Вы находитесь на странице: 1из 11

Acta Mech Sin (2015) 31(3):292302

DOI 10.1007/s10409-015-0398-5

REVIEW PAPER

A review of research on nanoparticulate flow undergoing


coagulation
Jianzhong Lin1,2 Linlin Huo1

Received: 12 September 2014 / Revised: 20 October 2014 / Accepted: 29 October 2014 / Published online: 15 May 2015
The Chinese Society of Theoretical and Applied Mechanics; Institute of Mechanics, Chinese Academy of Sciences and Springer-Verlag Berlin
Heidelberg 2015

Abstract Nanoparticulate flows occur in a wide range of


natural phenomena and engineering applications and, hence,
have attracted much attention. The purpose of the present
paper is to provide a review of the research conducted over
the last decade. The research covered relates to the Brownian
coagulation of monodisperse and polydisperse particles, the
Taylor-series expansion method of moment, and nanoparticle
distributions due to coagulation in pipe and channel flow,
jet flow, and the mixing layer and in the process of flame
synthesis and deposition.
Keywords

Nanoparticulate flow Coagulation Review

1 Introduction
Nanoparticulate flows occur in a wide range of natural
phenomena and engineering applications and have been
extensively investigated during the past 20 years. Many physical properties of nanoparticulate flows and their behavior
involving diffusion and condensation are strongly dependent
on their size distribution. One of the most significant factors
affecting particle size distribution is coagulation. Individual
nanoparticles suspended in a fluid may come into contact,
collide with one another, and stick together to form larger
particles because of their Brownian motion or as a result of
their relative motion arising from gravity, electrical force,

Jianzhong Lin
jzlin@sfp.zju.edu.cn

Institute of Fluid Measurement and Simulation,China Jiliang


University, Hangzhou 310018, China

Institute of Fluid Engineering, Zhejiang University,


Hangzhou 310027, China

123

hydrodynamic force, or other external forces. The net result


of particle coagulation is a decrease in particle number and
an increase in particle size. Therefore, in actual applications,
the evolution of particle number concentration and size distribution in nanoparticulate flow undergoing coagulation is
of fundamental importance and interest.

2 Brownian coagulation of monodisperse particles


2.1 Equation for particle number
Based on the assumption that particles will adhere when they
collide and that particle size changes slowly, Uchowski [1]
first derived the coagulation equation for monodisperse particles by solving the diffusion equation around a particle and
by obtaining the flux of other particles toward it:
dn
= k0 n 2 ,
dt

(1)

where n is the particle number, t the time, k0 the ideal rate


without considering the interactions between particles, and
is the collision efficiency and represents the probability
of coagulation between two particles [2]. By integrating the
coagulation equation, the particle number concentration can
be explicitly determined as a function of time.
2.2 Particle collision efficiency
The collision efficiency in Eq. (1) is an important physical
quantity. Some studies in the literature address the collision
efficiency. By considering the van der Waals force, Russel
et al. [2] computed the coagulation rate. The collision effi-

A review of research on nanoparticulate flow undergoing coagulation

ciency was first derived by Han et al. [3] by considering the


interparticle forces and hydrodynamics and then combining
the shift of the centroids of the curves for the particle size
distribution with the total number of particles to quantify the
coagulation rate [4]. Shear-induced coagulation under the
effect of van der Waals interaction was studied by Vanni and
Baldi [5]. An expression of the collision efficiency for various
collision angles was derived by Chin et al. [6] by considering
the electrostatic, van der Waals, magnetic dipole, and interparticle forces and hydrodynamics. A geometric model was
proposed by Olsen et al. [7] to predict the collision efficiency
for systems consisting of two oppositely charged species. The
effects of van der Waals forces and noncontinuum lubrication
forces on the coagulation process were explored by Chun and
Koch [8]. An expression of the collision efficiency considering elastic deformation and van der Waals forces was derived
by Feng and Lin [9]. New formulas for the central oblique
collision efficiency of dioctyl phthalate particles with diameters ranging from 100 to 800 nm were derived by Wang
et al. [10] by considering the elastic deformation and van der
Waals force. Those researchers showed that the elastic deformation between two particles should not be neglected. There
exists an obvious difference in the collision efficiencies for
central normal and central oblique collisions. New formulas
of the collision efficiency of dioctyl phthalate particles with
diameters ranging from 50 to 500 nm under the effect of the
Stokes resistance, van der Waals, elastic deformation, and
lubrication forces were built by Chen and You [11]. They
found that the collision efficiency increased with decreasing
particle size.
The collision efficiency in cases where a nanoparticle
collides with a wall was studied by Wang and Lin [12]
by considering the Stokes resistance, van der Waals, and
elastic deformation forces. New formulas for the collision
efficiency of dioctyl phthalate particles with diameters ranging from 100 to 800 nm were derived for different initial
angles of attack. The results showed that, on the whole,
collision efficiency increases with decreasing particle size,
but there is a maxmum collision efficiency for particles
with diameters of around 550 nm. Whether or not the elastic deformation force is considered has a significant efffect
on the collision efficiency, whereas the difference in the
collision efficiency in cases where particles collide with a
wall vertically or horizontally can be neglected. Molecular dynamics simulation was used by Zhang et al. [13] to
explore the effect of interactions between two dipoles on the
coagulation of charge-neutral TiO2 nanoparticles in the free
molecule regime, and the results showed that particle coagulation increases with decreasing particle diameter because
the dipoledipole force gradually becomes larger than the
van der Waals force. However, the effect of the interaction between two dipoles on coagulation drops drastically
with as temperature increases because the average dipole

293

moment becomes correspondingly small. A model of central oblique collisions between two nanoparticles at different
initial angles of attack was proposed by Wang and Lin [14]
to derive the collision efficiency under the effect of elastic
deformation and van der Waals forces. New formulas for
the friction coefficient between two nanoparticles were first
proposed by Chen et al. [15] using a statistical method and
then modifying the relative diffusion coefficient, taking into
account the slip effect. The results showed that the modified diffusion coefficient enhanced the collision efficiency
to some extent whether or not the van der Waals force was
considered.

3 Brownian coagulation of polydisperse particles


The coagulation equation for polydisperse particles becomes
much more complicated and no explicit solution exists
because the coagulation rate depends on the range of particle sizes. The coagulation mechanisms of unequal sized
silicon nanoparticles of volume ratios between 0.053 and 1
were explored by Hawa and Zachariah [16] using molecular dynamics simulation. It was found that the convection
process and the diffusion process dominate the coagulation
process for liquidlike particles and for near solidlike particles, respectively. For liquidlike particles, the deformation
of smaller particles also has a significant effect on the coagulation process. Two particles with a much smaller ratio of
sizes coagulate much faster. The collision efficiency of two
unequal sized dioctyl phthalate particles with diameters ranging from 100 to 750 nm was derived by Wang and Lin [17]
by considering the elastic deformation and van der Waals
forces. They found that, on the whole, the collision efficiency increases as particle sizes decrease. A coagulation
rate constant that is dependent on the time and particle size
was derived by Kelkar et al. [18]. The results showed that the
size dependence of the coagulation rate constant affects in
particular predictions for initially polydisperse particle systems.
To predict the coagulation of polydisperse particles more
generally, the following population balance equation was
proposed by Mller [19] to represent changes in the polydisperse particle size distribution:
1
n (v, t)
=
t
2

v
(v1 , v v1 ) n (v1 , t) n (v v1 , t) dv1
0


n (v, t)

(v1 , v) n (v1 , t) dv1 ,

(2)

where n(v, t)dv is the number of particles whose volume is


between v and v + dv, and (v1 , v) is the collision kernel

123

294

J. Lin, L. Huo

for two particles of volume v and v1 , which depends on the


collision mechanism and the sizes of two colliding particles.
The first and second terms on the right-hand side denote the
increase and decrease in particle number, respectively.
As far as the Brownian coagulation is concerned, for those
particles much smaller or much larger than the mean free path
length of the gas molecules, the kinetic theory of gases or the
continuum diffusion theory must be used to derive the collision kernel, respectively. Therefore, the Knudsen number
K n = l/r (l is the mean free path length of the gas molecules
and r is the particle radius) is usually used to define the particle size regime. In the free molecule regime with K n > 50,
the collision kernel is [20]
1/3

fm (v, v1 ) = B1 (1/v + 1/v1 )1/2 (v 1/3 + v1 )2 ,

(3)

where B1 = (3/4 )1/6 (6kb T /)1/2 (kb is the Boltzmann


constant, T is the temperature, and is the particle density)
is the Brownian coagulation coefficient.
In the continuum regime with K n < 1, the collision kernel
is derived by the continuum diffusion theory as [20]

co (v, v1 ) = B2



C(v) C (v1 )  1/3
1/3
v
,
+
+
v
1
1/3
v 1/3
v1

(4)

where B2 = 2kb T / ( is the viscosity), and C = 1 +


K n[1.142+0.558exp(0.999/Kn)] is the gas slip correction
factor [21].
In the transition regime with 1 < K n < 50, the collision
kernel is described neither by the kinetic theory of gases nor
by the continuum diffusion theory. Based on the semiempirical solution of the collision kernel found by Fuchs [22],
Otto et al. [23] proposed the collision kernel in the transition
regime as
entire (v1 , v) = co f (K n),

(5)

where f (Kn) is called the enhancement function.

4 Taylor-series expansion method of moment


The evolution of nanoparticle distribution can be traced by
solving the population balance equation (2). However, the
numerical calculations often become impractical because of
the huge computational cost. Some methods have been used
to alleviate the computational cost; one of the common methods is called the method of moment [2426], which has been
extensively used and has become a powerful tool owing to
the relative simplicity of its implementation and relatively
low computational cost. In the method of moment, Eq. (2) is
first transformed into an ordinary differential equation with
respect to the moment m k by multiplying both sides of Eq. (2)

123

by v k and then integrating over the entire particle size range


[27]:
1
m k
=
t
2

 

(v + v1 )k v k
0
0

v1k (v, v1 ) n (v, t) n (v1 , t) dvdv1 ,
(k = 0, 1, 2, . . .),

(6)

with


mk =

v k n(v)dv.

(7)

The moment equation (6) must be closed before being solved.


A number of different methods for closing Eq. (6) have
emerged. The method of moment proposed by Pratsinis
(PMM) [28] makes a prior assumption about the profile of
the particle size distribution. The PMM can be used to obtain
such important physical variables as total particle concentration, average particle size, and polydispersity by solving the
zero-, one-, and two-order moment equations. The quadrature
method of moment proposed by McGraw [29] approximates
the integral moment by an n-point Gaussian quadrature and
can be applied to cases governed by laws of particle growth
and collision kernels. With this method, the moment equation can be closed by the quadrature approximation without
making a prior assumption for the profile of the particle size
distribution when the population balance equation is derived
in terms of one internal coordinate.
The Taylor-series expansion method of moment (TEMOM)
proposed by Yu et al. [30] uses the Taylor-series expansion
technique to close the moment equation. The collision kernel
in the free molecule regime is shown by Eq. (3), in which we
use the notation C = (1/v + 1/v1 )1/2 . Equation (6) is not
closed when we substitute Eq. (3) into Eq. (6) because the
variable C exists in quadratic form. For the PMM, Eq. (6)
becomes integrable by using the following approximation
[28,31]:


1/2
,
C = b 1/v 1/2 + 1/v1

(8)

in which b is strongly dependent on the range of the particle size spectrum. For the TEMOM, Eq. (3) of the collision
kernel can be rewritten as
fm = B1 (v + v1 )1/2


1/2
1/6
1/6
v 1/6 v1
,
+ 2v 1/6 v1
+ v 1/2 v1

(9)

and then (v + v1 )1/2 in Eq. (9) is expanded in a Taylor series


about point (v = u, v1 = u):

A review of research on nanoparticulate flow undergoing coagulation

(v + v1 )

1/2

2 (v u)
2 (v1 u)
2u +
+

4 u
4 u

2 (v u)2
2 (v1 u) (v u)

3/2
32u
16u 3/2

2
2 (v1 u)

+ .
(10)
32u 3/2

Substituting Eqs. (9) and (10) into Eq. (6), we can close the
moment equation (6).
The aforementioned method is also used to deal with collision kernels in the continuum regime and close the moment
equation (6).
No prior assumption about the profile of the particle size
distribution is necessary for TEMOM, and the number of
moment equations required equals the order of the Taylorseries expansion. In TEMOM the precision of a solution is
enhanced and the computational cost is increased as the order
of the Taylor-series expansion increases.
The TEMOM has been used to deal with various problems in the past 5 years. The binary homogeneous nucleation
of watersulfuric acid and the growth of nanoparticles were
explored by Yu and Lin [32]. They analyzed the competition between nucleation, coagulation, and condensation in
both cases with and without background particles and showed
that the production rate of sulfuric acid is an important factor affecting nucleation kinetics and particle dynamics. The
effect of particle diffusion on coagulation was studied by
Wang et al. [33], who showed the distributions of particle number and mass concentration and the average particle
volume under coagulation and diffusion. The TEMOM was
modified by Lin and Chen [34] to match the property of real
self-preserved nanoparticles under coagulation based on the
numerical results given by the sectional method in both free
molecule and continuum regimes. The results showed that
when the particle size distribution attains a self-preserved
state or the coagulation time is long enough, the modified
TEMOM can give more precise predictions on the results of
the zeroth and second moments than the original TEMOM.
They also proposed a special kind of coordinate diagram
to describe qualitatively the errors produced by different
methods of moment based on the zeroth, first and second
moments. Meanwhile, a new set of moment equations in the
free molecule regime were built for the case in which the
particle size distribution attained a self-preserved state with
log-normal form. Some fundamental issues of the TEMOM
were investigated by Xie and He [35], who clarified the
uniqueness of Taylor-series expansions, the availability of
closuring moment equations with fractional moments, the
convergence of the analytical solutions, and so on. A direct
expansion method of moment over the entire size regime
was proposed by Chen et al. [36] using the exact Dahnekes
formula as the collision kernel. They showed that the direct

295

expansion method is can be used to describe the evolution


of the zeroth and second moments and predict more precise results than those predicted by the quadrature method
of moment when the initial geometric standard deviation is
relatively small.
In Sects. 58 the nanoparticle distributions due to coagulation in the pipe and channel flow, jet flow, and mixing
layer and in the process of flame synthesis and deposition
are demonstrated based on the Brownian coagulation of
monodisperse and polydisperse particles partly through the
TEMOM.

5 Pipe and channel flows


5.1 Straight pipe and channel flow
The formation, growth, and transport of particles with diameters ranging from 1 to 50 nm in a channel capacitively
coupled radio-frequency silane discharge were investigated
by De Bleecker et al. [37], who found that the concerted
actions of particle charging and transport greatly affect the
location of particle growth because of coagulation. Particle
contamination can be controlled by thermophoretic force.
The distributions of nanoparticle number, average size, and
volume at different pipe lengths and Reynolds numbers were
measured by Yin and Lou [38]. The results show that for
particles with diameters ranging from 5.6 to 560 nm, the
increment of the pipe length and the turbulence intensity will
accelerate the deposition process. The smaller the particles,
the lower the penetration efficiency. Nanoparticle migration
in a turbulent pipe flow was studied by Lin et al. [39]. It
was found that particles distribute nonuniformly in a cross
section. Coagulation leads to a gathering of larger particles
at the pipe center. As the Schmidt number increases, the
particle mass and number concentrations, as well as the polydispersity, decrease in the pipe center. The increase in the
Damkohler number leads to an increase in the particle size
and geometric standard deviation. The transport and penetration efficiency of particles with diameters ranging from 50
to 450 nm in a turbulent flow were studied by Lin et al. [40]
by considering the Brownian diffusion, turbulent diffusion,
coagulation, and breakage. They showed that the particle
number concentration is distributed nonuniformly in cross
sections. with higher values in the pipe center. The particle
diameter decreases gradually from wall to pipe center. The
particle penetration efficiency changes from 65 % to 95 %
when the Reynolds number changes from 4426 to 8500 and
the ratio of pipe length to diameter changes from 375 to 625.
The particle penetration efficiency increases and decreases
with increases in the particle size and the Reynolds number, respectively. The researchers also derived the relation of
particle penetration efficiency to related parameters.

123

296

5.2 Curved pipe and channel flow


Nanoparticle transport and deposition in a curved pipe at different Reynolds numbers and Dean numbers were simulated
by Lin and Lin [41]. It was found that the particle distribution was symmetrical with respect to the pipe bottom and
top sides. There was a large and a small amount of deposition on the outside and inside walls, respectively. Particles
at higher flow Reynolds numbers deposited quickly on the
wall, whereas pipe curvature had an insignificant effect on
particle deposition. The researchers also studied the effect
of Schmidt numbers on particle transport and deposition and
found that the Schmidt number, pipe curvature, and Reynolds
number had first-order, second-order, and forth-order effects
on particle deposition, respectively [42]. Nanoparticle transport and coagulation in a curved pipe were simulated by
Lin et al. [43], who showed that the particle number and
mass concentration, polydispersity, average diameter, and
geometric standard deviation increased with time. The particle number concentration increased, whereas the particle
polydispersity, average diameter, and geometric standard
deviation decreased with increasing initial particle diameters
and Reynolds numbers. The effect of particle coagulation on
particle distribution was more pronounced in the initial stage
than in subsequent stages.
5.3 Rotating curved pipe and channel flow
Nanoparticle transport and deposition in a rotating curved
pipe at different angular velocities, Schmidt numbers, and
Dean numbers were explored by Lin et al. [44]. It was found
that the particle distribution was basically dependent on the
axial velocity when the Schmidt number was small, whereas
the secondary flow controlled the particle distribution when
the Schmidt number was much larger than 1, as shown in
Fig. 1. The particle deposition became uniform along the
entire edge of the wall with an increase in the Dean number.
The rotation direction of the pipe had an important effect on
the particle deposition. There was a large amount of deposition when secondary flow appeared in the pipe.

Fig. 1 Distribution of nanoparticle mass fraction at different Schmidt


numbers

123

J. Lin, L. Huo

The evolution of nanoparticle distribution in a rotating


curved pipe at different rotation numbers, Schmidt numbers,
and Reynolds numbers was explored by Lin et al. [45]. The
results showed that the particles gathered in the region close
to the outside edge of the pipe when the Coriolis force and
the centrifugal force were in the same direction, while particles gathered in the region close to the inside edge when the
Coriolis force was in the opposite direction to the centrifugal
force and the Coriolis force was much larger than the centrifugal force. Particle mass and number concentrations increased
quickly in the early stage than that in subsequent stages and
finally attained a stable value. The particle distribution was
dominated by the competition between the pipe curvature and
the rotation number. The particle mass and number concentrations increased, whereas particle polydispersity, diameter,
and geometric standard deviation decreased with increases
in the Reynolds and Schmidt numbers.

6 Jet flows
6.1 Round jet
Nanoparticle coagulation and dispersion in a round jet were
simulated by Chan et al. [46], who showed that particle mass
concentration in the jet core decreases with increases in the
distance from the jet exit. The particle number concentration decreases quickly in the exit region of the jet and then
decreases slowly as the distance from the jet exit increases,
until an asymptotic state is reached. The growth rate of particle polydispersity is greatest in the exit region of the jet, while
particles with the largest diameter are found in the core region
of the jet. The particle diameter changes slightly across the
width of the jet, except in the interface region between the
jet and the outside. The particle geometric standard deviation increases in the exit region of the jet and then slowly
approaches an asymptotic value in the core region of the jet,
as shown in Fig. 2. Across the width of the jet the particle
diameter increases as the the Damkohler number increases,
as shown in Fig. 3, and the particle number concentration
increases as the Schmidt number increases.
Nanoparticle coagulation and dispersion was studied by
Lin et al. [47] on the basis of unimodal lognormal particle size
distributions. It was found that coherent structures appeared
in the interface region between the jet and the outside, which
promoted particle coagulation and changed the particle concentration distribution across the width of the jet, resulting
in a nonuniform dispersion along the flow direction. With
increases in the Schmidt number, the region occupied by particles became narrower, the particle polydispersity decreased,
and particle number concentration increased. Across the
width of the jet, the particle size and geometric standard
deviation increased as the Damkohler number increased.

A review of research on nanoparticulate flow undergoing coagulation

2.0
x/D =0;
x/D =6;

1.8

x/D =2;
x/D =8

x/D =4;

1.6
1.4
1.2
1.0
-1.6

-1.2

-0.8

-0.4

0.0

0.4

0.8

1.2

1.6

r/D
Fig. 2 Distribution of geometric standard deviation across width of jet

2.0

dp /(nm)

1.8

297

and the smaller the number of particles distributed in the


region close to the impinging plane. The optimum dilution
conditions were explored by Fujitani et al. [50] by considering particle growth via coagulation and other factors in a
diesel exhaust. It was found that a short residence time could
prevent particles from growing via coagulation following
the primary dilution or before the diluted exhaust reached
the inhalation chamber following the secondary dilution. The
effects of some related factors on the gas-to-nanoparticle conversion and distribution of gas released from the tailpipe of a
diesel vehicle were explored by Chen et al. [51]. The results
showed that the coagulation process was slowest compared
with nucleation and turbulent dispersion processes. A higher
vehicle velocity decreased the particle number concentration.
Nanoparticle coagulation and dispersion in a turbulent round
jet were studied by Zhu et al. [52]. It was found that the particle size was distributed uniformly and remained constant
in the potential core but became large in the region of high
turbulence intensities. The distribution of particle size along
the flow direction changed slightly at first, and then quickly,
finally attaining a steady state.

1.6
1.4

6.2 Planar jet


Da =0.5
Da =1
Da =2

1.2
1.0
-1.6

-1.2

-0.8

-0.4

0.0

0.4

0.8

1.2

1.6

r/D
Fig. 3 Distribution of particle diameter across width of jet

The formation of nanoparticles arising from the homogeneous binary nucleation of H2 SO4 and H2 O vapors in twin
round jets was simulated by Yin et al. [48], who showed that
the velocity ratio of ambient wind to exhaust gas plays an
important role in affecting particle concentration and size
distribution. The particle formation rate decreased and particle size increased with increases in the ambient wind velocity.
The formation of nanoparticles in twin impinging jets in different spaces between the two jets at different distances from
nozzle exit to impingement plane was studied by Yin and
Lin [49]. It was found that the maximum particle size and
the maximum particle number concentration appeared in the
region of the free jet and in the region close to the impinging plane, respectively. The particle concentration and size
distribution in the interface region of the two jets were dependent on the space between the two jets. The larger the space
between the two jets, the more the particles were produced.
The greater the distance from the nozzle exit to the impingement plane, the lower the particle number concentration,

Nanoparticle coagulation and growth in a planar jet was studied by Yu et al. [53], who showed that the distributions
of particle diameter, number intensity, and polydispersity
are dominated by the formation and evolution of coherent
structures. Along the flow direction, the particle number concentration decreases in the core region of the jet and increases
in the peripheral region. The particle mass concentration
remains unchanged on the whole flow, whereas particle size
and geometric standard deviation increase and attain their
maximum in the interface region between the jet and the
outside. The effect of Damkohler number and Schmidt number on nanoparticle distribution was investigated by Yu et al.
[54]. It was found that the larger the Schmidt number is, the
narrower the region across the width of the jet in which particles are distributed. Particles with smaller diameters show
a higher polydispersity because they coagulate and disperse
more easily and grow quickly. The characteristic time of particle coagulation is so short that particle collision and coagulation happen frequently, which results in an increase in particle size. Particle polydispersity is directly proportional to the
Damkohler number. The distributions of nanoparticles arising from the homogeneous binary nucleation of H2 SO4 and
H2 O vapors in a submerged constraint jet was simulated by
Liu [55]. It was found that the particle number concentration
clearly increased within the core region of the coherent vortex produced by the rolling up of the interface between jet and
ambient. The presence of a coherent vortex promoted particle
coagulation by increasing the possibility of particle collision.

123

298

The effects of turbulence on nanoparticle growth was


explored by Das and Garrick [56] by considering Brownian
coagulation, nucleation, and condensation in a reacting flow.
They showed that particles with large diameters appeared in
regions away from the core region of the jet. The particle
average diameter increased with increases in the precursor
concentration. The particle growth rate was higher within
the core region of the eddy and increased continuously along
the flow direction. The evolution of the TiO2 nanoparticle
distribution in a turbulent reacting planar jet was simulated
by Loeffler et al. [57] by neglecting the effect of unresolved
fluctuations on coagulation. They suggested that neglect of
the interactions between particles would result in an increase
in the particle growth rate as the precursor concentration
increased. The TiO2 nanoparticle synthesis in a planar, nonpremixed diffusion flame was simulated by Garrick et al. [58]
by considering coagulation, nucleation, coalescence, and
condensation. They explored the effect of mixing and finiterate sintering on particle distributions and showed that highly
agglomerated particles were located on the periphery of the
coherent eddy, where particle collisions resulting in coagulation took place more quickly than particle coalescence.

7 Mixing layers
Based on the assumption that particle size has a lognormal distribution, nanoparticle coagulation in an isothermal
mixing layer was studied by Settumba and Garrick [59],
who discussed the spatiotemporal evolution of the particle
number and mass concentration, mean size, and geometric
standard deviation for Damkohler numbers 0.2, 1, and 2.
The researchers also performed a numerical simulation of
nanoparticle coagulation in a mixing layer [60] and found that
the polydispersity calculated based on the diffusion coefficient and the local average volume was slightly higher; hence,
the method of diffusion coefficient and local average volume
can be used to reduce the spatial resolution, which allows for
more affordable computations. Nanoparticle distribution in
an isothermal shear layer under the influences of coagulation,
convection, and diffusion at a Reynolds number of 200 and
Damkohler numbers of 1 and 10 was investigated by Garrick
et al. [61], who showed that the nonuniformity of the particle
concentration distribution grew with time, which resulted in
an increase in the particle geometric standard deviation. With
the increase in the Damkohler number, the particle growth
increased and the particle size distributions became wider
than the self-preserving limit. The formation and growth of
TiO2 nanoparticles, up to and including particles 128 nm in
diameter, in a mixing layer was studied by Wang and Garrick [62]. It was found that particle formation and growth
were limited by the mixing and particle size increased more
quickly with increases in the initial reactant levels. The

123

J. Lin, L. Huo

Fig. 4 Distribution of particle mean diameter

Fig. 5 Distribution of volume concentration

particle geometric standard deviation showed a greater variation throughout the mixing layer. A numerical simulation of
nanoparticles arising from a homogeneous binary nucleation
of H2 SO4 and H2 O vapors in a mixing layer was performed
by Lin and Liu [63]. It was found that the particle number
and volume concentration distributions, as well as the mean
diameter, were dominated by the coherent vortex structure.
The particle number concentration decreased, whereas the
volume concentration and mean diameter increased along
the flow direction, as shown in Figs. 4 and 5. The particle
number and volume concentration, as well as mean diameter, were distributed nonuniformly across the width of the
mixing layer. The coagulation process was longer than the
nucleation process and was mainly affected by the number
concentration.
Particle coagulation in a temporal mixing layer was studied by Xie et al. [64]. The results showed that the particle
number concentration decreased while the particle mean volume increased with time. The particle mass and number concentration and mean volume were distributed nonuniformly
owing to the existence of a coherent structure. The flow
advection had an insignificant effect on particle coagulation
in the region far from the coherent structures. The particles
had an obvious wavelike distribution within the core region
of the coherent vortex. The formation and evolution of the
coherent structure had a greater effect on particle coagulation,
which in turn affected the distributions of the particle mass
and number concentration, as well as the mean diameter.

8 Flame synthesis and deposition


The generalizations of a Gaussian quadrature-based ninemoment method was used by Rosner and Pyykonen [65]

A review of research on nanoparticulate flow undergoing coagulation

Inlet TiCl4 loading (mol/min)


0.4 10-4

m0

1021
10

20

10

19

10

18

10

17

10

1.6 10-4
5.8 10-4

16

0.05

0.1

x
Fig. 6 Distribution of particle number concentration
Inlet TiCl4 loading (mol/min)
0.4 10-4

500

1.6 10-4

400

5.8 10-4

da

to study the bivariate particle balance equation in a laminar flame reactor. The evolution of alumina nanoparticles
was predicted under the combined effects of coagulation,
thermophoresis, and sintering. The formation and growth
of polydisperse nonspherical silica nanoparticles in an oxyhydrogen coflow diffusion flame was simulated Kim et al.
[66] by considering particle generation, coagulation, diffusion, convection, and coalescence. The results showed that
the distribution of nonspherical particle sizes was highly
spatially nonuniform. The attainment of size distributions
narrower than those predicted by the self-preserving theory of nanoparticle coagulation was studied by Tsantilis
and Pratsinis [67]. They simulated the TiO2 formation via
titanium-tetra-isopropoxide (TTIP) or TiCl4 oxidation by
considering the gas/surface reactions and coagulation and
illustrated the effects of temperature, pressure, and initial
precursor molar fraction on the distributions of TiO2 particle
size and geometric standard deviation. A new constitutive law
for fractal aggregates resulting from coagulation was generalized by Kostoglou et al. [68], who added a restructuring
mechanism to the population balance model. By solving the
bivariate coagulation equation, they found that the presence
of restructure resulted in an evolution dynamics of the fractal aggregate distribution that is much richer than that given
by previous coagulation models. A mass-flow-type stochastic particle algorithm that is used to simulate nanoparticle
growth in flames and reactors was derived and tested by
Morgan et al. [69]. The newly derived algorithm combines
the effect of coagulation with particle source and surface
growth. A particle model and stochastic methods to simulate
nanoparticle size distributions in a premixed flame was used
by Morgan et al. [70]. The results provided further evidence
of the interplay among nucleation, coagulation, and surface
rates. The formation of monomodal polyelectrolyte complex
nanoparticles under different ionic strengths and applied centrifugation regimes was studied by Starchenko et al. [71],
who used a novel particle coagulation model to simulate the
aggregation process under different typical colloidal parameters and analyzed the effects of these parameters on the mean
final size and the size distribution function of the particles.
A new numerical algorithm was developed by Yu et al. [72]
to explore the effects of precursor loading on TiO2 nanoparticle synthesis in a flame reactor. The precursor TiCl4 oxidation
leading to particle formation is modeled using a one-step
chemical kinetics approach. It was found that the particle
number concentration and diameter increased with increases
in precursor loading, as shown in Figs. 5 and 6. The particle
surface fractal dimension was weakly dependent on the inlet
precursor loading. When the inlet precursor loading remained
unchanged, the larger the carrying gas rate was, the smaller
the agglomerated particles, the larger the total specific surface area, and the wider the particle size distribution. The
researchers took the particle size and particle surface area

299

300
200
100

0.05

0.1

x
Fig. 7 Distribution of particle diameter

as two independent variables to investigate the distributions


of particle-volume-equivalent diameters, the particle number
evolution, the geometric standard deviation based on particle
volume and the fractal nature of the agglomerated particles,
and the number of primary particles per agglomerate [73]
(Fig. 7) .
A low Reynolds number turbulent model was used by
Aristizabal et al. [74] to simulate the synthesis process of
aluminum nanoparticles by taking the particle coagulation as
the dominant mechanism in particle growth. They found that
the flow changes from laminar to turbulent in scale-up with a
constant residence time, which has a significant/insignificant
effect on particle properties for the reactors of small/large
length-to-diameter ratios, respectively. The synthesis of TiO2
nanoparticles in a low-pressure flat stagnation flames was
studied by Zhao et al. [75] studied using TTIP as precursor. It was found that both larger aggregate particles and
smaller primary particles were produced at higher pressures.
When the precursor-loading rates were higher, the aggregate
particle size was larger, whereas the primary particle size
remained constant in the experiment and decreased slightly
in the numerical simulation. Particle growth processes in the

123

300

postflame region of a premixed ethylene-air flame was investigated by De Filippo et al. [76]. The particle size distributions
were initially unimodal and then became bimodal in the postflame region. The results showed that the smaller mode could
be satisfactorily predicted with a size-dependent coagulation
model.
The aggregation and deposition kinetics of fullerene C60
nanoparticles was studied by Chen and Elimelech [77]. The
results showed that the aggregation kinetics of the particles exhibited reaction-limited and diffusion-limited regimes
at critical coagulation concentrations. The deposition rate
increased with increases in the electrolyte concentrations.
The Brownian and coagulation rates and deposition coefficients of nanoparticles in a closed chamber were measured
by Kim et al. [78]. The results showed that the deposition processes were overwhelmed by the coagulation
processes in the case of high particle number concentration. The larger turbulent coefficients made the turbulent
coagulation stronger. Particle coagulation rates were different in the small particle size range. The removal of
dispersant-stabilized carbon nanotube suspensions by polyaluminum chloride alum was studied by Liu [79]. It was
found that the polyaluminum chloride had a different effect
with alum in the removal of the suspensions. The reduction of particle number concentration first increased then
decreased with increases in coagulant dosage. The flocculation and coagulation of carbon nanotube suspensions
by polyaluminum chloride were regulated mainly by the
mechanism of adsorption charge neutralization, whereas the
coagulation by alum mainly involved electrical double-layer
compression.

9 Conclusions
This review presented an overview of the research progress
made in the Brownian coagulation of monodisperse and
polydisperse nanoparticles, the TEMOM, nanoparticle distributions due to coagulation in pipes and channel flows, jet
flows, and mixing layers, and in the process of flame synthesis and deposition.
Future investigations must include the following aspects:
(1) Contributions to nanoparticle coagulation resulting from
fluctuating particle concentrations;
(2) Determination of a collision kernel for two nonspherical
nanoparticles;
(3) Distributions of nanoparticles under nucleation, convection, diffusion, coagulation, and breakage in turbulent
flows.
Acknowledgments The project was supported by the Major Program
of the National Natural Science Foundation of China (Grant 11132008).

123

J. Lin, L. Huo

References
1. Uchowski, V.: Versuch einer mathematischen theorie der koagulation skinetik kollider losungen. Z. Phys. Chem. 92, 129168
(1917)
2. Russel, W.B., Saville, D.A., Schowalter, W.R.: Colloidal Dispersions. Cambridge University Press, Cambridge (1989)
3. Han, M.Y., Lee, H., Lawler, D.F., Choi, S.: Collision efficiency
factor in Brownian coagulation (Br) including hydrodynamics
and interparticle forces. Water Sci. Technol. 36, 6975 (1997)
4. Han, M.Y., Lee, H.: Collision efficiency factor in Brownian
coagulation(Br): calculation and experimental verification. Colloids Surf. A 202, 2331 (2002)
5. Vanni, M., Baldi, G.: Coagulation efficiency of colloidal particles
in shear flow. Adv. Colloid Interface Sci. 97, 151177 (2002)
6. Chin, C.J., Lu, S.C., Yiacoumi, S.: Fractal dimension of particle
aggregates in magnetic fields. Sep. Sci. Technol. 39, 28392862
(2004)
7. Olsen, A., Franks, G., Biggs, S.: An improved collision efficiency
model for particle aggregation. J. Chem. Phys. 125, 184906 (2006)
8. Chun, J., Koch, D.L.: The effects of non-continuum hydrodynamics
on the Brownian coagulation of aerosol particles. J. Aerosol Sci.
37, 471482 (2006)
9. Feng, Y., Lin, J.Z.: The collision efficiency of spherical dioctyl
phthalate aerosol particles in the Brownian coagulation. Chin. Phys.
B 17, 45474553 (2008)
10. Wang, Y.M., Lin, J.Z., Feng, Y.: The central oblique collision efficiency of spherical nanoparticles in the Brownian coagulation.
Mod. Phys. Lett. B 24, 15231531 (2010)
11. Chen, Z.L., You, Z.J.: New expression for collision efficiency
of spherical nanoparticles in Brownian coagulation. Appl. Math.
Mech (English Edition). 31, 851860 (2010)
12. Wang, Y.M., Lin, J.Z.: Attachment efficiency of polydisperse
nanoparticles wall-deposition. KONA Powder Part. J. 29, 158167
(2011)
13. Zhang, Y.Y., Li, S.Q., Yan, W., Yao, Q., Tse, S.D.: Role of dipole
dipole interaction on enhancing Brownian coagulation of chargeneutral nanoparticles in the free molecular regime. J. Chem. Phys.
134, 084501 (2011)
14. Wang, Y.M., Lin, J.Z.: The oblique collision efficiency of nanoparticles at different angles in Brownian coagulation. Comput. Math.
Appl. 61, 19171922 (2011)
15. Chen, Z.L., Jiang, R.J., Ku, X.K.: Collision efficiency of brownian
coagulation for nanoparticles taking into account the slip boundary
condition on the particle surface. Mod. Phys. Lett. B 26, 1250135
(2012)
16. Hawa, T., Zachariah, M.R.: Coalescence kinetics of unequal sized
nanoparticles. J. Aerosol Sci. 37, 115 (2006)
17. Wang, Y.M., Lin, J.Z.: Collision efficiency of two nanoparticles
with different diameters in Brownian coagulation. Appl. Math.
Mech. (English Edition). 32, 10191028 (2011)
18. Kelkar, A.V., Dong, J.N., Franses, E.I., Corti, D.S.: New models and
predictions for Brownian coagulation of non-interacting spheres.
J. Colliod Interface Sci. 389, 188198 (2013)
19. Muller, H.: Zur Allgemeinen Theorie der Raschen Koagulation.
Kolloideihefte 27, 223250 (1928) (in German)
20. Friedlander, S.K.: Smoke, Dust and Haze: Fundamentals of Aerosol
Behavior. Wiley, New York (2000)
21. Allen, M.D., Raabe, O.G.: Slip correction measurement of spherical solid aerosol particles in an improved millican apparatus.
Aerosol Sci. Technol. 4, 269286 (1985)
22. Fuchs, N.A.: The Mechanics of Aerosols. Pergamon, New York
(1964)
23. Otto, E., Fissan, H., Park, S.H., Lee, K.W.: The log-normal size
distribution theory of brownian aerosol coagulation for the entire

A review of research on nanoparticulate flow undergoing coagulation

24.

25.

26.

27.

28.
29.
30.

31.
32.

33.

34.

35.

36.

37.

38.
39.

40.

41.

42.

43.

44.

45.

particle size range: part II-analytical solution using Dahnekes


coagulation kernel. J. Aerosol Sci. 30, 1734 (1999)
Pratsinis, S.E.: Simultaneous nucleation, condensation, and coagulation in aerosol reactor. J. Colloid Interface Sci. 124, 416417
(1988)
Hulbert, H.M., Katz, S.: Some problems in particle technology: a
statistical mechanical formulation. Chem. Eng. Sci. 19, 555574
(1994)
Lin, J.Z., Chan, T.L., Liu, S., Zhou, K., Zhou, Y., Lee, S.C.: Effects
of coherent structures on nanoparticle coagulation and dispersion
in a round jet. Int. J. Nonlinear Sci. Numer. Simul. 8, 4554 (2007)
Upadhyay, R.R., Ezekoye, O.A.: Evaluation of the 1-point quadrature approximation in QMOM for combined aerosol growth laws.
J. Aerosol Sci. 34, 16651683 (2003)
Pratsinis, S.E.: Receptor models for ambient carbonaceous
aerosols. Aerosol Sci. Technol. 10, 258266 (1989)
Mcgraw, R.: Description of aerosol dynamics by the quadrature
method of moments. Aerosol Sci. Technol. 27, 255265 (1997)
Yu, M.Z., Lin, J.Z., Chan, T.L.: A new moment method for solving
the coagulation equation for particles in Brownian motion. Aerosol
Sci. Technol. 42, 705713 (2008)
Lee, K.W., Chen, H.: Coagulation rate of polydisperse particles.
Aerosol Sci. Technol. 3, 327334 (1984)
Yu, M.Z., Lin, J.Z.: Binary homogeneous nucleation and growth
of water-sulfuric acid nanoparticles using a TEMOM model. Int.
J. Heat Mass Transfer 53, 635644 (2010)
Wang, W.X., He, Q., Chen, N.A., Xie, M.L.: A simple moment
model to study the effect of diffusion on the coagulation of nanoparticles due to Brownian motion in the free molecule regime. Therm.
Sci. 16, 13311338 (2012)
Lin, J.Z., Chen, Z.L.: A modified TEMOM model for Brownian
coagulation of nanoparticles based on the asymptotic solution of the
sectional method. Sci. China Technol. Sci. 56, 30813092 (2013)
Xie, M.L., He, Q.: The fundamental aspects of TEMOM model
for particle coagulation due to Brownian motion. Part 1: in the free
molecule regimes. Int. J. Heat Mass Transfer 70, 11151120 (2014)
Chen, Z.L., Lin, J.Z., Yu, M.Z.: Direct expansion method of
moments for nanoparticle Brownian coagulation in the entire size
regime. J. Aerosol Sci. 67, 2837 (2014)
De Bleecker, K., Bogaerts, A., Goedheer, W.: Modelling of
nanoparticle coagulation and transport dynamics in dusty silane
discharges. New J. Phys. 8, 178181 (2006)
Yin, Z.Q., Lou, M.: Experimental study on nanoparticle deposition
in straight pipe flow. Therm. Sci. 16, 14101413 (2012)
Lin, J.Z., Liu, S., Chan, T.L.: Nanoparticle migration in a fully
developed turbulent pipe flow considering the particle coagulation.
Chin. J. Chem. Eng. 20, 679685 (2012)
Lin, J.Z., Yin, Z.Q., Gan, F.J., Yu, M.Z.: Penetration efficiency and
distribution of aerosol particles in turbulent pipe flow undergoing
coagulation and breakage. Int. J. Multiph. Flow 61, 2836 (2014)
Lin, P.F., Lin, J.Z.: Transport and deposition of nanoparticles in
bend tube with circular cross-section. Prog. Nat. Sci. 19, 3339
(2009)
Lin, P.F., Lin, J.Z.: Prediction of nanoparticle transport and deposition in bends. Appl. Math. Mech. (English Edition). 30, 957968
(2009)
Lin, J.Z., Lin, P.F., Yu, M.Z., Chen, H.J.: Nanoparticle transport and
coagulation in bends of circular cross section via a new moment
method. Chin. J. Chem. Eng. 18, 19 (2010)
Lin, J.Z., Lin, P.F., Chen, H.J.: Research on the transport and deposition of nanoparticles in a rotating curved pipe. Phys. Fluids 21,
122001 (2009)
Lin, J.Z., Lin, P.F., Chen, H.J.: Nanoparticle distribution in a rotating curved pipe considering coagulation and dispersion. Sci. China
Phys. Mech. Astron. 54, 15021513 (2011)

301
46. Chan, T.L., Lin, J.Z., Zhou, K., Chan, C.K.: Simultaneous numerical simulation of nano and fine particle coagulation and dispersion
in a round jet. J. Aerosol Sci. 37, 15451561 (2006)
47. Lin, J.Z., Chan, T.L., Liu, S., Zhou, Y., Lee, S.C.: Effects of coherent structures on nanoparticle coagulation and dispersion in a round
jet. Int. J. Nonlinear Sci. Numer. Simul. 8, 4554 (2007)
48. Yin, Z.Q., Lin, J.Z., Zhou, K., Chan, T.L.: Numerical simulation
of the formation of pollutant nanoparticles in the exhaust twin-jet
plume of a moving car. Int. J. Nonlinear Sci. Numer. Simul. 8,
535543 (2007)
49. Yin, Z.Q., Lin, J.Z.: Numerical simulation of the formation of
nanoparticles in an impinging twin-jet. J. Hydrodyn. 19, 533541
(2007)
50. Fujitani, Y., Hirano, S., Kobayashi, S., Tanabe, K., Suzuki, A.,
Furuyama, A., Kobayashi, T.: Characterization of dilution conditions for diesel nanoparticle inhalation studies. Inhalation Toxicol.
21, 200209 (2009)
51. Chan, T.L., Zhou, K., Lin, J.Z., Liu, C.H.: Vehicular exhaust gasto-nanoparticle conversion and concentration distribution in the
vehicle wake region. Int. J. Nonlinear Sci. Numer. Simul. 11, 581
593 (2010)
52. Zhu, J.Z., Qi, H.Y., Wang, J.S.: Nanoparticle dispersion and coagulation in a turbulent round jet. Int. J. Multiph. Flow 54, 2230
(2013)
53. Yu, M.Z., Lin, J.Z., Chen, L.H., Chan, T.L.: Large eddy simulation
of a planar jet flow with nanoparticle coagulation. Acta Mech. Sin.
22, 293300 (2006)
54. Yu, M.Z., Lin, J.Z., Chen, L.H.: Nanoparticle coagulation in a planar jet via moment method. Appl. Math. Mech. (English Edition).
28, 14451453 (2007)
55. Lu, Y.H.: Nanoparticle nucleation and coagulation in a submerged
jet: theoretical prediction and simulation. Int. J. Nonlinear Sci.
Numer. Simul. 10, 11891200 (2009)
56. Das, S., Garrick, S.C.: The effects of turbulence on nanoparticle growth in turbulent reacting jets. Phys. Fluids 22, 103303
(2010)
57. Loeffler, J., Das, S., Garrick, S.C.: Large eddy simulation of titanium dioxide nanoparticle formation and growth in turbulent jets.
Aerosol Sci. Technol. 45, 616628 (2011)
58. Garrick, S.C., Wang, G.H.: Modeling and simulation of titanium
dioxide nanoparticle synthesis with finite-rate sintering in planar
jets. J. Nanopart. Res. 13, 973984 (2011)
59. Settumba, N., Garrick, S.C.: Direct numerical simulation of
nanoparticle coagulation in a temporal mixing layer via a moment
method. J. Aerosol Sci. 34, 149167 (2003)
60. Settumba, N., Garrick, S.C.: A comparison of diffusive transport
in a moment method for nanoparticle coagulation. J. Aerosol Sci.
35, 93101 (2004)
61. Garrick, S.C., Lehtinen, K.E.J., Zachariah, M.R.: Nanoparticle
coagulation via a NavierStokes/nodal methodology: evolution of
the particle field. J. Aerosol Sci. 37, 555576 (2006)
62. Wang, G.H., Garrick, S.C.: Modeling and simulation of titania formation and growth in temporal mixing layers. J. Aerosol Sci. 37,
431451 (2006)
63. Lin, J.Z., Liu, Y.H.: Nanoparticle nucleation and coagulation in a
mixing layer. Acta Mech. Sin. 26, 521529 (2010)
64. Xie, M.L., Yu, M.Z., Wang, L.P.: A TEMOM model to simulate
nanoparticle growth in the temporal mixing layer due to Brownian
coagulation. J. Aerosol Sci. 54, 3248 (2012)
65. Rosner, D.E., Pyykonen, J.J.: Bivariate moment simulation of coagulating and sintering nanoparticles in flames. AIChE J. 48, 476491
(2002)
66. Kim, H.J., Jeong, J.I., Park, Y.: Modeling of generation and growth
of non-spherical nanoparticles in a co-flow flame. J. Nanopart. Res.
5, 237246 (2003)

123

302
67. Tsantilis, S., Pratsinis, S.E.: Narrowing the size distribution of
aerosol-made titania by surface growth and coagulation. J. Aerosol
Sci. 35, 405420 (2004)
68. Kostoglou, M., Konstandopoulos, A.G., Friedlander, S.K.: Bivariate population dynamics simulation of fractal aerosol aggregate
coagulation and restructuring. J. Aerosol Sci. 37, 11021115
(2006)
69. Morgan, N.M., Wells, C.G., Goodson, M.J., Kraft, M., Wagner,
W.: A new numerical approach for the simulation of the growth of
inorganic nanoparticles. J. Comput. Phys. 211, 638658 (2006)
70. Morgan, N., Kraft, M., Balthasar, M., Wong, D., Frenklach, M.,
Mitchell, P.: Numerical simulations of soot aggregation in premixed laminar flames. Proc. Combust. Inst. 31, 693700 (2007)
71. Starchenko, V., Muller, M., Lebovka, N.: Growth of polyelectrolyte
complex nanoparticles: computer simulations and experiments. J.
Phys. Chem. C 112, 88638869 (2008)
72. Yu, M.Z., Lin, J.Z., Chen, L.H., Chan, T.L.: Effect of precursor
loading on non-spherical TiO2 nanoparticle synthesis in a diffusion
flame reactor. Chem. Eng. Sci. 63, 23172329 (2008)
73. Yu, M.Z., Lin, J.Z., Chan, T.L.: Numerical simulation of nanoparticle synthesis in diffusion flame reactor. Powder Technol. 181, 920
(2008)

123

J. Lin, L. Huo
74. Aristizabal, F., Munz, R.J., Berk, D.: Turbulent modeling of the
production of ultra fine aluminum particles: scale-up. Aerosol Sci.
Technol. 42, 556565 (2008)
75. Zhao, H., Liu, X.F., Tse, S.D.: Effects of pressure and precursor
loading in the flame synthesis of titania nanoparticles. J. Aerosol
Sci. 40, 919937 (2009)
76. De Filippo, A., Sgro, L.A., Lanzuolo, G., DAlessio, A.: Probe
measurements and numerical model predictions of evolving size
distributions in premixed flames. Combust. Flame 156, 17441754
(2009)
77. Chen, K.L., Elimelech, M.: Aggregation and deposition kinetics of
fullerene (c-60) nanoparticles. Langmuir 22, 1099411001 (2006)
78. Kim, D.S., Hong, S.B., Kim, Y.J., Lee, K.W.: Deposition and coagulation of polydisperse nanoparticles by Brownian motion and
turbulence. J. Aerosol Sci. 37, 17811787 (2006)
79. Liu, N., Liu, C.L., Zhang, J., Lin, D.H.: Removal of dispersantstabilized carbon nanotubes by regular coagulants. J. Environ. Sci.
24, 13641370 (2012)

Вам также может понравиться