensity, allowing us to ask the question at what Reynolds numbers (ReRe) does turbulence modulation emerge? And does the modulation effect persist as ReRe\to\infty? Crucially, we connect our results to real-world flows by investigating isotropically-shaped particles (spheres) and anisotropically-shaped particles (fibres). The spheres have a single characteristic length, given by their diameter, whereas the fibres have two: their length and thickness. This naturally allows us to ask how do the particle’s characteristic lengths impact the scales of the turbulent flow?

A number of works have investigated particle shape; see Voth and Soldati [12] for a review of the behaviour of anisotropic particles in turbulence, including oblate spheroids, prolate spheroids and fibres. In 1932 Wadell [13] defined the sphericity of a particle as its surface area divided by the area of a sphere with equivalent volume; Wadell used sphericity to classify quartz rocks according to their shape. More recently, Zhao et al. [14] made direct numerical simulations of oblate and prolate spheroids in a channel flow and found that away from the channel walls, prolate spheroids tend to rotate about their symmetry axis (spinning), whereas oblate particles rotate about an axis perpendicular to their symmetry axis (tumbling). Ardekani et al. [15] showed that oblate spheroids can reduce drag in a turbulent channel by aligning their major axes parallel to the wall. Yousefi et al. [16] simulated spheres and oblate spheroids in a shear flow and found that the rotation of the spheroids can enhance the kinetic energy of the flow. At the other shape extreme from oblate spheroids, we have fibres, which are long and thin. In this article, we choose to study rigid fibres (also known as rods) as they are a simple example of a particle with high anisotropy. Concerning fibres, Paschkewitz et al. [17], Gillissen et al. [18] and others showed that rigid fibres align with vorticity in channel flows, dissipating the vortex structures, and drag reductions up to 26% have been measured [17]. In this article, we wish to investigate how the shapes and length scales of particles interact with turbulence, so we choose the triperiodic flow geometry. This geometry avoids the effects of walls and other large structures on the flow, enabling us to focus on the emergent length scales of the flow.

A few works have investigated the effect of turbulence intensity on particle-laden flows. Lucci et al. [19] simulated spheres of various diameters cc in decaying isotropic turbulence and found that spheres reduce the fluid kinetic energy and enhance dissipation when c>ηc>\eta, where η\eta is the Kolmogorov length scale. In this article, we also vary the ratio c/ηc/\eta. However, unlike [19], we do it by varying the fluid viscosity, not the particle size. This enables us to keep the particle volume, particle surface area and number of particles constant across our cases. Oka and Goto [20] studied sphere diameters in the range 7.8<c/η<647.8<c/\eta<64, in turbulent flows with Reλ94Re_{\lambda}\approx 94 and showed that vortex shedding and turbulence attenuation occur when cλρ/ρ~c\gtrsim\lambda\rho/\tilde{\rho}, where λ\lambda is the Taylor length-scale, ρ\rho and ρ~\tilde{\rho} are the fluid and solid densities, respectively. Shen et al. [21] investigated the effect of solid-fluid density ratio for flows with 38<Reλ<7438<Re_{\lambda}<74, and 8.8<c/η<188.8<c/\eta<18, showing that higher density spheres cause an increase in the normalised dissipation rate and a greater turbulence attenuation. Peng et al. [22] parametrized the attenuation of turbulent kinetic energy by spheres with 7.1<c/η<157.1<c/\eta<15 and 41<Reλ<6341<Re_{\lambda}<63. They found that the particle mass fraction is indeed a strong indicator of attenuation, and there is negligible dependence on the Taylor Reynolds number ReλRe_{\lambda} for this range. Compared to the above works, we choose a wider range of turbulence intensities, such that 11.7<c/η<12511.7<c/\eta<125 and 12.8<Reλ<44212.8<Re_{\lambda}<442, and we extend the study to include non-spherical particles.

This article is structured as follows: In the following section, we describe the numerical methods and the parameters used in our study. In section III, we present and discuss the results of our simulations, and section IV concludes with a summary and outlook on future research.

II Methods and setup

We tackle the problem using large direct numerical simulations on an Eulerian grid of 102431024^{3} points with periodic boundaries in all three directions. To obtain the fluid velocity uiu_{i} and pressure pp, we solve the incompressible Navier-Stokes equations for a Newtonian fluid with kinematic viscosity ν\nu and density ρ\rho,

tui+j(uiuj)\displaystyle\partial_{t}u_{i}+\partial_{j}(u_{i}u_{j}) =νjjuiip/ρ+fiABC+fisf,\displaystyle=\nu\partial_{jj}u_{i}-\partial_{i}p/\rho+f^{ABC}_{i}+f^{sf}_{i}, (1)
juj\displaystyle\partial_{j}u_{j} =0,\displaystyle=0, (2)

where indices i,j{1,2,3}i,j\in\{1,2,3\} denote the Cartesian components of a vector, and repeated indices are implicitly summed over. The turbulent flow is sustained by an ABC forcing [23, 24], which is made of sinusoids with a wavelength 2πL2\pi L equal to the domain size,

f1ABC=Csin(x3/L)+Ccos(x2/L),f2ABC=Csin(x1/L)+Ccos(x3/L),f3ABC=Csin(x2/L)+Ccos(x1/L).\begin{split}f^{ABC}_{1}=C\sin({x_{3}}/{L})+C\cos({x_{2}}/{L}),\\ f^{ABC}_{2}=C\sin({x_{1}}/{L})+C\cos({x_{3}}/{L}),\\ f^{ABC}_{3}=C\sin({x_{2}}/{L})+C\cos({x_{1}}/{L}).\end{split} (3)

The amplitude CC of the forcing is used to define the forcing Reynolds number ReABCC1/2L3/2/ν{Re_{ABC}\equiv C^{1/2}L^{3/2}/\nu}. To discern the effect of increasing turbulence intensity, we conduct a number of simulations with various values of ReABCRe_{ABC}, given in table 1.

When the single-phase flows reach a statistically steady state, we add the solid particles at random (non-overlapping) locations and orientations in the domain. We allow the flow to reach a statistically steady state again, and measure the statistics presented in section III, which were averaged in time over multiple large-eddy turnover times τf2πL/urms\tau_{f}\equiv 2\pi L/u_{{rms}}, where urmsu_{{rms}} is the root mean square of the fluid velocity. This procedure is repeated for every solid mass fraction investigated, defined as Mms/(ms+mf)M\equiv m_{s}/(m_{s}+m_{f}) where msm_{s} and mfm_{f} are the total mass of solid and fluid. The single-phase cases have M=0M=0, and the cases with fixed particles have M=1M=1. The solid phase is two-way coupled to the fluid using the immersed boundary method, and the back-reaction of the particles on the fluid is enforced by 𝐟𝐬𝐟\mathbf{f^{sf}} in equation 1. As can be seen in figure 1, we simulate two types of particles, spheres and fibres; to isolate the effect of particle isotropy on the flow, we choose the fibres and spheres to have the same size: the spheres have diameter cc, and the fibres have length cc. We choose c=L/2{c=L/2}, which lies inside the inertial range of scales for all of our cases.

We use the in-house Fortran code Fujin to solve the flow numerically. Time integration is carried out using the second-order Adams-Bashforth method, and incompressibility (equation 2) is enforced in a pressure correction step [25], which uses the fast Fourier transform. Variables are defined on a staggered grid; velocities and forces are defined at the cell faces, while pressure is defined at the cell centres. Second-order finite differences are used for all spatial gradients. See https://groups.oist.jp/cffu/code for validations of the code.

marker ReABCRe_{ABC} MM η/L\eta/L ReλRe_{\lambda} ϵ/urms3\epsilon\mathcal{L}/u_{{rms}}^{3}
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.0 4.06×𝟏𝟎𝟑\mathbf{4.06\times 10^{-3}} 433 0.399
\blacksquare 447447 0.0 6.31×1036.31\times 10^{-3} 308 0.382
\blacktriangle 224224 0.0 1.11×1021.11\times 10^{-2} 204 0.400
\blacklozenge 112112 0.0 1.95×1021.95\times 10^{-2} 116 0.473
\mdlgblkcircle\mdlgblkcircle 55.9\mathbf{55.9} 0.0 2.99×𝟏𝟎𝟐\mathbf{2.99\times 10^{-2}} 101 0.428

Table 1: Single-phase flows. ReABCRe_{ABC} is the forcing Reynolds number, and MM is the solid mass fraction. We measure the Kolmogorov length scale ην3/4/ϵ1/4\eta\equiv\nu^{3/4}/\epsilon^{1/4}, the Taylor Reynolds number ReλRe_{\lambda}, and the dissipation ϵ\epsilon, which has been normalised using the integral length scale \mathcal{L} and the root mean square velocity urmsu_{{rms}} of the fluid. The largest and smallest values of each parameter are shown in bold.

II.1 Motion of the spheres

The sphere motion and forces are modelled using an Eulerian-based immersed boundary method developed by Hori et al. [26]. The velocity 𝐔\mathbf{U} and rotation rate 𝛀\mathbf{\Omega} of each sphere are found by integrating the Newton-Euler equations in time

mtUi\displaystyle m\partial_{t}U_{i} =\oiintS(ρν(iuj+jui)pδij)njdSFcolni,\displaystyle=\oiint_{S}\left(\rho\nu(\partial_{i}u_{j}+\partial_{j}u_{i})-p\delta_{ij}\right)n_{j}dS-F^{col}n_{i}, (4)
ItΩi\displaystyle I\partial_{t}\Omega_{i} =\oiintSεijkc2njρν(kul+luk)nldS,\displaystyle=\oiint_{S}{\varepsilon_{ijk}}\frac{c}{2}n_{j}\rho\nu(\partial_{k}u_{l}+\partial_{l}u_{k})n_{l}dS, (5)

where δij\delta_{ij} is the Kronecker delta and εijk{\varepsilon_{ijk}} is the Levi-Civita symbol, SS is the surface of the sphere, and 𝐧\mathbf{n} is its normal. m=ρ~πc3/6{m=\tilde{\rho}\pi c^{3}/6} and I=mc3/20{I=mc^{3}/20} are the mass and moment of inertia of the sphere with diameter cc and density ρ~\tilde{\rho}. A soft-sphere collision force Fcol𝐧F^{col}\mathbf{n} is applied in the radial direction when spheres overlap [26].

Table 2 shows our choice of parameters for the sphere-laden flows. In all cases, we use 300 spheres, giving the solid phase a volume fraction of 0.0790.079. The characteristic time of the spheres is τs=ρ~c2/(18ρν)\tau_{s}=\tilde{\rho}c^{2}/(18\rho\nu), from which we can define the Stokes number Stτs/τfSt\equiv\tau_{s}/\tau_{f} of our flows. The particle Reynolds number of the spheres is Rep=c𝚫𝐮n𝚫𝐮nn/νRe_{p}=c\sqrt{\langle\mathbf{\Delta u}_{n}\cdot\mathbf{\Delta u}_{n}\rangle_{n}}/\nu, where 𝚫𝐮n\mathbf{\Delta u}_{n} is the difference between the velocity of the nnth particle and the local fluid velocity, averaged in a ball of diameter 2c2c centred on the sphere. The angled brackets n\langle\rangle_{n} denote an average over all spheres in the simulation.

marker ReABCRe_{ABC} MM ρ~/ρ\tilde{\rho}/\rho η/L\eta/L ReλRe_{\lambda} ϵ/urms3\epsilon\mathcal{L}/u_{{rms}}^{3} StSt RepRe_{p}
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.1 1.29 4.09×𝟏𝟎𝟑\mathbf{4.09\times 10^{-3}} 431 0.397 7.4 618
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.3 4.99 4.18×1034.18\times 10^{-3} 397 0.395 26.9 857
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.6 17.4 4.09×1034.09\times 10^{-3} 346 0.507 89.1 1120
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.9 105 4.18×1034.18\times 10^{-3} 280 0.625 471 1180
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 1.0 \boldsymbol{\infty} 4.29×1034.29\times 10^{-3} 247 0.708 \boldsymbol{\infty} 1190
\blacksquare 447447 1.0 \boldsymbol{\infty} 7.31×1037.31\times 10^{-3} 161 0.786 \boldsymbol{\infty} 559
\blacktriangle 224224 0.1 1.29 1.13×1021.13\times 10^{-2} 219 0.374 1.91 142
\blacktriangle 224224 0.3 4.99 1.16×1021.16\times 10^{-2} 181 0.477 6.49 198
\blacktriangle 224224 0.6 17.4 1.18×1021.18\times 10^{-2} 157 0.586 20.9 243
\blacktriangle 224224 0.9 105 1.19×1021.19\times 10^{-2} 123 0.751 109 266
\blacktriangle 224224 1.0 \boldsymbol{\infty} 1.26×1021.26\times 10^{-2} 101 0.936 \boldsymbol{\infty} 260
\blacklozenge 112112 1.0 \boldsymbol{\infty} 2.14×1022.14\times 10^{-2} 61.5 1.19 \boldsymbol{\infty} 117
\mdlgblkcircle\mdlgblkcircle 55.9\mathbf{55.9} 1.0 \boldsymbol{\infty} 3.60×𝟏𝟎𝟐\mathbf{3.60\times 10^{-2}} 34. 9 1.81 \boldsymbol{\infty} 44.6

Table 2: Sphere-laden flows. We set the solid-fluid density ratio ρ~/ρ\tilde{\rho}/\rho to obtain a range of solid mass fractions MM. Bulk statistics for the particles are the Stokes number StSt and the particle Reynolds number RepRe_{p}. The largest and smallest values of each parameter are shown in bold.

II.2 Motion of the fibres

For the motion and coupling of the fibres, we also use an immersed boundary method, but this one is Lagrangian and solves the Euler-Bernoulli equation for the position 𝐗\mathbf{X} of the beam with coordinate ll along its length [27, 28, 29]

π4d2(ρ~ρ)t2Xi=l(TlXi)+γl4XiFifs+Ficol,\frac{\pi}{4}d^{2}(\tilde{\rho}-\rho)\partial^{2}_{t}X_{i}=\partial_{l}(T\partial_{l}X_{i})+\gamma\partial^{4}_{l}X_{i}-F^{fs}_{i}+F^{col}_{i}, (6)

where TT is the tension, enforcing the inextensibility condition;

tXitXi=1.\partial_{t}X_{i}\partial_{t}X_{i}=1. (7)

The volumetric density of the fibre is ρ~\tilde{\rho}, and its stiffness is γ\gamma. Note that the fluid density ρ\rho in equation 6 cancels the inertia of the fluid in the fictitious domain inside the fibre [30]. To exclude deformation effects in our study, we choose a stiffness which limits the fibre deformation below 1%. The collision force 𝐅𝐜𝐨𝐥\mathbf{F^{col}} is the minimal collision model by Snook et al. [31]. The fluid-solid coupling force 𝐅𝐟𝐬\mathbf{F^{fs}} enacts the non-slip condition at the particle surface, and an equal and opposite force (𝐟𝐬𝐟\mathbf{f^{sf}} in equation 1) acts on the fluid. The spreading kernel onto the Eulerian grid has a width of three grid spaces, giving the fibre diameter d=3Δxd=3\Delta x in units of the Eulerian grid spacing [32, equation 14]. The fibre diameter 0.4η<d<5η0.4\eta<d<5\eta is on the order of the Kolmogorov length η\eta in all cases, i.e., it is smaller than the turbulent eddies in the energy-containing range of the flow. This allows us to consider the fibres as infinitesimally thin objects with a high degree of anisotropy. Table 3 shows our choice of parameters for the fibre-laden flows. In all cases, we use 10410^{4} fibres, giving the solid phase a volume fraction of around 5.4×1035.4\times 10^{-3}. The characteristic time of the fibres is calculated using a formulation which takes their aspect ratio βc/d\beta\equiv c/d into account [33],

τs=ρ~d218ρνβln(β+β21)β21,\tau_{s}=\frac{\tilde{\rho}d^{2}}{18\rho\nu}\frac{\beta\ln\left(\beta+\sqrt{\beta^{2}-1}\right)}{\sqrt{\beta^{2}-1}}, (8)

from which we can define the Stokes number Stτs/τfSt\equiv\tau_{s}/\tau_{f} of our flows. The particle Reynolds number of the fibres is Rep=d𝚫𝐮n𝚫𝐮nn/νRe_{p}=d\sqrt{\langle\mathbf{\Delta u}_{n}\cdot\mathbf{\Delta u}_{n}\rangle_{n}}/\nu, where 𝚫𝐮n\mathbf{\Delta u}_{n} is the difference between the velocity of the midpoint of the nnth fibre and the local fluid velocity, averaged in a ball of diameter 2c2c centred on the fibre’s midpoint. The angled brackets n\langle\rangle_{n} denote an average over all fibres in the simulation.

marker ReABCRe_{ABC} MM ρ~/ρ\tilde{\rho}/\rho η/L\eta/L ReλRe_{\lambda} ϵ/urms3\epsilon\mathcal{L}/u_{{rms}}^{3} StSt RepRe_{p}
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.2 47.2 4.09×1034.09\times 10^{-3} 422 0.425 1.45 34.9
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.3 81.8 4.02×𝟏𝟎𝟑\mathbf{4.02\times 10^{-3}} 442 0.432 2.6 41.2
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.6 279 4.17×1034.17\times 10^{-3} 340 0.713 7.52 46.4
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 0.9 1690 4.26×1034.26\times 10^{-3} 223 1.15 36.1 47.2
\pentagonblack\pentagonblack 𝟖𝟗𝟒\mathbf{894} 1.0 \boldsymbol{\infty} 4.40×1034.40\times 10^{-3} 201 1.29 \boldsymbol{\infty} 46.0
\blacksquare 447447 1.0 \boldsymbol{\infty} 7.44×1037.44\times 10^{-3} 115 1.78 \boldsymbol{\infty} 20.6
\blacktriangle 224224 0.1 21.3 1.16×1021.16\times 10^{-2} 192 0.434 0.155 8.31
\blacktriangle 224224 0.3 81.8 1.16×1021.16\times 10^{-2} 196 0.465 0.605 9.9
\blacktriangle 224224 0.6 279 1.19×1021.19\times 10^{-2} 167 0.63 1.85 11.2
\blacktriangle 224224 0.9 1690 1.23×1021.23\times 10^{-2} 72.5 2.25 7.16 9.11
\blacktriangle 224224 1.0 \boldsymbol{\infty} 1.29×1021.29\times 10^{-2} 54.9 3.17 \boldsymbol{\infty} 8.22
\blacklozenge 112112 1.0 \boldsymbol{\infty} 2.33×1022.33\times 10^{-2} 28.8 4.68 \boldsymbol{\infty} 3.32
\mdlgblkcircle\mdlgblkcircle 55.9\mathbf{55.9} 1.0 \boldsymbol{\infty} 4.27×𝟏𝟎𝟐\mathbf{4.27\times 10^{-2}} 12.8 8.60 \boldsymbol{\infty} 1.51

Table 3: Fibre-laden flows. We set the solid-fluid density ratio ρ~/ρ\tilde{\rho}/\rho to obtain a range of solid mass fractions MM. Bulk statistics for the particles are the Stokes number and particle Reynolds number. The largest and smallest values of each parameter are shown in bold.

III Results and discussion

III.1 Bulk statistics

Refer to caption
Figure 2: (a) Effect of the solid mass fraction MM on the Taylor Reynolds number ReλRe_{\lambda}. Flows with ReABC=224Re_{ABC}=224 are marked using triangles, and flows with ReABC=894Re_{ABC}=894 are marked using pentagons. (b) Effect of the forcing Reynolds number ReABCRe_{ABC} on the Taylor Reynolds number ReλRe_{\lambda} for the single phase and particle-laden cases with fixed particles (M=1M=1). In both plots, we show the single-phase flows in black, flows with spheres in blue, and flows with fibres in yellow; the shaded regions give the root-mean-square of ReλRe_{\lambda} in time. Each case is plotted using its marker, which is listed in table 1, 2, or 3.

The Taylor Reynolds number is defined as Reλurmsλ/νRe_{\lambda}\equiv u_{rms}\lambda/\nu, where urmsuiui𝐱,t/3u_{rms}\equiv\sqrt{\langle u_{i}u_{i}\rangle_{\mathbf{x},t}}/3 is the root-mean-square velocity, λ=urms15ν/ϵ\lambda=u_{{rms}}\sqrt{15\nu/\epsilon} is the Taylor length scale, ϵ2νsijsij𝐱,t\epsilon\equiv 2\nu\langle s_{ij}s_{ij}\rangle_{\mathbf{x},t} is the viscous dissipation rate, sij(iuj+jui)/2{s_{ij}\equiv(\partial_{i}u_{j}+\partial_{j}u_{i})/2} is the strain-rate tensor, and angled brackets 𝐱,t\langle\cdot\rangle_{\mathbf{x},t} indicate an average over space 𝐱\mathbf{x} and time tt.

The Taylor-Reynolds number is an indicator of the intensity of turbulence in the flow, and in figure 2, we show how ReλRe_{\lambda} compares for all of our cases. Figure 2a shows that increasing the solid mass fraction MM causes a reduction in ReλRe_{\lambda} at both high and low forcing Reynolds numbers. Both spheres and fibres produce a comparable decrease in ReλRe_{\lambda}, but the reduction effect is more substantial for very heavy fibres. Looking at the trend in ReλRe_{\lambda} with the forcing Reynolds number ReABCRe_{ABC} in figure 2b, we see, as expected, a monotonic increase in all three cases, i.e., single phase, sphere-laden, and fibre-laden flows.

Refer to caption
Figure 3: The turbulent kinetic energy KK^{\prime}, normalised using the forcing amplitude CC and length LL. (a) shows the effect of solid mass fraction MM, and (b) shows the effect of forcing Reynolds number ReABCRe_{ABC}. Each case is plotted using its marker, which is listed in table 1, 2, or 3. We show the single-phase flows in black, flows with spheres in blue, and flows with fibres in yellow; the shaded regions give the root-mean-square of KK^{\prime} in time.

A second way to quantify turbulence is the turbulent kinetic energy Kuiui𝐱,t/2K^{\prime}\equiv\langle u_{i}^{\prime}u_{i}^{\prime}\rangle_{\mathbf{x},t}/2. Similarly to Oka and Goto [20], we compute KK^{\prime} using the fluctuating part of the fluid velocity ui(𝐱,t)ui(𝐱,t)ui(𝐱,𝐭)tu_{i}^{\prime}(\mathbf{x},t)\equiv u_{i}(\mathbf{x},t)-\langle{u_{i}(\mathbf{x,t})\rangle_{t}}. Figure 3a shows the dependence of turbulent kinetic energy on particle mass fraction. We see that adding spheres and fibres reduces KK^{\prime} relative to the single-phase flow, with fibres having a slightly greater attenuation effect at both high and low forcing Reynolds numbers. From figure 3b, we see that the turbulent kinetic energy shows only a weak dependence on ReABCRe_{ABC}, the single-phase (M=0)(M=0) cases remain roughly constant (around K6CLK^{\prime}\approx 6CL), and flows with fixed (M=1) particles show a large attenuation of KK^{\prime} for all ReABCRe_{ABC}. This agrees with the observations of Peng et al. [22], who made simulations of spherical particles in homogeneous isotropic turbulence and found that the attenuation of turbulent kinetic energy has little dependence on the Reynolds number.

Refer to caption
Figure 4: The dependence of the normalised dissipation ϵ/urms3\epsilon\mathcal{L}/u_{{rms}}^{3} on the Taylor-Reynolds number ReλRe_{\lambda}. Flows with spheres are marked blue on panel (a), flows with fibres are marked in orange on panel (b), and single-phase flows are marked in black on both panels. Each case is plotted using its marker, which is listed in table 1, 2, or 3. Dashed lines show the anomalous value of dissipation ϵ/urms3=0.4\epsilon\mathcal{L}/u_{{rms}}^{3}=0.4 measured by Donzis et al. [34]. Solid lines show fits of equation 10 with A=0.2A=0.2. The fitted values of BB are given in the inset, where Donzis’ result (B=92)(B=92) is marked with an “X”.

The dissipative anomaly is a well-studied feature in single-phase turbulent flows, first proposed by Taylor [35]. It states that, even in the limit of vanishing viscosity (ReλRe_{\lambda}\to\infty), the normalised dissipation ϵ/urms3\epsilon\mathcal{L}/u_{{rms}}^{3} remains finite. The integral length scale is given by,

=π2urms20Eκdκ,\mathcal{L}=\frac{\pi}{2u_{{rms}}^{2}}\int_{0}^{\infty}\frac{E}{\kappa}\mathrm{d}\kappa, (9)

where EE is the turbulent kinetic energy spectrum and κ\kappa is the wavenumber. In 2005, Donzis et al. [34] parametrized the dissipative anomaly, based on the mathematical derivation of Doering and Foias [36], they fit the function

ϵurms3=A[1+1+(B/Reλ)2]\frac{\epsilon\mathcal{L}}{u_{{rms}}^{3}}=A\left[1+\sqrt{1+(B/Re_{\lambda})^{2}}\right] (10)

to a number of single-phase flows, obtaining A=0.2A=0.2 and B=92B=92. Figure 4 shows the dependence of the normalised dissipation for our flows. We see that as ReλRe_{\lambda}\to\infty, flows with spheres and fibres at all mass fractions appear to converge to the same anomalous value of dissipation ϵ/urms30.4\epsilon\mathcal{L}/u_{{rms}}^{3}{\to}0.4. Hence, we fit equation 10 with A=0.2A=0.2 to each mass fraction of spheres and fibres, obtaining a value for BB in each case, shown in the insets of figure 4. The fit to our single phase flows agrees closely with Donzis et al. [34]’s result. However, both spheres and fibres cause an increase in the value of BB as their mass fraction increases, with fibres producing roughly double the effect. To understand how spheres and fibres modify the dissipation in the flow, we must look at how they transport energy from large to small scales into the dissipative range.

III.2 Scale-by-scale results

Refer to caption
Figure 5: Energy spectra of all cases, normalised using the forcing amplitude CC and length LL. On the left, we show flows with spheres, and the vertical grey lines show the wavenumber κc=2π/c\kappa_{c}=2\pi/c of the sphere diameter. On the right, we show flows with fibres, and we shade the region between the wavenumber κc\kappa_{c} of the fibre length and the wavenumber κd=2π/d\kappa_{d}=2\pi/d of the fibre diameter. (a) and (b) show ReABC=894Re_{ABC}=894 and ReABC=224Re_{ABC}=224, with the latter shifted downwards by a factor of 100 on the y-axis, for various mass fractions. (c) and (d) show fixed (M=1)(M=1) and single phase (M=0)(M=0) cases for various Reynolds numbers. Each case is marked at the wavenumber corresponding to the Taylor length scale, using the marker listed in table 1, 2, or 3.

Figure 5 shows each flow’s turbulent kinetic energy spectrum EE. Single-phase flows exhibit the canonical Kolmogorov scaling Eκ5/3E\sim\kappa^{-5/3} for one or two decades, depending on the forcing Reynolds number. Adding spheres reduces EE at wavenumbers up to the sphere diameter (κ<κc\kappa<\kappa_{c}) and increases EE in the dissipative range. We mention, in passing, the oscillations at the wavenumber of the sphere diameter; these are an artefact resulting from the discontinuity in the velocity gradient at the sphere boundary [19]. The addition of fibres causes a reduction in EE across a broad band of wavelengths (κL100\kappa L\lesssim 100) and a pronounced increase in the dissipative range. As was previously seen by Olivieri et al. [10], the energy scaling EκβE\sim\kappa^{-\beta} becomes flatter as the mass fraction MM of fibres increases. Figure 5 also shows that both spheres and fibres cause the Taylor length scale to shift to higher wavenumbers due to the increase of the energy in the dissipative scales due to the presence of particles.

Refer to caption
Figure 6: Scale-by-scale energy balance for all cases. Panels (a) and (b) show ReABC=894Re_{ABC}=894 cases for various mass fractions. Panels (c) and (d) show ReABC=224Re_{ABC}=224 cases for various mass fractions. Panels (e) and (f) show fixed (M=1M=1) and single-phase (M=0M=0) cases for various Reynolds numbers. On the left, we show flows with spheres, and the vertical grey lines show the wavenumber κc=2π/c\kappa_{c}=2\pi/c of the sphere diameter. On the right, we show flows with fibres, and we shade the region between the wavenumber κc\kappa_{c} of the fibre length and the wavenumber κd\kappa_{d} of the fibre diameter. For each case, we plot three terms: the Dissipation DD, the energy flux Π\Pi due to convection, and the energy flux Πsf\Pi_{sf} due to the solid-fluid coupling. Each curve is marked where the gradient is largest, using the marker listed in table 1, 2, or 3.

Figure 6 shows how each term in the Navier-Stokes equation interacts with the energy spectrum, as expressed by the spectral energy balance,

inj(κ)+Π(κ)+Πsf(κ)+D(κ)=ϵ,\mathcal{F}_{inj}(\kappa)+\Pi(\kappa)+\Pi_{sf}(\kappa)+{D}(\kappa)=\epsilon, (11)

where inj,Π,Πsf,\mathcal{F}_{inj},\Pi,\Pi_{sf}, and DD are the rate of energy transfer by the ABC forcing, advection, solid-fluid coupling, and dissipation, respectively. See the supplementary information of Ref. [37] for a derivation of this equation. For the single-phase flows, energy is carried by the advective term Π\Pi from large to small scales, where it is dissipated by the viscous term DD. When particles are added, we see that the solid-fluid coupling term Πsf\Pi_{sf} acts as a “spectral shortcut” [38, 10]; it bypasses the classical turbulent cascade, removing energy from large scales and injecting it at the length scale of the particles, through their wakes. In keeping with the spectra, the power of the solid-fluid coupling (shown by the peak value of Πsf\Pi_{sf}) increases with mass fraction MM of both spheres and fibres. For spheres, the coupling Πsf\Pi_{sf} dominates only at wavenumbers less than that of the sphere diameter κ<κc\kappa<\kappa_{c} (i.e., large length scales). Around κc\kappa_{c}, the sphere wake returns energy to the fluid, and the classical cascade resumes for κ>κc\kappa>\kappa_{c}. For fibres instead, Πsf\Pi_{sf} extends deep into the viscous range. In fact, markers on the Πsf\Pi_{sf} curves show that spheres mostly inject energy around the length scale of their diameter, and fibres mostly inject energy at a length scale between their thickness and length. The images of dissipation in figure 1 support this observation; around the spheres, we see wakes comparable in size to the sphere diameter, while around the fibres, we see wakes comparable in size to the fibre thickness. We presume that the fluid-sphere coupling Πsf\Pi_{sf} would extend into the viscous range if the sphere diameter was smaller. Lastly, we consider the effect of Reynolds number on energy transfers in the flow, reducing ReABCRe_{ABC} limits the range of scales at which the advective flux Π\Pi occurs for single-phase flows. However, ReABCRe_{ABC} has little effect on the wavenumber range of the solid-fluid coupling term Πsf\Pi_{sf}, suggesting that the size of the particle wakes is governed mainly by particle geometry, with a lesser effect from fluid properties like viscosity.

Refer to caption
Figure 7: Structure functions of order p=2p=2 (dashed lines) and p=6p=6 (solid lines). Panels (a) and (b) show cases with ReABC=894Re_{ABC}=894 for various mass fractions. Panels (c) and (d) show cases with ReABC=224Re_{ABC}=224 for various mass fractions. Panels (e) and (f) show fixed (M=1M=1) and single-phase (M=0M=0) cases for various Reynolds numbers. Flows with spheres are on the left, and flows with fibres are on the right. Grey lines show the scalings predicted by Kolmogorov [39] for the single-phase structure functions. We mark each line at r=cr=c using the marker listed in table 1, 2, or 3.

Moving from wavenumber space to physical space, we show the longitudinal structure functions

Sp(r)[r^iui(𝐱+𝐫,t)r^iui(𝐱,t)]p𝐱,𝐫^,tS_{p}{(r)}\equiv\langle\left[\hat{r}_{i}u_{i}(\mathbf{x}+\mathbf{r}{,t})-\hat{r}_{i}u_{i}(\mathbf{x}{,t})\right]^{p}\rangle_{{\mathbf{x},\mathbf{\hat{r}},t}} (12)

of each flow in figure 7, where 𝐫\mathbf{r} is a separation vector between two points in the flow, it has magnitude rr and direction 𝐫^\mathbf{\hat{r}}, and angled brackets show an average over space 𝐱\mathbf{x}, direction 𝐫^\mathbf{\hat{r}}, and time t. For the second moment (p=2)(p=2), the single-phase flows closely follow Kolmogorov’s scaling [39] in the inertial range (rηr\gg\eta). However, when particles are added, S2S_{2} decreases relative to the single-phase case. For spheres, the decrease occurs for rcr\gtrsim c, while for fibres, it occurs at even smaller separations rr. Much like the EkβE\sim k^{-\beta} scalings seen in figure 5, the decreased regions in figure 7 show scalings S2rζ2S_{2}\sim r^{\zeta_{2}}, which become flatter for larger mass fractions MM. These energy spectrum and structure-function scalings are, in fact, just observations of the same phenomenon, as the exponents are related by ζ2=β1\zeta_{2}=\beta-1 for any differentiable velocity field [40, p.232]. In the higher moment structure function (p=6p=6), the flattening of the S6rζ6S_{6}\sim r^{\zeta_{6}} scaling by the particles is yet more apparent, indicating that the tails of the probability distribution of r^iui(𝐱+𝐫)r^iui(𝐱)\hat{r}_{i}u_{i}(\mathbf{x}+\mathbf{r})-\hat{r}_{i}u_{i}(\mathbf{x}) become wider for smaller separations rr. In other words, extreme values become more common, and the velocity field becomes more intermittent in space. In the case of single-phase homogeneous-isotropic turbulence, the intermittency of the velocity field has been linked to the non-space-filling nature of the fluid dissipation [41]. Next, we explore this link in the case of particle-laden turbulence by investigating the intermittency of dissipation ϵ\epsilon in space.

Refer to caption
Figure 8: The fluid dissipation averaged in balls of radius ll and squared. We mark each case at l=ηl=\eta using the marker listed in table 1, 2, or 3. We shade the regions where ϵl2\langle\epsilon_{l}^{2}\rangle shows a strong dependence on ll, for sphere-laden flows (a) this is 0.1Ll0.2L0.1L\leq l\leq 0.2L, whereas for fibre-laden flows (b) this is 0.04Ll0.08L0.04L\leq l\leq 0.08L.

Using the method described by Frisch [41, p.159], Meneveau and Sreenivasan [42] and others, we obtain the quantity ϵl\epsilon_{l} by averaging the viscous dissipation within a spherical region of radius ll (hereafter, we refer to this averaging region as a “ball” to avoid confusion with the particles). To observe the variation of ϵl\epsilon_{l} we take its square and average over many balls of radius ll at different locations and times, where the average is denoted using angled brackets \langle\cdot\rangle. Figure 8 shows how ϵl2\langle\epsilon_{l}^{2}\rangle depends on the radius ll of the balls. At large radii (l)(l\to\infty), ϵl\epsilon_{l} is by definition equivalent to the bulk value ϵ\epsilon, and so ϵl2/ϵ2\langle\epsilon_{l}^{2}\rangle/\epsilon^{2} tends to unity. However, for l<Ll<L, we see that ϵl2/ϵ2>1\langle\epsilon_{l}^{2}\rangle/\epsilon^{2}>1 in all of the cases plotted. This increase indicates that there are localised regions of high dissipation in the fluid. In other words, the dissipation fields are intermittent in space. For the single-phase flows, ϵl2\langle\epsilon_{l}^{2}\rangle reaches a plateau for l<ηl<\eta, giving an indication of the size of these flow structures. For the sphere-laden flows, we see that ϵl2\langle\epsilon_{l}^{2}\rangle reaches almost 100 times the bulk value and remains roughly constant for small radii (l<0.05L)(l<0.05L). This suggests that the spherical particles are creating high dissipation structures in the fluid, and balls with l<0.05Ll<0.05L fit entirely inside these structures. That is, reducing the ball radius below l=0.05Ll=0.05L has minimal effect on ϵl2\langle\epsilon_{l}^{2}\rangle because the majority of balls are sampling either entirely inside a high dissipation structure or entirely outside. In the case of fibres, high dissipation structures also produce an increase in ϵl2\langle\epsilon_{l}^{2}\rangle for small ll; here no plateau is seen, implying that even the smallest balls (l=6×103L)(l=6\times 10^{-3}L) cannot fit inside the high dissipation structures created by the fibres.

Refer to caption
Figure 9: Multifractal spectra of the dissipation in regions containing particles (dashed lines) and regions not containing particles (solid lines). Spectra in the left panels (a,c,e) were calculated using balls of radius 0.1Ll0.2L0.1L\leq l\leq 0.2L. Spectra in the right panels (b,d,f) were calculated using balls of radius 0.04Ll0.08L0.04L\leq l\leq 0.08L. We mark each line case where FF is maximum and at F=2F=2 using the marker listed in table 1, 2, or 3. Panels (a) and (b) show cases with ReABC=894Re_{ABC}=894 for various mass fractions. Panels (c) and (d) show cases with ReABC=224Re_{ABC}=224 for various mass fractions. Panels (e) and (f) show fixed (M=1M=1) and single-phase (M=0M=0) cases for various Reynolds numbers. Panel (e) shows a two-dimensional sheet intersecting with a ball, and panel (f) shows a one-dimensional tube intersecting with a ball.

To measure the fractal dimension of the high dissipation structures, we take a range of moments 6q6-6\leq q\leq 6 of ϵl\epsilon_{l}. This quantity has a scaling behaviour

ϵlqlτ,\langle\epsilon_{l}^{q}\rangle\sim l^{\tau}, (13)

if the dissipation field is multifractal [41]. Indeed, for the sphere-laden flows in figure 8a, ϵl2\langle\epsilon_{l}^{2}\rangle has a strong dependence on ll in the range 0.1Ll0.2L0.1L\leq l\leq 0.2L, and for the fibre-laden flows in figure 8b, ϵl2\langle\epsilon_{l}^{2}\rangle has a strong dependence on ll in the range 0.04Ll0.08L0.04L\leq l\leq 0.08L. Hence, we calculate τ\tau by making lines of best fit in these ranges, and we obtain the multifractal spectra F(α)F(\alpha) with a Legendre transformation

α\displaystyle\alpha =dτdq+1,\displaystyle=\frac{d\tau}{dq}+1, (14)
F\displaystyle F =q(α1)τ+3.\displaystyle=q(\alpha-1)-\tau+3. (15)

Figure 9 shows the multifractal spectra for all cases. Similarly to Mukherjee et al. [43], we split our analysis into separate regions: solid curves result from an ensemble average over balls not containing particles, while dashed curves result from an ensemble average over balls containing particles. We see that the multifractal spectra of the single-phase flows and the regions not containing particles have peaks at α1,F3\alpha\approx 1,F\approx 3, showing there is a background of space-filling dissipation in these regions [41, p.163]. In addition, the presence of tails in the spectra shows that the dissipation fields are not self-similar and violate the hypotheses made by Kolmogorov [39] for a turbulent flow. This can explain the flattening of the high-order structure functions (S6S_{6} in figure 7) relative to Kolmogorov’s predictions. We note in passing that the single-phase flows with lower Reynolds number (ReABC=55.9Re_{ABC}=55.9 and ReABC=112Re_{ABC}=112) show narrower tails in figure 9f, as viscous effects are present here.

The multifractal spectra of the regions containing spheres (blue dashed lines in figure 9) have peaks which are shifted to the left. This shift can be explained by considering the high dissipation in the boundary layers around the spheres. As imaged in figure 12, the boundary layers have a higher dissipation rate than the surrounding fluid, and are confined in space to relatively thin sheets. In figure 9e we sketch a ball of radius ll intersecting with a sphere’s boundary layer which has thickness δl\delta\ll l. The volume of the high dissipation region inside the ball is πl2δ\pi l^{2}\delta, hence the total dissipation in the ball roughly scales as l2\sim l^{2}. When we average over the volume of the ball (4πl3/3{4\pi l^{3}/3}) and take the qqth moment, we obtain ϵlql2ql3q\langle\epsilon_{l}^{q}\rangle\sim l^{2q}l^{-3q}. Comparing with equation 13, this scales as τ=q\tau=-q. Applying a Legendre transformation (equations 14 and 15) this corresponds to the point α=0,F=3\alpha=0,F=3, which is roughly where we find the peaks in the multifractal spectra of the regions containing spheres. This reinforces our observation that the dissipation in the regions around the spherical particles is confined to sheet-like structures.

The multifractal spectra of the regions containing fibres (orange dashed lines in figure 9) have peaks which are shifted even further to the left. Again, we explain this by considering the boundary layers around the particles. In figure 9f we sketch a ball of radius ll intersecting with a fibre’s boundary layer which has radius δl\delta\ll l. In this case, the volume of the high dissipation region is πδ2l\pi\delta^{2}l. Thus, the average dissipation in the ball scales as ϵlqlql3q\langle\epsilon_{l}^{q}\rangle\sim l^{q}l^{-3q}, so the resulting scaling is τ=2q\tau=-2q, which (by equations 14 and 15) corresponds to the point α=1,F=3\alpha=-1,F=3. In fact, in figure 9f, we see that the ensemble average over balls containing M=1M=1 fibres produces peaks around α=0.5\alpha=-0.5, suggesting that the wakes around the fibres are not exactly thin tubes, but more space-filling structures with a fractal dimension between one and two. This is also supported by the images in figure 12, as wakes are seen to extend behind the fibres.

The middle and upper panels of figure 9 show that (at ReABC=224Re_{ABC}=224 and ReABC=894Re_{ABC}=894) the sphere- and fibre-containing multifractal spectra tend toward the single phase multifractal spectrum as MM is reduced. This is expected as the particle-fluid relative velocity is reduced and the wakes become less prominent compared to the background space-filling turbulence. On the other hand, the ensembles of balls not containing particles produce multifractal spectra with low α\alpha tails that extend beyond those of the single-phase multifractal spectra, suggesting that the wakes of spheres and fibres extend into the bulk of the fluid and that they retain their non-space-filling nature.

III.3 Local flow structures

Refer to caption
Figure 10: Histograms of the alignment of the vorticity unit vector 𝝎^\boldsymbol{\hat{\omega}} with the eigenvectors 𝐚^𝟏,𝐚^𝟐,𝐚^𝟑\mathbf{\hat{a}_{1}},\mathbf{\hat{a}_{2}},\mathbf{\hat{a}_{3}} of the strain-rate tensor. Flows with spheres are on the left, and flows with fibres are on the right. Panels (a) and (b) show cases with ReABC=894Re_{ABC}=894 for various mass fractions. Panels (c) and (d) show cases with ReABC=224Re_{ABC}=224 for various mass fractions. Panels (e) and (f) show fixed (M=1M=1) and single-phase (M=0M=0) cases for various Reynolds numbers. Each case is plotted using its marker, which is listed in table 1, 2, or 3. We mark each line at |𝝎^𝐚^𝐢|=0,1|\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{i}}|=0,1 and |𝝎^𝐚^𝐢|\langle|\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{i}}|\rangle. The inset of panel (c) shows the directions of vorticity and the eigenvectors of the strain rate in a laminar shear flow, where 𝐚^𝟏\mathbf{\hat{a}_{1}} and 𝐚^𝟑\mathbf{\hat{a}_{3}} are confined to the shear plane, while 𝐚^𝟐\mathbf{\hat{a}_{2}} and 𝝎^\boldsymbol{\hat{\omega}} are perpendicular to it.

To look more closely at the flow structures produced by the particles, we compute the alignment of the unit-length eigenvectors 𝐚^𝟏,𝐚^𝟐,𝐚^𝟑\mathbf{\hat{a}_{1},\hat{a}_{2},\hat{a}_{3}} of the strain-rate tensor sij(iuj+jui)/2{s_{ij}\equiv(\partial_{i}u_{j}+\partial_{j}u_{i})/2} with the vorticity 𝝎εijkjuk𝐞^i{\boldsymbol{\omega}\equiv{\varepsilon_{ijk}}\partial_{j}u_{k}\mathbf{\hat{e}}_{i}} [44, 45], where 𝐞^𝟏,𝐞^𝟐,𝐞^𝟑\mathbf{\hat{e}_{1},\hat{e}_{2},\hat{e}_{3}} are the Cartesian unit vectors, and εijk\varepsilon_{ijk} is the Levi-Civita symbol. We choose 𝐚^𝟏\mathbf{\hat{a}_{1}} to be the eigenvector corresponding to the largest eigenvalue. For an incompressible fluid, the three eigenvalues of sijs_{ij} sum to zero, so the largest eigenvalue is never negative. Hence, 𝐚^𝟏\mathbf{\hat{a}_{1}} aligns with the direction of elongation in the flow. Similarly, we choose the eigenvector 𝐚^𝟑\mathbf{\hat{a}_{3}} to correspond with the smallest eigenvalue, which is never positive, so 𝐚^𝟑\mathbf{\hat{a}_{3}} aligns with the direction of compression in the flow. Finally, due to the symmetry of the strain-rate tensor, 𝐚^𝟐\mathbf{\hat{a}_{2}} is orthogonal to the other two eigenvectors and, depending on the flow, there can be compression or elongation along its axis. Figure 10 shows probability-density functions PP of the scalar product of the unit-length vorticity 𝝎^\boldsymbol{\hat{\omega}} with 𝐚^𝟏,𝐚^𝟐\mathbf{\hat{a}_{1},\hat{a}_{2}} and 𝐚^𝟑\mathbf{\hat{a}_{3}} for all of the flows studied. The quantity 𝝎^𝐚^𝐢\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{i}} is simply the cosine of the angle between the two vectors, and we take the modulus because 𝐚^𝐢\mathbf{\hat{a}_{i}} and 𝐚^𝐢-\mathbf{\hat{a}_{i}} are degenerate eigenvectors. In the single-phase flows, PP is maximum at |𝝎^𝐚^𝟐|=1|\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{2}|}=1, that is, we see a strong alignment of vorticity with the intermediate eigenvector 𝐚^𝟐\mathbf{\hat{a}_{2}}. This is a well-known feature of turbulent flows and has been attributed to the axial stretching of vortices [44]. Also in keeping with previous observations of single-phase turbulence, we see that the first eigenvector 𝐚^𝟏\mathbf{\hat{a}_{1}} shows very little correlation with 𝝎^\boldsymbol{\hat{\omega}}, and the last eigenvector is mostly perpendicular to the vorticity, producing a peak in PP at 𝝎^𝐚^𝟑=0\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{3}}=0. When spheres are added, vorticity aligns even more strongly with the intermediate eigenvector 𝐚^𝟐\mathbf{\hat{a}_{2}}, and becomes more perpendicular to the first and last eigenvectors 𝐚^𝟏\mathbf{\hat{a}_{1}} and 𝐚^𝟑\mathbf{\hat{a}_{3}}. This change is consistent with the existence of a shear layer at the surface of the spheres, or in the wakes behind the spheres. To demonstrate this, we draw a laminar shearing flow and label the directions of 𝐚^𝟏,𝐚^𝟐,𝐚^𝟑\mathbf{\hat{a}_{1},\hat{a}_{2},\hat{a}_{3}} and 𝝎^\boldsymbol{\hat{\omega}} in the inset of figure 10. The compression 𝐚^𝟑\mathbf{\hat{a}_{3}} and extension 𝐚^𝟏\mathbf{\hat{a}_{1}} are in the plane of the shear flow, while the vorticity 𝝎^\boldsymbol{\hat{\omega}} and the intermediate eigenvector 𝐚^𝟑\mathbf{\hat{a}_{3}} are perpendicular to the shear plane; this gives rise to the 𝝎^𝐚^𝟏=0,\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{1}}=0, |𝝎^𝐚^𝟐|=1|\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{2}}|=1 and 𝝎^𝐚^𝟑=0\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{3}}=0 modes in the sphere-laden flows in figure 10. The larger mass fraction spheres have a greater effect, as their motion relative to the fluid is larger. Also, the spheres’ effect is more pronounced at lower ReABCRe_{ABC}, which can be explained by an increased thickness in the shear layer as the fluid viscosity increases. For fibre-laden flows, peaks are also seen at 𝝎^𝐚^𝟏=0,\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{1}}=0, |𝝎^𝐚^𝟐|=1|\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{2}}|=1 and 𝝎^𝐚^𝟑=0\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{3}}=0, but the peaks are spread over a larger range of angles, presumably due to the high curvature of the fibre surfaces, which adds variance to the local direction of vorticity and shear. At small Reynolds numbers (ReABC=55.9Re_{ABC}=55.9 and 112112) the peaks are widest, and P(|𝝎^𝐚^𝐢|)P(|\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{i}}|) becomes an almost uniform distribution.

Refer to caption
Figure 11: Joint histograms of the invariants QQ and RR of the fluid velocity gradient tensor. Panels (a) and (b) show cases with ReABC=894Re_{ABC}=894 for various particle mass fractions. Panels (c) and (d) show cases with ReABC=224Re_{ABC}=224 for various particle mass fractions. Panels (e) and (f) show fixed (M=1M=1) and single-phase (M=0M=0) cases for various Reynolds numbers. Each case is plotted using its marker, which is listed in table 1, 2, or 3. The number on the marker shows the value of logp\log p at each contour, where pp is the joint probability density function. The grey curves show where the discriminant is zero (equation 18). The grey shaded regions in the left panels are the loci of equation 19, where we observe an increase pp due to the presence of spheres.

The local topology of a flow can be described entirely using the three principle invariants of the velocity gradient tensor iuj\partial_{i}u_{j} [46]. The first invariant iui\partial_{i}u_{i} is not interesting in our case because it is zero for an incompressible fluid. However, the second invariant,

Q=14ωiωi12sijsij,Q=\frac{1}{4}\omega_{i}\omega_{i}-\frac{1}{2}s_{ij}s_{ij}, (16)

expresses the balance of vorticity and strain, while the third invariant,

R=14ωisijωj13sijsjkski,R=\frac{1}{4}\omega_{i}s_{ij}\omega_{j}-\frac{1}{3}s_{ij}s_{jk}s_{ki}, (17)

is the balance of the production of vorticity with the production of strain [47]. The eigenvalues Λ\Lambda of iuj\partial_{i}u_{j} are the roots of the polynomial equation

Λ3+QΛ+R=0.\Lambda^{3}+Q\Lambda+R=0. (18)

Figure 11 shows the joint probability distributions p(R,Q)p(R,Q) for all of our flows. We also plot a line where the discriminant of equation 18 is zero: 27R2/4+Q3=027R^{2}/4+Q^{3}=0. Above this line, iuj\partial_{i}u_{j} has one real and two complex eigenvalues and vortices dominate the flow. Below this line, all three eigenvalues are real, and the flow is dominated by strain. The single-phase distributions have tails in the top-left and bottom-right quadrants; these correspond to stretching vortices and regions where the flow compresses along one axis, respectively [46]. When spheres are added, both QQ and RR are reduced, as the non-slip boundary condition at their surfaces dampens the flow structures. Incidentally, a similar QQ and RR reduction has been seen with droplets [48]. However, we see that the addition of spheres causes p(R,Q)p(R,Q) to increase in the shaded regions in the left panels of figure 11. These regions are defined by

27R24+Q3<4ϵ3ν3.\frac{27R^{2}}{4}+Q^{3}<-\frac{4\epsilon^{3}}{\nu^{3}}. (19)

These shaded regions correspond to a strain-dominated flow. In fact, for pure strain, we have Q=sijsij/2Q=-s_{ij}s_{ij}/2 and R=0R=0. Hence we can substitute for QQ and RR in equation 19 to estimate that the dissipation is

2νsijsij=4νQ<44/3ϵ6ϵ{2\nu s_{ij}s_{ij}=-4\nu Q<4^{4/3}\epsilon\approx 6\epsilon} (20)

in the straining flow. That is, the dissipation in regions around the spheres is roughly six times greater than the mean dissipation. Fibres also create straining regions in the flow, seen by the increased probability density functions in the lower parts (27R2/4+Q3<027R^{2}/4+Q^{3}<0) of the right-hand panels in figure 11. As the fibre mass fraction is increased, the variance of RR is reduced, showing the production of strain and vorticity is suppressed by the fibres. At high fibre mass fractions the distributions are approximately symmetrical in RR, in other words, QQ and RR are uncorrelated (much like the decorrelation of vorticity and strain observed for fibre-laden flows in figure 10). This suggests that the small fibre diameter produces many small scale structures which disrupt the typical turbulent flow. For both particle types, increasing the Reynolds number ReABCRe_{ABC} reduces the effect of the particles.

Refer to captionRefer to caption
Figure 12: Surfaces where the dissipation is 6ϵ6\epsilon, coloured according to the value of the flow topology parameter Σ\Sigma, which distinguishes rotational flow (green), shearing flow (white), and straining flow (purple). We show ReABC=224Re_{ABC}=224 cases with fixed spheres (left), single-phase (centre), and fixed fibres (right). The boxes have side length LL.

In figure 12 we visualise isosurfaces where the local fluid dissipation is six times its bulk value, i.e., where 2νsijsij=6ϵ2\nu s_{ij}s_{ij}=6\epsilon. As predicted by equation 20, the dissipation is greater than 6ϵ6\epsilon in the near-particle regions. And in keeping with our measurements of the fractal dimension of the dissipation field (figure 9), the high dissipation region near the sphere is approximately sheet-like and two-dimensional, while the high dissipation regions near the fibres are approximately tube-like and one-dimensional. In both cases, wakes extend behind the particles, which contribute to the fractal dimension of the dissipation fields. We see the single-phase dissipation structures take various shapes and sizes, giving rise to the multifractal behaviour discussed above. In figure 12 we colour the dissipation isosurfaces according to the local value of the flow topology parameter [49],

Σ2sijsijωiωi2sijsij+ωiωi,\Sigma\equiv\frac{2s_{ij}s_{ij}-\omega_{i}\omega_{i}}{2s_{ij}s_{ij}+\omega_{i}\omega_{i}}, (21)

which is essentially the second invariant of the velocity gradient tensor QQ normalised so that it is bounded between -1 and 1. Regions where Σ=1\Sigma=-1 (coloured green) correspond to pure rotational flow, regions where Σ=0\Sigma=0 (coloured white) correspond to pure shear, and regions where Σ=1\Sigma=1 (coloured purple) correspond to pure straining flow. From figure 12 we see fibres produce straining flow on the upstream side, and shearing flow in wakes on the downstream side, this manifests in figure 11 as an increase in the spread of QQ for the fibre-laden flows. At the surface of the sphere we see a mostly shearing flow, which supports our observations of strong shear-vorticity alignment (|𝝎^𝐚^𝟐|=1)(|\boldsymbol{\hat{\omega}}\cdot\mathbf{\hat{a}_{2}}|=1) in figure 10.

IV Conclusion

We made simulations of finite-size isotropically- and anisotropically-shaped particles (spheres and fibres with size cc within the inertial range of scales) in turbulence at various Reynolds numbers and solid mass fractions. We used bulk flow statistics to show that particles reduce turbulence intensity relative to the single-phase case at all Reynolds numbers, with the fibres producing a more significant reduction effect than the spheres. Regarding the trend in anomalous dissipation with Reynolds number, we see that the particle-laden flows tend to behave like single-phase flows as ReλRe_{\lambda}\to\infty, and the value of anomalous dissipation is ϵ/urms30.4\epsilon\mathcal{L}/u_{{rms}}^{3}\approx 0.4 for both single-phase and particle-laden flows. We see that spheres slow the convergence rate to the anomalous value, and fibres slow it further. Next, we analysed the flow at each scale of the simulation. Spheres and fibres provide a “spectral shortcut” to the flow, removing energy from the largest scales and injecting it at smaller scales. Spheres’ effect is mainly limited to the large scales, and they provide a spectral shortcut down to the length scale of their diameter. Fibres’ effect, on the other hand, occurs down to the scale of the fibre thickness, even at low Reynolds numbers when it is deep in the viscous range (d<ηd<\eta). This shortcut of energy to the dissipative scales slows the dissipation’s convergence to its anomalous value as ReλRe_{\lambda}\to\infty. Our scale-by-scale analysis also showed that particles cause the velocity field to become more intermittent in space. Multifractal spectra of the near-particle dissipation show that spheres enhance dissipation in two-dimensional sheets, and fibres enhance the dissipation in structures with a dimension greater than one and less than two. These lower dimensional structures are a possible source of intermittency in the flow. Zooming in closer to the flow, we looked at the shape of flow structures using their vorticity and shear. We saw that the particles enhance local shear, spheres suppress vortical flow structures, and fibres produce intense vortical and shearing structures, which overcome the usual vortex stretching behaviour. As Reynolds number increases, the flow structures created by the particles become less significant relative to the background turbulence.

To reiterate and answer our questions from section I; at what Reynolds numbers does turbulence modulation emerge? We see turbulence modulation at the lowest Reynolds number investigated (Reλ=12.8)(Re_{\lambda}=12.8). This gives us reason to believe that particles modulate turbulence in even the most weakly turbulent flows (Reλ0)(Re_{\lambda}\to 0). Secondly, does the modulation effect persist as ReRe\to\infty? The values of normalised dissipation and many other statistics presented here do indeed converge on the single-phase result as ReRe\to\infty. However, the modulation of turbulent kinetic energy KK^{\prime} appears to have little to no dependence on the Reynolds number. Larger Reynolds numbers must be investigated to test whether the attenuation of KK^{\prime} goes to zero as ReABCRe_{ABC}\to\infty. Finally, how do the particle’s characteristic lengths impact the scales of the turbulent flow? Both spheres and fibres take energy from the flow at scales larger than their size. Spheres re-inject energy around the characteristic length of their diameter cc, leaving the smaller scales relatively unperturbed, Indeed, the effect of the spheres on the local flow structures tends to zero as the ratio c/ηc/\eta increases. In contrast, fibres re-inject energy around the characteristic length of their thickness dd. The fibre thickness is at a small scale, so local flow structures are disrupted. None of the flow statistics shows any particular discerning feature at the length scale cc of the fibres.

These results relate to various environmental particle-laden flows, such as microplastics in the ocean, volcanic ash clouds, and sandstorms. We have explored the two extremes of isotropic and anisotropic particles, but further work is needed to investigate how intermediate aspect ratio particles such as ellipsoids interact with turbulent flows.

V Acknowledgements

The research was supported by the Okinawa Institute of Science and Technology Graduate University (OIST) with subsidy funding from the Cabinet Office, Government of Japan. The authors acknowledge the computer time provided by the Scientific Computing section of Research Support Division at OIST and the computational resources of the supercomputer Fugaku provided by RIKEN through the HPCI System Research Project (Project IDs: hp210229 and hp210269). SO acknowledges the support of grants FJC2021-047652-I and PID2022-142135NA-I00 by MCIN/AEI/10.13039/501100011033 and European Union NextGenerationEU/PRTR.

VI Data availability

References