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

Also at ]Ministry of Education Key Laboratory of Tropical Atmosphere-Ocean System, Zhuhai, China

Sharp depletion of radial distribution function of particles due to collision and coagulation inside turbulent flow: a systematic study.

Xiaohui Meng    Ewe-Wei Saw [email protected] [ School of Atmospheric Sciences and Guangdong Province Key Laboratory for Climate Change and Natural Disaster Studies, Sun Yat-Sen University, Zhuhai, China
Abstract

The clustering (preferential concentration) and collision of particles in turbulent flow is a significant process in nature and technical applications. We perform direct numerical simulation (DNS) to study the clustering of small, heavy, monodisperse particles subject to collision-coagulation in turbulent flow (i.e., colliding particles always coagulate (coalesce) into large ones). We find that collision-coagulation causes the radial distribution function (RDF) of the particles to decrease strongly at particle separation distances rr close to the particle diameter dd. However, we observe that the RDF do not decrease indefinitely but approach a finite value in the limit of rdr\to d. We study how the characteristics of this “depletion zone” relate to the particle Stokes number (St), particle diameter, and the Reynolds number of the turbulent flow. A collision-induced modulation factor γc\gamma_{c} is defined to represent the degree of RDF depletion due to collisions-coagulation. In the region where γc(r)\gamma_{c}(r) is a quasi-power-law, we find that the corresponding power-law exponent c~1\tilde{c}_{1} only depends weakly on StSt. We also find that the overall trend of c~1\tilde{c}_{1} with respect to StSt is similar to that of the classical power-law exponent c1c_{1} appearing in the RDF of non-colliding particles, i.e., the exponent increase at small StSt, peak around St0.7St\approx 0.7, and decrease thereafter. The same qualitative trend is also observed for the limiting values of γc\gamma_{c} at rdr\to d. A complementary investigation on the Stokes number trend of the full RDF in the depletion zone is conducted. The slope of RDF appear constant for St1St\ll 1 but is changing when StSt is getting large. The position where the RDF starts to decrease is found to be StSt-dependent. We found that he depletion zone is insensitive to the flow Reynolds number and γc\gamma_{c} of different ReλRe_{\lambda} overlap. With changing particle diameter dd, the reduction of RDF occurs at scales that shift accordingly and always starts at around 2.4d3d2.4d-3d. We also find that the shape of γc(r)\gamma_{c}(r) is independent of changes in dd.

I Introduction

The fluctuation of particle concentration has a profound effect on inter-particle collision. This effect has a pivotal role in both natural sciences and industrial engineering. For example, the collision-coagulation process for small droplets (particles) determines their spatial and temporal size distribution. These microscale properties have significant influences on the formation of precipitation [1]. Small-scale turbulence in clouds makes an important contribution to the collision and coagulation of droplets [2] [1]. Another example is the formation of planetesimal. The collision of dust in protoplanetary disks sets the stage for planet formation. Research shows that the turbulent motion will concentrate dust in the dissipation scale, increasing the collision rate of dust particles [3] [4] [5].

The preferential concentration of inertia particle has become a prevalent research topic since the end of the 20th century. Squires and Eaton [6] found that the inertial particles concentrate preferentially in regions of low vorticity and high strain rate. The degree of particle clustering can be characterized via the radial distribution function (RDF), which is defined as the ratio of the probability of finding a particle pair at a distance rr normalized by the probability of the same event for random particle distribution. Reade & Collins [7] found that the clustering of small particles occurs on a scale that is much smaller than the Kolmogorov length scale and the RDF of particles follows a negative power law with the inter-particle separation distance rr. Chun et al. [8] developed a drift-diffusion theory to predict the RDF in turbulent flows for particles in the limit of small particle Stokes number. Their results indicate that RDF of particle is proportional to c0(r/η)c1c_{0}(r/\eta)^{c_{1}}, where η\eta is the Kolmogorov length scale. They also find that the exponent c1c_{1} is proportional to St2St^{2}. The Stokes number (St) is an important measurement for particle inertia, and is defined as the ratio of particle relaxation time τp\tau_{p} and the Kolmogorov time scale τη\tau_{\eta}. The dissipation-scale clustering of particles becomes stronger as the Stokes number increase and the RDF reaches a peak near the Stokes number of order unit [9]. This relationship between the RDF and the Stokes number is corroborated by both numerical and experimental studies [10] [11].

The preferential concentration of particle is expected to enhance particle collision but the calculation of collision kernel is still an open question. In the work of Sundaram and Collins [12], the RDF was first introduced into the formula of collision kernel [13]: K=4πd2g(d)W(d)K=4\pi d^{2}g(d)\langle W(d)\rangle, where dd is particle diameter, g(d)g(d) is the value of RDF at r=dr=d and W(d)\langle W(d)\rangle is the mean radial relative velocity of particle at r=dr=d. The works on the RDF of particles mentioned thus far had used the ghost particle assumption and ignored the effect of collision and coagulation among particles. Saw and Meng [14] found that the RDF will drop profoundly at rr close to dd in the presence of collision-coagulation. This finding is interesting because it highlights the importance of accounting for actual particle collision and it also questions the formula of collision kernel mentioned above.

In this paper, we use direct numerical simulation (DNS) to study the RDF of the inertial, colliding, particles. DNS, which solves the Navier-Stokes equation fully resolving the spatial and temporal scales of the problem without using any turbulence modeling, is an efficient numerical tool to study the particle-laden turbulent flow. The RDF of particle considering the effect of the collision-coagulation is investigated. The influence of the particle and turbulent parameters on the decreasing of RDF is also studied in this paper. The paper is organized as follows: section 2 provides a summary of simulation methods and the relevant turbulent and particle parameters. The statistical results and discussion are in section 3 and section 4. Finally, the results of the influence of the turbulent and particle parameters on the RDF are summarized in section 5.

II Simulation method

We performed direct numerical simulation (DNS) of the particle-laden turbulent flow. The incompressible Navier-Stokes equations are shown below.

ut+uu=1ρp+ν2u+f(x,t)\frac{\partial\vec{u}}{\partial t}+\vec{u}\cdot\nabla\vec{u}=-\frac{1}{\rho}\nabla p+\nu\nabla^{2}\vec{u}+\vec{f}(\vec{x},t) (1)
u=0\nabla\cdot\vec{u}=0 (2)

The N-S equations are numerically solved on N3N^{3} grids cube using a pseudo-spectral method on the periodic domain, the length of which is 2π2\pi. The turbulent velocity u\vec{u} is transformed from physical space to wavenumber space. The aliasing error rising from the convection part of N-S equation is removed by the 2/3-method [15]. f(x,t)\vec{f}(\vec{x},t) in the N-S equation is an external forcing conducted at low-wavenumber in order to maintain statistically stationary [16]. In order to study the influence of the (Taylor scaled) Reynolds number on RDF, simulations with different ReλRe_{\lambda} are conducted: Reλ=84Re_{\lambda}=84, 124124 and 189189, the detailed turbulent parameters are shown in Table 1. For different Reynolds number, the grid size is N=2563N=256^{3} (for Reλ=84Re_{\lambda}=84 and 124124) and 5123512^{3} (for Reλ=189Re_{\lambda}=189). The small scale resolution may be characterized by kmaxη=1.59k_{max}\eta=1.59, 1.21 and 1.38 respectively, where kmax=N2/3k_{max}=N\sqrt{2}/3 is the maximum resolved wavenumber magnitude. The 2-order Runge-Kutta method is used to conduct time advancement in N-S equation. The Courant number C=0.0248C=0.0248, 0.04010.0401 and 0.08650.0865.

Table 1: The DNS parameters and time-averaged statistics. NN is the simulation grid size, ν\nu is the kinematic viscosity of turbulence, ϵ\epsilon is the dissipation rate of turbulent flow, uu^{\prime} is the root-mean-square velocity of turbulent flow, λ\lambda is the Taylor length scale, η\eta and τη\tau_{\eta} are the Kolmogorov length and time scale, LL and TLT_{L} are the integral length and time scale, ReλRe_{\lambda} is the Taylor scaled Reynolds number.
NN ν\nu ϵ\epsilon uu^{\prime} λ\lambda η\eta τη\tau_{\eta} LL TLT_{L} ReλRe_{\lambda}
flow 1 256 0.001 0.0326 0.3519 0.2386 0.0132 0.1750 0.5073 1.4416 84
flow 2 256 0.001 0.1013 0.5684 0.2187 0.0100 0.0993 0.6151 1.0822 124
flow 2 512 0.001 0.9472 1.226 0.1544 0.0057 0.0325 0.7398 0.6034 189

The particles we take are small (the diameter of particle dd is smaller than the Kolmogorov length scale η\eta) and heavy (the particle’s density is larger than the flow’s). The gravitational effect and inter-particle hydrodynamic interactions are not included in our DNS model because only basic questions are discussed in this paper. Under these circumstances, the particles are only subjected to viscous Stokes drag force, and the motion equation of particles is shown below [17]:

dvdt=uvτp\frac{d\vec{v}}{dt}=\frac{\vec{u}-\vec{v}}{\tau_{p}} (3)

In which, v\vec{v} is the particle velocity, u(x,t)\vec{u}(\vec{x},t) is the fluid velocity at particle position. τp\tau_{p} is the particle inertia response time, defined τp=118ρpρd2ν\tau_{p}=\frac{1}{18}\frac{\rho_{p}}{\rho}\frac{d^{2}}{\nu} where ρp\rho_{p} and ρ\rho are density of particles and flow respectively, dd is the particle diameter and ν\nu is dynamic viscosity of turbulent flow. The linear interpolation method is used for interpolating flow’s velocity in particle position and the 2-order Runge-Kutta method with “exponential integrators” is used for time advancement [18].

Spherical and mono-dispersed particles are randomly introduced in the simulation. Particles collide when their volumes overlap and a new particle is formed conserving volume and momentum. New particles are injected continuously and randomly in the system so that the particle system is in a steady state after a transient period. In order to study the influence of particle parameters on RDF, particles with different Stokes numbers from 0.010.01 to 2.02.0 are introduced in each simulation. The size of particle is another important parameter related to particle collision. Three different sizes of particles are introduced in each simulation: d=13dd=\frac{1}{3}d_{*}, d=dd=d_{*} and d=3dd=3d_{*}, dd is the diameter of particle and d=9.4868×104d_{*}=9.4868\times 10^{-4}. The details of simulations are listed in Table 2. The statistics are calculated for mono-dispersed particles.

Table 2: Characteristics of the runs discussed here. ReλRe_{\lambda} is the Taylor microscale Reynolds number of the flow. dd is the particle diameter, d=9.4868×104d_{*}=9.4868\times 10^{4}. StSt is the particle Stokes number.
Run ReλRe_{\lambda} dd StSt
1 124 dd_{*} 0.01
2 124 dd_{*} 0.05
3 124 dd_{*} 0.1
4 124 dd_{*} 0.2
5 124 dd_{*} 0.5
6 124 dd_{*} 0.7
7 124 dd_{*} 1.0
8 124 dd_{*} 2.0
9 124 13d\frac{1}{3}d_{*} 0.1
10 124 3d3d_{*} 0.1
11 84 dd_{*} 0.1
12 189 dd_{*} 0.1

III Results and Discussion

III.1 Stokes number dependence

The statistics from Run 1 to Run 8 in Table 2 are used to study the influence of Stokes number on RDF. The RDFs for particles with different Stokes numbers are shown in Figure 1. What is striking in this figure is the strong decrease of RDF when the particle separation distance rr is close to the particle diameter dd. Figure 1 also show that there is an increase in the slope and the magnitude of RDF at the scales r/η110r/\eta\sim 1-10 when StSt is increased from 0.01 to 0.7. We note that the slope of the RDF for the St=1.0St=1.0 case is smaller than that of the St=0.7St=0.7 case, even though its values are everywhere larger. Beyond St=1.0St=1.0, the slope and magnitude are both decreasing from St=1.0St=1.0 to St=2.0St=2.0. At larger scales (r/η10r/\eta\sim 10), the RDF curves flatten and converge to 1. To elucidate the trend of RDF when the separation distance rr is close to the particle diameter dd, the RDFs are plotted as the function of rdr-d in Figure 2. The most interesting aspect of this graph is that the relationship between RDF and rdr-d exhibits a quasi-power-law trend in the range 4×105rd3×1044\times 10^{-5}\lesssim r-d\lesssim 3\times 10^{-4}. As rr continues to decrease towards the particle diameter, the slope of RDF gradually becomes smaller and approaches zero, i.e., the RDF approaches a plateau.

Refer to caption
Refer to caption
Figure 1: The RDFs versus r/ηr/\eta for particles with different Stokes number. The diameter of particle is d=9.49×104d=9.49\times 10^{-4} and the Taylor scaled Reynolds number is Reλ=124Re_{\lambda}=124. The RDF drops significantly when rr is close to dd.
Refer to caption
Refer to caption
Figure 2: The RDFs versus rdr-d for particles with different Stokes number, d=9.49×104d=9.49\times 10^{-4} is the particle’s diameter. The Taylor-scaled Reynolds number is Reλ=124Re_{\lambda}=124 in this case. Error bars represent one standard error. The RDF follows a quasi-power-law with rdr-d in the range 0.04drd0.3d0.04d\lesssim r-d\lesssim 0.3d and the slope of RDF decreases to zero gradually as rr continues to decrease.

Figure 3 illustrates more closely the trend of RDF as rr decrease towards dd, using the cases of St=0.1St=0.1, St=0.2St=0.2, and St=0.5St=0.5 as examples. This shows more clearly that after rdr-d is less than 4×1064\times 10^{-6}, RDF no longer decreases but levels off to a fixed value. The limiting value of the RDF at particle contact (r=dr=d) is proportional to the particle collision rate [12], thus a finite collision rate in our simulations implies that the value of RDF should be equal to a fixed value but not an infinitesimal value. The results shown above agree with this analysis.

We shall call the region where the RDF decreases, as seen in Fig. 3, the depletion zone. Subsequent discussions will mainly focus on this region.

Refer to caption
Figure 3: High resolution plots of RDFs versus rdr-d for cases of St=0.1St=0.1, 0.20.2, and 0.50.5, where the regime of smaller rdr-d values is resolved more clearly. In order to compare them clearly, the RDF for St=0.5St=0.5 is translated down vertically to half of its original height Error bars represent statistical error of one standard deviation. Within the range of uncertainty, the RDF no longer decreases after rd<0.04dr-d<0.04d and levels off to a fixed value.

III.2 The Collisional Modulation Factor

To further characterize this “depletion zone” of RDF due to particle collisions, following the work of (Saw and Meng, 2022) [14], we assume that the RDF could be factorized such that g(r)=γc(r)gn(r)g(r)=\gamma_{c}(r)g_{n}(r), where γc(r)\gamma_{c}(r) represent the effect of collision-coagulation in the form of a “modulation factor”, while gn(r)g_{n}(r) is the RDF for non-colliding (ghost) particles under the same physical environment. It is well known that, for mono-dispersed particle, gn(r)g_{n}(r) is a power-law of rr [19][8]. However, as shown in Figure 4, when plotted against rdr-d, gng_{n} level off to plateau as rr decrease towards dd as a result of finiteness of gn(d)g_{n}(d). The collision induced modulation factor γc\gamma_{c} which, by definition, equals g(r)g(r) compensated (divided) by gn(r)g_{n}(r), is calculated as such in each case and shown in Figure 5. As expected, at large rr, γc\gamma_{c} universally converge to unity, signifying that collisional effect are only appreciable at rdr\sim d. In the rdr\sim d regime on the other hand, we see that as StSt increases from a minute value (i.e., 0.010.01), γc\gamma_{c} gradually decreases, with smaller rr affected more strongly. The rate of this decrease (with respect to StSt) is at first very weak, consistent with the hypothesis in [14] that γc\gamma_{c} is independent of StSt in the limit of small StSt. The observed rate of decrease becomes pronounced when StSt increases from 0.10.1 to 0.50.5. Beyond St=0.5St=0.5, γc\gamma_{c} seem almost stagnant again until it reverse the trend and start to increases significantly when StSt is greater than 1.01.0. This implies that when the Stokes number of particles is much smaller than 1.0, the influence of collision on the RDF is insensitive to the Stokes number. As the Stokes number increases, the influence of collision grows and peaks at St=1.0St=1.0, which is a trend similar to the one of the power (c1c_{1}) of inertial clustering [19]. In order to see the trend of γc\gamma_{c} at small rr more clearly, using the case of St=0.1St=0.1, 0.2, and 0.5 as examples, we plot in inset of Figure 5, γc\gamma_{c} versus the gap distance rdr-d. Again, we see that γc\gamma_{c} follows a semi-power-law for rdr-d in the range of 16×1041-6\times 10^{-4}. At smaller rdr-d, the curves flatten and level off to a finite value.

Refer to caption
Figure 4: The RDF for non-colliding particles (gng_{n}) for different Stokes numbers. The particle diameter is d=9.49×104d=9.49\times 10^{-4} and Reλ=124Re_{\lambda}=124. In log-log axes, gng_{n} level off to plateau as the gap-distance rdr-d approaches zero.
Refer to caption
Refer to caption
Figure 5: The collisional modulation factor γc\gamma_{c} versus the gap-distance rdr-d for particles with different Stokes numbers. The particle diameter is d=9.49×104d=9.49\times 10^{-4} and Reλ=124Re_{\lambda}=124. Within the range of uncertainty, which is calculated as the standard error, γc\gamma_{c} is weakly dependent on StSt for St1.0St\ll 1.0 and it decreases as StSt increases from around 0.1 to 0.5.

For the sake of comparison, let us recall from earlier works [7][8][19] that gng_{n} is a negative power law of rr in the regime of r/η20r/\eta\lesssim 20, i.e., gn(r)=c0(r/η)c1g_{n}(r)=c_{0}(r/\eta)^{-c_{1}}, where c1c_{1} scales as St2St^{2} for St1St\ll 1. We now attempt to derive an analogous relationship between γc\gamma_{c} and the Stokes number. From Figure 5, γc\gamma_{c} seems to follow a quasi-power-law for gap distances (rd)(r-d) in the range 104rd7×10310^{-4}\lesssim r-d\lesssim 7\times 10^{-3}. We assume that in this region γc=c~0(rd)c~1\gamma_{c}=\tilde{{c}}_{0}(r-d)^{\tilde{c}_{1}}. The relationship between c~1\tilde{c}_{1} and StSt is shown in Figure 6. For St1St\ll 1, considering the level of statistical uncertainty, there is a tentatively trend of increasing c~1\tilde{c}_{1} with Stokes number (c~1\tilde{c}_{1} follows a quasi-linear relationship with StSt for St1St\ll 1 in linear axes). This positive trend is clearer at larger StSt. c~1\tilde{c}_{1} reach a peak value at around St=0.7St=0.7 and thereafter decreases slowly at larger StSt.

For comparison, we also show c1(St)c_{1}(St) and γc(rd)\gamma_{c}(r\to d) in Figure 6, where γc(rd)\gamma_{c}(r\to d) is the limiting value of γc\gamma_{c} at particle contact (r=dr=d). The latter is of interest since it is closely related to the collision rate (in practice, we use the value of γc\gamma_{c} at rd2×106r-d\sim 2\times 10^{-6} as this limit). Comparing these three plots, we see that the general trend of both c~1\tilde{c}_{1} and γc(rd)-\gamma_{c}(r\to d) as a function of Stokes number is similar to that of c1c_{1}.

The above results indicate that when St1.0St\ll 1.0, the drop of RDF caused by particle collision-coagulation is roughly independent of the Stokes number. While for large StSt (>0.2>0.2), the RDF in depletion zone is related to it. Furthermore, the relationship between the decrease of RDF and StSt is similar to that between the preferential concentration of particle and StSt. This implies that the Stokes number dependence of the RDF could not be complete decoupled from γc\gamma_{c} and that this issue merit further investigations.

Refer to caption
Figure 6: The value of exponent of the power-law c~1\tilde{c}_{1} in γc=c~0(rd)c~1\gamma_{c}=\tilde{c}_{0}(r-d)^{\tilde{c}_{1}} in the range 1×104rd7×1041\times 10^{-4}\lesssim r-d\lesssim 7\times 10^{-4}, which is shown as blue \triangle. Its error bars represent the statistical uncertainty in the nonlinear regression method. The value of γc\gamma_{c} at rdr\to d (rd0.02dr-d\sim 0.02d) is shown as red \square and its error bars represent one standard error. Inset) The value of the exponent of the power-law c1c_{1}, as defined in gn=c0(r/eta)c1g_{n}=c_{0}(r/eta)^{-c_{1}}, obtained from DNS results in the range 0.1r/η10.1\lesssim r/\eta\lesssim 1. The statistical uncertainty of c1c_{1} is smaller than the size markers, therefore it is not shown in this figure. The vertical axes and the horizontal axis are logarithmic. c~1\tilde{c}_{1} and γc(rd)\gamma_{c}(r\to d) increase slowly for St1St\ll 1 and it grows significantly from St=0.2St=0.2 to 0.5. The general trend of c~1\tilde{c}_{1} and γc(rd)-\gamma_{c}(r\to d) as a function of StSt is similar to that of c1c_{1}.
{comment}
Refer to caption
Figure 7: The relationship between c~1\tilde{c}_{1} in γc=c~0(rd)c~1\gamma_{c}=\tilde{c}_{0}(r-d)^{\tilde{c}_{1}} with the Stokes number in the range of 3×104rd6×1043\times 10^{-4}\lesssim r-d\lesssim 6\times 10^{-4}.
Refer to caption
Figure 8: The relationship between the value of γc\gamma_{c} at rd1×104r-d\sim 1\times 10^{-4} with the Stokes number.
Refer to caption
Figure 9: The relationship between the value of γc\gamma_{c} at the separation rd=2×106r-d=2\times 10^{-6} and the Stokes number.

III.2.1 Comparative study of the full RDF

In view of the significant Stokes number dependence of γc\gamma_{c}, we also conducted a complementary investigation on the Stokes number trend of the full RDFs (i.e., g(r)g(r)) as a comparative study. The RDFs for various Stokes numbers are translated in vertically to overlap with the g(r)g(r) for St=0.05St=0.05 at rd2×104r-d\sim 2\times 10^{-4} to compare their shape. In order to show the influence on StSt more clearly, the RDFs for St=0.05St=0.05 and 0.7 are shown in Figure 10 and the RDFs for St1St\ll 1 are shown in the inset. It can be seen that the slope of RDF is almost constant for St1St\ll 1 but is changing when considering larger StSt. Furthermore, the position of the peak of the RDF has a significant StSt-dependence. This shows that both γc(r)\gamma_{c}(r) and g(r)g(r) has nontrivial Stokes number dependence which should be further investigated in future works.

Refer to caption
Figure 10: The RDF for St=0.7St=0.7 is translated vertically down to 6.5%6.5\% of its original height to compare the slope with St=0.05St=0.05. The translated RDFs for St=0.01St=0.01, 0.05, and 0.1 are in the inset. The slope is the same for St1St\ll 1 but it is different for St=0.7St=0.7 and 0.05. The position where the peak of RDF is related to StSt.

III.3 Reynolds number dependence

Statistics of Run 3, Run 11, and Run 12, listed in Table 2, are used to investigate the influence of the Taylor-scaled Reynolds number ReλRe_{\lambda} on RDF. The RDFs for different ReλRe_{\lambda} are shown in Figure 11. The Stokes number is 0.1 and the parameters for three simulations are shown in Table 1. We see that in the range of small rr, the RDFs of all cases overlap but for larger rr, the RDFs are separated. However, if rr is normalized by the Kolmogorov length scale η\eta, as shown in the inset of Figure 11, the RDFs now overlap at large rr. These results suggest that in the range of Reλ=84189Re_{\lambda}=84-189, the statistics of the depletion zone related to particle collision is not influenced by ReλRe_{\lambda}, while the power regime related to the classical inertial clustering depends on ReλRe_{\lambda} but only through the change in Kolmogorov length-scale η\eta.

The modulation factor γc\gamma_{c} for these cases are shown in Figure 12. We find that they overlap within statistical uncertainty. This again suggests that the main characteristics of the depletion zone is insensitive to flow Reynolds number from Reλ=84Re_{\lambda}=84 to 189.

According to the Kolmogorov 1941 hypothesis [20], when the Reynolds number is large enough, the statistics of the small-scale of turbulent flow will not be influenced by of the large-scale. Collision occurs on a scale of particle diameter, which is here much smaller than the Kolmogorov length scale, therefore, the decrease of RDF is insensitive to ReλRe_{\lambda}. This view is consistent with the above findings.

Refer to caption
Figure 11: The RDFs for particles in three cases with different ReλRe_{\lambda}. The particle Stokes number is 0.1 and the diameter of particle is d=9.49×104d=9.49\times 10^{-4}. Inset) the RDFs as the function of rr normalized by the Kolmogorov length η\eta. The observed overlaps implies that Renoylds number effect is very weak.
Refer to caption
Figure 12: The modulation factor γc\gamma_{c} versus rdr-d for particles in three cases with different ReλRe_{\lambda}. The Stokes number is 0.1 and the particle diameter d=9.49×104d=9.49\times 10^{-4}. γc\gamma_{c} for three cases are overlapped within a range of uncertainty. Error bars represents standard deviation.

III.4 Particle diameter dependence

The RDFs for particles with different diameters are shown in Figure 13, the statistics used are from Run 3, Run 9, and Run 10. What is striking in Figure 13 is that the position where RDF starts decreasing is consistent with particle diameter.

The modulation factor γc\gamma_{c} for particles with different diameter as a function of the rescaled gap distance (rd)/d(r-d)/d is shown in Figure 14. In this case, the modulation factors γc\gamma_{c} are coincident.

The collision process of particles is strongly related to the particle size. Since we only consider the RDFs of mono-dispersed particles, we expected that the position where depletion zone begins is close to the particle diameter. The results shown here are in line with our expectations.

Refer to caption
Figure 13: The RDFs for particles with different diameters, d=13dd=\frac{1}{3}d_{*}, d=dd=d_{*} and d=3dd=3d_{*} respectively, in which d=9.49×104d_{*}=9.49\times 10^{-4}. The Stokes number of particles is 0.1 and the Reynolds number of the flow is Reλ=124Re_{\lambda}=124. The position where the RDF decreases closely follows to the particle diameter.
Refer to caption
Figure 14: The collisional modulation factor γc\gamma_{c} for particles with different diameters, d=13dd=\frac{1}{3}d_{*}, d=dd=d_{*} and d=3dd=3d_{*}, in which d=9.49×104d_{*}=9.49\times 10^{-4}. The Stokes number of particles is 0.1 and the Reynolds number of turbulent flow is Reλ=124Re_{\lambda}=124. γc\gamma_{c} for all three cases coincide substantially.
{comment}

IV Discussion

When the collision and coagulation occur, the RDF of particle drops significant at the separation rr is close to the particle diameter dd. The RDF is calculated by the probability of finding another particle at a separation distance rr away from the primary particle normalized by the same probability for randomly particle distribution. Because of the collision and coagulation, this probability decreases, which results in the decrease of the RDF. After few eddy turnover time, the preferential concentration and the collision of particle reach a dynamic equilibrium. The RDF will not decrease infinitely with the decreasing of rr. This is consistent with our results in figure x and y. The decreasing part of the RDF, which will be mainly discussed here, is named the depletion zone, in the range of 2×106rd1×1042\times 10^{-6}\lesssim r-d\lesssim 1\times 10^{-4}.

For particles with different Stokes number, the RDFs have the same shape in the depletion zone. And when particles in three different flows with three turbulent Reynolds number, the depletion zone of RDF will not be influenced by the Reynolds number of turbulent flow. According to the Kolmogorov’s hypothesis [20], when the Reynolds number is large enough, the statistics of the small-scale motions will not be influenced by of the large-scale. If the Stokes number of particle is small (St<O(10)St<O(10)), the clustering of particles is mainly affected by the flow in small scale. The former research [9] [collins2004reynolds] also shows that the RDF versus r/ηr/\eta is going to be saturation with increasing Reynolds number. Since the particles diameters in this paper are much smaller than the Kolmogorov length scale η\eta, the scale where particles collide is also much smaller than η\eta, and the Reynolds number will not affect the decline of RDF. When the diameters of the particles are different, the position of the RDF depletion zone is consistent with the particle diameter. From the above results, we can say that the depletion zone of RDF is self-similar.

In this article, we propose the modulation factor γc\gamma_{c} in the RDF: g(r)=γc(r)gn(r)g(r)=\gamma_{c}(r)g_{n}(r) represents the impact of particle collision on the RDF. Unexpectedly, γc\gamma_{c} is sensitive to the Stokes number, and the relationship with the Stokes number is opposite to the relationship with gng_{n}, indicating that the higher the degree of particle preferential concentration, the effect of collision on RDF bigger. In the different range of rdr-d, the functional relationship between γc\gamma_{c} and the Stokes number is also complicated. Furthermore, γc\gamma_{c} is not influenced by the Reynolds number of flows, and its depletion zone is also consistent with the particle diameter. In the study of this paper, the function and significance of the modulation factor remains to be discovered, and whether it can help to study more factors that affect particle collision remains to be studied in the future.

V Conclusion

This paper studies the change of radial distribution function (RDF) of particles subjected to the collision and coagulation (coalescence) interaction. We investigate the relationship between the RDF and particle Stokes number, particle diameter, and Reynolds number. We find that the RDF diminishes significantly at small particle separation distances rr. When viewed as a function of rdr-d, we found evidence that the RDF do not decrease infinitely, but levels off to a fixed value in the limit of rdr\to d. To study the relationship between the degree of reduction of the RDF and the particle and turbulent parameters, we separate the RDF into two multiplicative parts i.e., g=γcgng=\gamma_{c}\,g_{n}, where gng_{n} is the RDF for non-colliding particles under equivalent environment and γc\gamma_{c} is a collisional modulation factor that reflects the effect of particle collision on particle preferential concentration. We see that gn(rd)g_{n}(r-d) levels off to a plateau as the argument rdr-d approaches zero. On the other hand, the collisional factor γc\gamma_{c} universally converges to unity at large rr and levels off to a fixed value at rdr\approx d. We find that γc\gamma_{c} is dependent on the Stokes number. Specifically, assuming a power-law model for γc\gamma_{c} in the region 0.1drd0.7d0.1d\lesssim r-d\lesssim 0.7d (i.e., γc=c~0(rd)c~1\gamma_{c}=\tilde{c}_{0}(r-d)^{\tilde{c}_{1}}), we find that in the small Stokes number limit, the value of c~1\tilde{c}_{1} is only very weakly dependent on StSt, while the overall Stokes number trend of c~1\tilde{c}_{1} is qualitatively similar to the power law exponent c1c_{1} in the RDF of non-colliding (ghost) particles (i.e., gn(r))g_{n}(r)). The magnitude of γc\gamma_{c} at the limit rdr\to d varies with the Stokes number following a trend similar to that of c~1(St)\tilde{c}_{1}(St).

The preceding findings motivate a comparative investigation into the StSt trend of the full RDF, which has the result that the slope of the RDF g(r)g(r) in the depletion zone is the same for St1St\ll 1 but it is different for large Stokes number (i.e., St0.05St\geq 0.05). Besides this, the location of the peak of RDF is found to be significantly StSt-dependent. These findings imply that the Stokes number dependence of the RDF could not be completely decoupled from γc\gamma_{c} (nor from g(r)g(r)) except in the regime of St1St\ll 1 where StSt dependence is weak.

We find that the shape of the RDF in the depletion zone (rdr\sim d) do not change with the variation of the flow Reynolds number within the range studied i.e., Reλ=84189Re_{\lambda}=84-189 and the collisional modulation factor γc\gamma_{c} from different ReλRe_{\lambda} overlap.

On the effect of particle diameter dd, we find that larger dd leads to the falling edge in γc\gamma_{c} occurring at larger values of rdr-d (and rr) such that the results coincide when γc\gamma_{c} is plotted against (rd)/d(r-d)/d.

Acknowledgements.

This work was supported by the National Natural Science Foundation of China (grant 11872382) and by the Thousand Young Talents Program of China.

References

  • Shaw [2003] R. A. Shaw, Particle-turbulence interactions in atmospheric clouds, Annual Review of Fluid Mechanics 35, 183 (2003).
  • Grabowski and Wang [2013] W. W. Grabowski and L.-P. Wang, Growth of cloud droplets in a turbulent environment, Annual review of fluid mechanics 45, 293 (2013).
  • Cuzzi et al. [2001] J. N. Cuzzi, R. C. Hogan, J. M. Paque, and A. R. Dobrovolskis, Size-selective concentration of chondrules and other small particles in protoplanetary nebula turbulence, The Astrophysical Journal 546, 496 (2001).
  • Johansen et al. [2007] A. Johansen, J. S. Oishi, M.-M. M. Low, H. Klahr, T. Henning, and A. Youdin, Rapid planetesimal formation in turbulent circumstellar disks, Nature 448, 1022 (2007).
  • Pan et al. [2011] L. Pan, P. Padoan, J. Scalo, A. G. Kritsuk, and M. L. Norman, Turbulent clustering of protoplanetary dust and planetesimal formation, The Astrophysical Journal 740, 6 (2011).
  • Squires and Eaton [1991] K. D. Squires and J. K. Eaton, Preferential concentration of particles by turbulence, Physics of Fluids A: Fluid Dynamics 3, 1169 (1991).
  • Reade and Collins [2000] W. C. Reade and L. R. Collins, Effect of preferential concentration on turbulent collision rates, Physics of Fluids 12, 2530 (2000).
  • Chun et al. [2005] J. Chun, D. L. Koch, S. L. Rani, A. Ahluwalia, and L. R. Collins, Clustering of aerosol particles in isotropic turbulence, Journal of Fluid Mechanics 536, 219 (2005).
  • Ireland et al. [2016] P. J. Ireland, A. D. Bragg, and L. R. Collins, The effect of reynolds number on inertial particle dynamics in isotropic turbulence. part 1. simulations without gravitational effects, Journal of Fluid Mechanics 796, 617 (2016).
  • Saw et al. [2008] E. W. Saw, R. A. Shaw, S. Ayyalasomayajula, P. Y. Chuang, and A. Gylfason, Inertial clustering of particles in high-reynolds-number turbulence, Physical review letters 100, 214501 (2008).
  • Salazar et al. [2008] J. P. Salazar, J. De Jong, L. Cao, S. H. Woodward, H. Meng, and L. R. Collins, Experimental and numerical investigation of inertial particle clustering in isotropic turbulence, Journal of Fluid Mechanics 600, 245 (2008).
  • Sundaram and Collins [1997] S. Sundaram and L. R. Collins, Collision statistics in an isotropic particle-laden turbulent suspension. part 1. direct numerical simulations, Journal of Fluid Mechanics 335, 75 (1997).
  • Wang et al. [2000] L.-P. Wang, A. S. Wexler, and Y. Zhou, Statistical mechanical description and modelling of turbulent collision of inertial particles, Journal of Fluid Mechanics 415, 117 (2000).
  • Saw and Meng [2022] E.-W. Saw and X. Meng, Intricate relations among particle collision, relative motion and clustering in turbulent clouds: computational observation and theory, Atmospheric Chemistry and Physics 22, 3779 (2022).
  • Rogallo [1981] R. S. Rogallo, Numerical experiments in homogeneous turbulence, Vol. 81315 (National Aeronautics and Space Administration, 1981).
  • Eswaran and Pope [1988] V. Eswaran and S. B. Pope, An examination of forcing in direct numerical simulations of turbulence, Computers & Fluids 16, 257 (1988).
  • Maxey and Riley [1983] M. R. Maxey and J. J. Riley, Equation of motion for a small rigid sphere in a nonuniform flow, The Physics of Fluids 26, 883 (1983).
  • Ireland et al. [2013] P. J. Ireland, T. Vaithianathan, P. S. Sukheswalla, B. Ray, and L. R. Collins, Highly parallel particle-laden flow solver for turbulence research, Computers & Fluids 76, 170 (2013).
  • Saw et al. [2012] E.-W. Saw, J. P. Salazar, L. R. Collins, and R. A. Shaw, Spatial clustering of polydisperse inertial particles in turbulence: I. comparing simulation with theory, New Journal of Physics 14, 105030 (2012).
  • Kolmogorov [1941] A. N. Kolmogorov, The local structure of turbulence in incompressible viscous fluid for very large reynolds numbers, Cr Acad. Sci. URSS 30, 301 (1941).