ArticlePDF Available

Mass and heat transfer between evaporation and condensation surfaces: Atomistic simulation and solution of Boltzmann kinetic equation

Authors:

Abstract and Figures

Boundary conditions required for numerical solution of the Boltzmann kinetic equation (BKE) for mass/heat transfer between evaporation and condensation surfaces are analyzed by comparison of BKE results with molecular dynamics (MD) simulations. Lennard–Jones potential with parameters corresponding to solid argon is used to simulate evaporation from the hot side, nonequilibrium vapor flow with a Knudsen number of about 0.02, and condensation on the cold side of the condensed phase. The equilibrium density of vapor obtained in MD simulation of phase coexistence is used in BKE calculations for consistency of BKE results with MD data. The collision cross-section is also adjusted to provide a thermal flux in vapor identical to that in MD. Our MD simulations of evaporation toward a nonreflective absorbing boundary show that the velocity distribution function (VDF) of evaporated atoms has the nearly semi-Maxwellian shape because the binding energy of atoms evaporated from the interphase layer between bulk phase and vapor is much smaller than the cohesive energy in the condensed phase. Indeed, the calculated temperature and density profiles within the interphase layer indicate that the averaged kinetic energy of atoms remains near-constant with decreasing density almost until the interphase edge. Using consistent BKE and MD methods, the profiles of gas density, mass velocity, and temperatures together with VDFs in a gap of many mean free paths between the evaporation and condensation surfaces are obtained and compared. We demonstrate that the best fit of BKE results with MD simulations can be achieved with the evaporation and condensation coefficients both close to unity.
Content may be subject to copyright.
SPECIAL FEATURE: PERSPECTIVE
Mass and heat transfer between evaporation and
condensation surfaces: Atomistic simulation and
solution of Boltzmann kinetic equation
Vasily V. Zhakhovsky (Василий Жаховский)
a,1
, Alexei P. Kryukov
b
, Vladimir Yu. Levashov
b,c
,
Irina N. Shishkova
b
, and Sergey I. Anisimov
d
Edited by William A. Goddard III, California Institute of Technology, Pasadena, CA, and approved March 21, 2018 (received for review
December 25, 2017)
Boundary conditions required for numerical solution of the Boltzmann kinetic equation (BKE) for mass/heat
transfer between evaporation and condensation surfaces are analyzed by comparison of BKE results with
molecular dynamics (MD) simulations. LennardJones potential with parameters corresponding to solid
argon is used to simulate evaporation from the hot side, nonequilibrium vapor flow with a Knudsen
number of about 0.02, and condensation on the cold side of the condensed phase. The equilibrium density
of vapor obtained in MD simulation of phase coexistence is used in BKE calculations for consistency of BKE
results with MD data. The collision cross-section is also adjusted to provide a thermal flux in vapor identical
to that in MD. Our MD simulations of evaporation toward a nonreflective absorbing boundary show that
the velocity distribution function (VDF) of evaporated atoms has the nearly semi-Maxwellian shape be-
cause the binding energy of atoms evaporated from the interphase layer between bulk phase and vapor is
much smaller than the cohesive energy in the condensed phase. Indeed, the calculated temperature and
density profiles within the interphase layer indicate that the averaged kinetic energy of atoms remains
near-constant with decreasing density almost until the interphase edge. Using consistent BKE and MD
methods, the profiles of gas density, mass velocity, and temperatures together with VDFs in a gap of many
mean free paths between the evaporation and condensation surfaces are obtained and compared. We
demonstrate that the best fit of BKE results with MD simulations can be achieved with the evaporation and
condensation coefficients both close to unity.
evaporation
|
condensation
|
Boltzmann kinetic equation
|
molecular dynamics
Evaporation and condensation can be realized in
different natural phenomena and technologies. A
peculiarity of these processes is the coupled mass
and heat transfer from evaporation to condensation
surface. Nowadays, a correct description of the trans-
port processes across the interfacial surfaces is required
for the development of new advanced technologies
and solution of the known engineering problems.
Among them, the problem of removing heat from
space vehicles, developing technologies based on the
interaction of matter in the form of cryogenic corpus-
cular targets with high-energy beams, evaporation of
droplets on superhydrophobic surfaces using liquid
droplets as the molecular concentrators of ultradilute
solutions, the development of effective vacuum drying
methodsall of these tasks are inextricably linked with
the solution of the evaporationcondensation problem.
The main goal in considering the coupled mass/heat
transfer is an accurate evaluation of the masses of
evaporated or condensed material. Determination of
these quantities is important because a significant heat
can be diverted from the interfacial surface, which, as a
consequence, leads to cooling of the condensed phase.
For example, when considering the evaporation of a
liquid droplet placed in a steamgas mixture, the inac-
curacies in calculating the evaporation rate can lead to
a
Center for Fundamental and Applied Research,Dukhov Research Instituteof Automatics, Moscow127055, Russia;
b
Departmentof Low Temperatures,
Moscow Power Engineering Institute, Moscow 111250, Russia;
c
Institute of Mechanics, Lomonosov Moscow State University, Moscow 119192,
Russia; and
d
Landau Institute for Theoretical Physics, Russian Academy of Science, Chernogolovka 142432, Russia
Author contributions: S.I.A. designed research; V.V.Z., A.P.K., V.Y.L., and I.N.S. performed research; V.V.Z., A.P.K., V.Y.L., and S.I.A. analyzed data;
and V.V.Z., A.P.K., and V.Y.L. wrote the paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
Published under the PNAS license.
1
To whom correspondence should be addressed. Email: 6asi1z@gmail.com.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1714503115/-/DCSupplemental.
www.pnas.org/cgi/doi/10.1073/pnas.1714503115 PNAS Latest Articles
|
1of9
SPECIAL FEATURE:
PERSPECTIVE
errors in determining the pattern of temperature changes at the in-
terfacial surface, which may result in inaccurate forecasts of the drop-
let temperature and a complete evaporation time.
As noted in ref. 1, despite the simplicity of the formulation of
the evaporationcondensation problem, its solution encounters
certain difficulties in the general case. Traditionally, it is assumed
that the heat coming to the interphase boundary is spent on
evaporation and heating of the particles, and the resulting vapor
is diverted from the evaporation surface by diffusion. It is believed
that the concentration of the evaporated gas near the interphase
boundary is equal to the equilibrium vapor concentration. This
assumption was first used in ref. 2 to consider evaporation of a
spherical drop. However, this is not the case, since the vapor near
the interfacial surface will be saturated only if a diffusion rate of
gas escape is lower than an arrival rate of molecules from the
interphase boundary. Ref. 3 showed the existence of a concentra-
tion jump near the interphase boundary. One of the refinements
of the evaporationcondensation theory can be achieved by in-
voking the kinetics of interaction of vapor molecules with the
surface of the liquid phase and also with each other. In refs. 4
and 5, using the methods of molecularkinetic theory, a mass flux
of evaporated molecules was evaluated. A disadvantage of the
formula proposed in these papers lies in the fact that it was
obtained for a free molecular flow of the evaporated gasthat
is, for the conditions where emitted particles do not interact with
molecules presented near the surface.
The next stage in the study of evaporationcondensation dates
back to the 1960s, when the dynamics of rarefied gases was de-
veloped rapidly. The needs of technology development led to the
emergence of more rigorous calculation techniques based on the
exact or approximate solution of the Boltzmann kinetic equation
(BKE). A linear theory was being formulated at this time. The
beginning was laid by the work in ref. 6, in which the form of
the velocity distribution function (VDF) near the interphase
boundary was adopted rather than in the derivation of the
HertzKnudsen formula (4, 5). The authors suggested that the
VDF for molecules moving to this boundary is the same as for a
negative half-space of velocities at a considerable distance
from the interface of the phases. Then, by writing down the
expression for the mass flow from definition, they obtained a
result that was two times different from the mass flow calcu-
lated by the HertzKnudsen formula.
Linearized, or more simply a linear theory of, evaporation and
condensation was developed by Labuntsov and Muratova in refs.
7 and 8. At about the same time, many researchers (916) and
Anisimov et al. (1719) were focused on the solution of nonlinear
evaporationcondensation problems with the use of the kinetic
theory of gases.
With the advancement of computers, the direct numerical
solution of the BKE began to be applied to the evaporationcon-
densation problems (2024). For solving the BKE, it is necessary to
specify the correct boundary conditions for the VDFs of evapo-
rated and condensed molecules. The shapes of VDFs together
with the evaporation and condensation coefficients, determining
the corresponding fluxes through the evaporation and condensa-
tion surfaces, are involved in those boundary conditions. In the
previously listed works, a semi-Maxwellian VDF with zero trans-
port velocity was taken as such a function. However, as noted in
ref. 25, no serious theoretical conclusion of such a boundary
condition is known to us.
The measured evaporation and condensation coefficients may
vary greatly from experiment to experiment. The condensation
coefficient for water ranges from 0.01 to 1 as noted in the review
(26). It seems likely that such a wide spread can be explained by
the fact that those coefficients were measured not at the interface
but over a distance of many mean free paths in the vapor. More-
over, even the small differences between the experimental con-
ditions (such as chemical impurities on the interface and variation
of surface temperatures of liquids being investigated) may have a
dramatic effect on the measurement results. The dependence of
evaporation and condensation coefficients from the surface tem-
perature is also reported in simulation works (27, 28). It was found
that those coefficients, which are close to unity at low tempera-
tures, begin to decrease if the surface temperature is increased
well above the triple point.
It should be noted that the molecularkinetic approach allows
us to correctly describe the change in the macroparameters of the
vapor/gas near the interfacial surface, but it is assumed that the
state of the condensed phase remains unchanged. On the other
hand, the VDF of molecules escaping from the surface can be
affected by processes occurring near this surface, both from the
liquid side and from the vapor side. In this connection, the ap-
proach in which both the condensed and vapor phases are con-
sidered within the framework of a single modeling method is
obvious. As a research method, the molecular dynamics (MD) simu-
lation has been widely used presently. There are several known works
in which the calculation of the VDF of molecules emitted from the
interphase boundary layer is performed by the MD method (25, 29
31) and the analysis of the simulation results lead to the conclusion
about the proximity of the VDF of vapor molecules flyingfrom the
interphase to the Maxwellian distribution.
In refs. 28 and 32, the problem of recondensation through a
small vapor gap with the thickness of about 7 nm, which leads to a
Knudsen flow with Kn 1 roughly, is investigated via MD simula-
tion and by the solution of the EnskogVlasov equation with the
Direct Simulation Monte Carlo (DSMC) method. The authors came
to the following conclusion: On the basis of the results of this
study, we constructed the kinetic boundary conditions (KBC) for
the hard-sphere molecules in consideration of the liquid temper-
ature dependence in the course of a steady net evaporation and
condensation; however, the application of the KBC during un-
steady net evaporation-condensation is extremely important.
Furthermore, the same two-surface method used in refs. 28 and
32 was applied in ref. 33 to study the evaporationcondensation
in an unsteady transient regime on a short MD time scale. The
authors of ref. 34 noted that there is a deviation from the Maxwel-
lian VDF for particles flying in the direction normal to the interfa-
cial surface: However, the detailed mechanism of the deviation
has not been clarified yet.
Numerous papers using the MD method indicate that an
important question in determining the KBC is at which position of
the boundary between the liquid and gas phases those KBC
should be determined (27, 32, 34, 35). The authors of ref. 35
discuss the influence of the KBC position on the mass flow of
the evaporating substance.
It seems obvious the way in which a joint (cross-linked) version
of the description is used. That is, the liquid phase and the region
near the interphase boundary layer are described by the MD
method, next the methods of the kinetic theory of gases are used,
and finally the continuum mechanics approaches are applied on
distances larger than the 10 to 20 mean free paths. However, such
attempts are faced with certain problems and difficulties. The
characteristic time scale of the MD processes and the kinetic re-
laxation time differ approximately by a factor of 10
4
. Therefore, more
than 10
5
simulation steps must be taken to trace the behavior of an
atomic system during the average interatomic collision time. The MD
method deals with the coordinates and velocities of the particles, but
for a molecularkinetic approach, this description of the motion of
2of9
|
www.pnas.org/cgi/doi/10.1073/pnas.1714503115 Zhakhovsky et al.
the gas is unnecessarily complete. Therefore, we must resort to a less
complete statistical description of the behavior of the systemas a
note in ref. 36. Also, the results of molecularkinetic calculations in the
form of VDF cannot be used to obtain information on the particle
coordinates and velocities necessary for MD simulation. Accordingly,
the mutual exchange of results of calculations obtained by molecular
kinetic and MD methods becomes problematic.
In this work, a nonequilibrium vapor flow from the evaporation
to condensation boundary layer is considered in detail on an
atomic scale using the MD method, which provides the classical
trajectories of interacting atoms starting from a hot bulk material
and reaching a cold material after traveling through a gas gap. To
simplify the modeling of gas flow, the exact atom positions r
i
and
velocities v
i
can be replaced by a probability density function
fðr,v,tÞto find dN molecules having positions near rand veloci-
ties near vwithin a phase-space volume drdvat time tsuch as
dN ¼fðr,v,tÞdrdv. Upon integrating fðr,v,tÞ, also called a distri-
bution function, over a whole range of velocity, we obtain a local
number density nðr,tÞof molecules. In a volume with spatially
uniform distribution of molecules, a VDF Fðv,tÞcan be obtained
by integrating fðr,v,tÞdrover the volume.
The BKE (36) describing evolution of the fðr,v,tÞat the ab-
sence of external forces is given by
f
t+vf
r¼I,[1]
where rðx,y,zÞare Cartesian coordinates, vðυx,υy,υzÞis the
molecule velocity in a laboratory coordinate system at time t,
and Iis a collision integral. The different forms of the collisional
integral are considered in refs. 20 and 21. Here in this work, the
simplest form for the hard sphere collisions is used.
Our prime goal is to make a bridge from atomistic simulation to
a probabilistic approach through a more penetrating insight into
the atomic-scale mechanism of evaporation and condensation,
which determines the VDFs at the corresponding interfaces. We
performed the large-scale MD simulations to find the best
boundary conditions, including their positions, evaporation and
condensation coefficients, and the shapes of VDFs, which can be
used in the solution of BKE providing the best agreement with data
obtained from the MD simulations. For consistency of BKE with MD
results, the saturated vapor density and transport cross-section
were precomputed by MD and then used in the BKE method.
Simulation Techniques
Equilibrium evaporation and condensation processes take place in an
interphase transition layer between the coexisting condensed phase
and its vapor. For MD simulation of nonequilibrium evaporation
condensation, the two surfaces of condensed phase at different
temperatures are required. This condition can be complied with using
both sides of a single film as illustrated in Fig. 1, where the condensed
phase of argon is placed in the middle of the MD computational
domain. Periodical boundary conditions are imposed on all three di-
mensions of the domain, which has typical dimensions of Lx¼400nm
and Ly¼Lz¼200nm. The total number of atoms was about
29.15 million in our MD simulations of evaporationcondensation.
To establish both evaporation and condensation processes in a
single simulation, a temperature gradient is maintained by the Lan-
gevin thermostat with target temperature TðxÞdepending on atom
position in the film. The thermostat is only applied to atoms moving in
a gray zone nearby the center of film as shown in Fig. 1; thus, the
interphase layers between the condensed phase and vapor are not
acted upon by the thermostat forces. The Langevin forces are given by
mid~
υi=dt ¼~
ξiγ~
υi~
ux,[2]
where the force acting on each atom iis a sum of a Gaussian-
distributed random force~
ξiand a frictional damping term (37).
The last is determined by a friction coefficient γand a thermal
velocity of atom in reference to a target flow velocity ~
ux.To
reproduce the temperature gradient within the film, the dis-
persion hξ2iof random forces and the friction coefficient must
satisfy the condition hξ2iΔt=2γ¼TðxÞ, where Δtis an MD sim-
ulation time step. Then, the dispersion becomes a function of
-200 -100 0 100 200
position x(nm)
0.1
1
10
0.2
0.3
0.5
2
3
5
20
0.05
number density n (atoms/nm3)
65
70
75
80
85
temperatures Tx and Ty (K)
thermostat
evaporationcondensation
200 nm
Th
Tc
n
Tx
Ty
Fig. 1. MD computational domain for evaporation from the right hot
side of liquid film and condensation on the left cold side of the same
film placed in periodical conditions. Profiles of atom number density,
longitudinal T
x
,andtransverseT
y
temperatures are taken from MD
simulation providing Tc¼72.3 for a cold surface and Th¼80.4K for a hot
one. The Langevin thermostat maintains a required temperature gradient
for atoms in a gray zone xi[7,7nm] inside the film with a thickness of
32.7nm. The thermostat also keeps the mass center of film at rest by
adjusting the mass flow velocity from the cold to the hot side. Here this
flow velocity is ux¼0.129m=satthex¼0. See details in Fig. S1.
T
-4
-3
-2
-1
0
1
ns3
-1
0
1
2
3
4
P
P
P
ns
Fig. 2. Pressure and atom number density n
s
of saturated vapor in
equilibrium with the condensed phase of argon. Red line fitted to the
vapor density in the liquidvapor system and the blue line fitted to
the vapor density in the solidvapor system are used for estimating
vapor density between data points. The experimental data from ref.
38, pp 6-128 and 15-10. The triple point of simulated argon
Tt.p.¼76.7K is lower than the experimental Tt.p.¼83.8058K.
Zhakhovsky et al. PNAS Latest Articles
|
3of9
atom position x
i
because the friction coefficient is set to a
constant in our MD simulation.
To maintain the steady positive atom flux across the MD do-
main and keep the film at rest, the Langevin target velocity ~
uxis
coupled with a displacement of film mass center using a negative
feedback control. After reaching the stationary regime of vapor
flow, such feedback control provides almost a fixed position of the
film with negligible irregular fluctuations of mass center in the
range less than ±0.01nm, which is much smaller than a thickness
of the interphase layer of 2nm for liquid argon at T¼80K. For
steady evaporationcondensation, the atom flux in direction xis a
constant everywhere regardless of local density. As a result, the mass
velocity at the center of the film is less than gas flow speed by a factor
equal to a density ratio between condensed phase and vapor. In the
simulation shown in Fig. 1, the vapor mass velocity is about 40m=sin
the middle of the gap between evaporation and condensation sur-
faces, while the liquid flows with 0.129m=s near the center of film,
with the thickness of 32.7nm defined as a distance between positions
of hot and cold surfaces having temperatures T
h
and T
c
, respectively.
Determination of surface positions is discussed in Evaporation Co-
efficient from MD Simulation. The detailed profiles of target tem-
perature and flow variables in the film and surrounding vapor are
presented in SI Appendix.
MD simulations were performed with a smoothed Lennard
Jones (LJ) potential (39) given by
ϕðrÞ¼4"hð=rÞ12 ð=rÞ6i+a2x2a3x3,[3]
where ris an interatomic distance and x¼ðr=Þ2ðr0=Þ2. The
coefficients "¼1.0312kJ=mol, ¼0.33841nm, a2¼1.3647
104kJ=mol, and a3¼2.4614104kJ=mol are fitted to repro-
duce argon fcc crystal at zero temperature, which has the
lattice parameter of 0.524673 nm and the cohesive energy
of 7.74005 kJ/mol. The position of the potential minimum
r0¼21=6is identical to that in the original LJ potential.
The smoothing coefficients a
2
and a
3
were chosen to satisfy
the conditions ϕðrcÞ¼0andϕðrcÞ¼0 at the cutoff radius
rc¼0.8125nm.
A density of vapor in equilibrium with the condensed phase is
required to run the BKE calculation of evaporation and conden-
sation because the density nsðTÞof saturated vapor is used to set a
boundary condition for VDF. Using the smoothed L-J potential Eq.
3the vapor pressure and atom number density are evaluated in
several MD simulations of equilibrium liquid-vapor and solid-
vapor systems. Fig. 2 shows the calculated and experimental
values for argon. The visible difference in vapor pressure indicates
that the smoothed LJ potential, fitted to the experimental
parameters of solid Ar, overestimates the pressure. Using the
phase-coexistence MD method, we obtained all three phases
in equilibrium at the triple point Tt.p.¼76.7K of simulated ar-
gon, which is appreciably lower than the experimental triple
point Tt.p.¼83.8058K (38).
Numerical solution of BKE of the evaporationcondensation
problem is performed only for vapor gaps between two interfaces
defined from MD simulated profiles as the outer boundaries of in-
terphase transition layers from the bulk condensed phase to vapor.
While the results of BKE calculations are almost insensitive to the
delimitation of interfaces, they depend highly on the definition of
interface temperatures. Those temperatures are also evaluated
from MD simulated profiles at some position inside the transition
layer. The interface definitions are introduced in the next section.
With the vapor density function, positions of interfaces, and their
temperatures provided by MD simulations, the BKE Eq. 1can be
solved numerically in the vapor gap. We use a finite-difference
computational method described in refs. 20, 21, and 40, in which
a spherical velocity domain is represented by a discrete 3D mesh of
velocity nodes. The discretized BKE equation for each node is solved
in two steps. First, the spatial displacements are calculated without
collisions, and the Courant condition used for a time step Δtguar-
antees that even fastest nodes cannot move more than one spatial
step Δx. Then the collisions are calculated and taken into account.
The collisional integral is evaluated by the quasi Monte Carlo
method using the Korobovs pseudorandom sequences (21). For
simplicity, the interatomic collision of LJ atoms is considered as
a collision between hard spheres of diameter d.Thelastisan
adjustable parameter that must be determined from MD simu-
lation for consistency between BKE and MD methods. The
collision cross-section determined by the diameter dis ad-
justed to provide a steady heat flux close to that derived from
-200 -100 0 100 200
position x(nm)
70
80
90
temperatures T
x
and T
y
(K)
0.028
0.032
0.036
density n (atoms/nm
-3
)
n
T
x
T
y
246 kW/m
2
Fig. 3. MD simulation of steady heat flux from the left heating
thermostat with Th¼90K to the right cooling thermostat with
Tc¼70K applied in the gray zones. A transport cross-section for BKE
is fitted to get the same heat flux of 246kW=m2produced by the
temperature gradient of 0.05K=nm imposed at x¼0.
x
0.3
3
0.03
n
3
T
x
T
y
T
x
T
y
n
0
1
2
3
4
5
K
x
T
h
Fig. 4. Number density and temperatures profiles in a hot interphase
layer and a small vapor gap lgap ¼1.5nm bounded by the Maxwell
demon. The left boundary of the nonequilibrium part of the
interphase layer (the left vertical dashed line) is determined by a
point of divergence between longitudinal T
x
and transverse T
y
temperatures. This point is used for definition of the surface
temperature T
h
. The right boundary (the right dashed line)
corresponds to a point where density and T
x
starts to decrease
linearly. The Maxwell demon ensures that all atoms in the gray zone
have υx>0. Numbers along the density profile indicate positions
where velocity data for VDFs are gathered. Atoms pass the
interphase almost without loss of their longitudinal component of
kinetic energy K
x
expressed in the temperature units.
4of9
|
www.pnas.org/cgi/doi/10.1073/pnas.1714503115 Zhakhovsky et al.
MD simulation of heat transfer between hot and cold zones of
gas gap shown in Fig. 3. LJ atoms within the left hot and right
coldzonesintheMDdomainwitharigidwallat±200nm are
subjected to two Langevin thermostats supporting 90K and 70K,
respectively. Initially atoms were placed in the domain at a constant
density of n¼0.032nm3. After reaching a steady regime, the heat
flux, the number density, and temperature gradient were measured
at the center x¼0. Using those MD data, the cross-section was
found to equal d¼0.55nm to reproduce the almost identical heat
flux of 245kW=m2in BKE calculations.
For the vapor density n¼0.0963nm3in equilibrium with LJ
liquid at T¼80K, a mean free path in a hard sphere system is
¼1=ffiffiffi
2
pnd2¼7.73nm, which gives the Knudsen number
Kn ¼=Lgap ¼0.021 for the vapor gap length of Lgap ¼364.3nm
shown in Fig. 1.
To obtain evaporation and condensation coefficients, which
are the basic parameters governing the boundary conditions in
the BKE method, we use the flow profiles and VDFs gained from
atomistic trajectories simulated by the MD method. The most
appropriated coefficients can be found via comparison of VDFs
from numerical solution of BKE with VDFs obtained from the MD
simulation. The VDFs in a steady vapor flow at xpositions are
calculated as Fðx,υx,tÞ¼Rdydz Rdυydυzfðr,v,tÞ. For steady
flow, the distributions are independent of time, which allows us to
accumulate atom position and velocity statistics during MD sim-
ulation. Further integration over υxgives an atom number density
profile nðxÞ¼RdυxFðx,υxÞ.
The steady profiles of density, mass flow velocity, and tem-
perature are obtained by averaging the corresponding values in
spatial slabs with the small thickness of 0.02nm along the xaxis
and during the entire time of productive MD simulation, which is
performed after the attainment of a steady regime. The profile of
flow velocity uxðxÞ¼hυxi, the longitudinal TxðxÞ¼m
kBυxuxÞ2i,
and transverse TyðxÞ¼ m
2kBhυ2
y+υ2
zitemperatures are calculated
from the corresponding components of atom velocities υx,υy,
and υz.
Evaporation Coefficient from MD Simulation
The transition of atoms from bulk phase to vapor can be
imagined as a jump over a potential barrier with the height
equal to the atom binding energy "
b
in the condensed phase.
It is assumed in this naive model that the barrier is infinitely
thin and an atom should spend its kinetic energy to over-
come the barrier. As a result, the atoms with kinetic energy "
l
in direction xtoward the vapor exceeding the binding en-
ergy "l>"
bcan go to the vapor phase, where the subscript l
indicates a condensed phase. Taking into account a new ki-
netic energy of atom in vapor "¼"l"bafter passing the
barrier, one can obtain d"¼mυdυ¼d"l¼mυldυl, where veloci-
ties are along the xaxis. Hence, assuming the Maxwellian VDF for
υl, the evaporated atoms obtain a new VDF given by ref. 41:
fxdυ¼Amυdυ
ffiffiffiffiffiffiffiffiffiffiffiffiffi
4kBT
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
"b+mυ2=2
pexp"b+mυ22
kBT,[4]
where a new atom velocity υis positive (i.e., directed toward
the vapor) and Ais a normalization factor (41). The VDFs f
y
and
f
z
do not change during evaporation.
The above-stated simple model ignores the density re-
distribution and energy transfer inside the interphase boundary
layer having the finite thickness, which makes a real evaporation
process not as easy as it seems. Using large-scale MD simulation
of nonequilibrium evaporation, we demonstrate below that
evaporated atoms are released from the interphase layer almost
without spending their kinetic energies, because they have a near-
zero binding energy at the interphase edge. The main work re-
quired for evaporation is provided by the bulk phase, which
supports via interatomic collisions a relatively slow drift of atoms
through the interphase by a temperature gradient. Because the
characteristic time of interatomic collisions in the condensed
phase is much shorter than the drift time through the interphase
(hundreds of picoseconds), the temperature remains in equilibrium
Fig. 5. Evolution of VDFs within the nonequilibrium part of the
interphase layer. Numbers indicate positions in the layer shown in
Fig. 4. The thin black line shows VDF for transverse velocity υy.The
last VDF at the edge of the interphase layer is marked by 1, while VDF
in vapor is marked by 0. Dashed line shows a model vapor VDF from
Eq. 4 with "b¼0.025kJ=mol. Major changes in VDF take place in the
interphase layer.
Fig. 6. VDFs in front of the absorbing boundary for different T
h
.A
decrease of vapor density with temperature results in a decreasing
rate of interatomic collisions resulting in reduction of a backward flux.
It also leads to lesser smearing of the VDF peak for small υx>0, which
leads to less changes of VDF in the vapor gap for smaller T
h
. The best
fit (red dashed line) of VDF at Th¼65.1K was obtained by Eq. 4 with
"b¼0.002kJ=mol, while a model VDF (black dashed line) normalized
to the same number density was built with "b¼6.65kJ=mol, which
corresponds to the cohesive energy of the atom in the solid argon at
T¼65.2K.
Zhakhovsky et al. PNAS Latest Articles
|
5of9
in the interphase until the density drops by one order of magnitude
near the interphase edge. Thus, VDF changes gradually with de-
creasing density inside the interphase from a symmetrical Max-
wellian form in the bulk of condensed phase to an almost semi-
Maxwellian VDF for evaporated atoms, which has a form described
by Eq. 4with the binding energy "b0.
Evaporation is always associated with condensation of evap-
orated atoms gaining a backward velocity due to interatomic
collisions in a vapor gap. Probability of such collisions increases
with the length of gap l
gap
. An additional flux j
ctoward the hot
surface is generated by evaporation from a cold surface facing the
hot surface. To eliminate the flux j
cfrom the opposite cold surface
and minimize the backward flux produced by collisions, we placed
a nonreflective absorbing boundary on the gap l
gap
of several
nanometers from the evaporation surface. In contrast to the large
vapor gap Lgap ¼365nm used in the preceding and next sections,
here the small vapor gaps lgap Lgap are used to achieve Kn 1
to reduce the probability of atom collisions in the gap. Such a
simple approach allows us to estimate an evaporation coefficient
from an intact evaporation flux undistorted by the atoms arriving
to the hot surface, and those may reflect back, without monitoring
of atom trajectories, which is hard to perform in the multimillion
atom simulations.
The absorbing boundary is implemented with the Maxwelldemon,
which watches over atom velocities in a gray zone beyond the
boundary at 1.5 nm from the interphase edge, as shown in Fig. 4.
If an atom gains a negative velocity υx, the Maxwell demon
subroutine replaced it by a small positive value. Atoms passing
the gray zone return back to the cold side of the film like in Fig. 1;
that is, Fig. 4 shows only part of the MD simulation domain to
provide a better spatial resolution for density and temperature
profiles in a thin interphase layer. Thus, a steady regime of evap-
oration, almost entirely uncoupled from condensation, is estab-
lished from the hot side of film.
The interphase layer can be divided into two almost equal
parts. First is the inner part between the bulk phase and a position
where the temperature profile splits into the longitudinal T
x
and
transverse T
y
temperatures, which is denoted by the left dashed
line in Fig. 4. The temperature equilibrium in the inner part of
interphase is well supported because the drift velocity is too small,
and atoms required more than 200ps to pass this part. It is rea-
sonable to take the position and temperature Th¼80.4K at the
point of temperature divergence as the surface/boundary pa-
rameters for the numerical solution of BKE.
With decreasing density by 3.5 times at the end of inner
equilibrium part, the temperature T
x
begins to drop, and the drift
velocity is accelerated to ux>0.005nm=ps, which is recognized as
a beginning of the outer nonequilibrium part of the interphase.
Acceleration of mass flow to about 0.1nm=ps to the edge of this
part shown by a right vertical line on Fig. 4 reduces a drift time of
atoms through the outer part to about 20 ps, which is comparable
with a collision time there. It leads to the large changes in VDF
shape with the approach to the right edge. Fig. 5 shows VDFs
constructed from velocity data accumulated in MD simulation
during about 1.2ns. VDF wings with negative velocities are re-
duced and shrunken much in approach to the right edge of the
interphase, while the width of positive VDF remains almost intact.
As a result of such evolution, the longitudinal T
x
drops dramati-
cally, but the averaged kinetic energy of evaporated atoms is little
affected, as it is illustrated in Fig. 4 by the longitudinal kinetic
energy Kx¼m
kBhυ2
xiexpressed in the temperature units. The trans-
verse Ky¼m
2kBhυ2
y+υ2
zTyby definition.
The VDF marked by 0 is accumulated in a thin layer with the
thickness of 0.5 nm at the end of vapor gap just before the ab-
sorbing boundary controlled by the Maxwell demon. Neverthe-
less, a negative velocity tail is formed here due to mostly pair
collisions. Such collision conserving the total energy and P
x
,P
y
,
and P
z
momenta may result in the negative velocity of one atom
along x, even though the colliding atoms have both positive ve-
locities υx. To get the negative atom velocity under such condi-
tions, the required kinetic energy is redistributed from the yand z
degrees of freedom of a colliding pair of atoms, which leads to a
decrease of T
y
seen in Fig. 4. Such collisions increase a backward
flux from the absorbing boundary to the interphase edge, but the
total flux remains constant everywhere. Thus, we see a larger
number of atoms having negative velocities in the interphase
edge VDF marked by 1 in Fig. 5.
The backward flux in vapor can be reduced by decreasing
the vapor density, which can be achieved with a decreasing
10 12 14 16 18 20 22 24
position x(nm)
30
40
50
60
70
80
temperatures T
x
and T
y
(K)
-5
-4
-3
-2
-1
0
potential energy E (kJ/mol)
0.1
1
10
0.2
0.5
2
5
20
0.05
density n (atoms/nm
3
)
0
20
40
60
80
100
120
140
flow velocity u
x
(m/s)
E
E
T
y
T
x
Maxwell demon
above 14 nm
u
x
n
Fig. 7. Profiles of number density, temperatures, flow velocity, and
averaged potential energy of atoms for different vapor gap lengths
but the fixed surface temperature Th¼80.4K. The absorbing
boundaries with the Maxwell demons are placed at 14, 16, 18, 20, 22,
and 24 nm in the MD simulation domain, which correspond to
lgap ¼1.5, 3.5, 5.5, 7.5, 9.5, and 11.5 nm, respectively.
lgap
α = j / jm
Fig. 8. Evaporation coefficient ~
α(lgap)¼j=jHK as a function of small
vapor gap length l
gap
. The fluxes are calculated from flow profiles
presented in Fig. 7. Error bars are from uncertainty in the vapor
density calculated from the red-line fit shown in Fig. 2.
6of9
|
www.pnas.org/cgi/doi/10.1073/pnas.1714503115 Zhakhovsky et al.
temperature of evaporation surface T
h
. The VDFs presented on
Fig. 6 were built for T
h
in the range of 65.1 to 80.4 K in the systems
similar to one shown on Fig. 4. The condensed phases were in
solid state at 65.1 and 70 K. The Maxwell demon is always placed
beyond the nonreflective absorbing boundary at the position of
14 nm, and the small vapor gaps l
gap
vary between 1.5 and 2 nm.
The collision rate drops for a larger Knudsen number at a lower
density in the vapor gap, resulting in a smaller negative wing of
VDF. It also prevents the large changes of VDF shape in the vapor
flow from the interphase to the absorbing boundary. The lesser
smearing of the VDF peak for small υx>0 for lower temperatures
produces a lesser impact on the almost semi-Maxwellian VDF for
atoms evaporated from the solid argon at Th¼65.1K. Such VDF
illustrates the idea of zero binding energy for evaporated atoms
even better than VDFs obtained for higher temperatures. In these
conditions, the simulated vapor flux j1.02jHK for Th¼65.1K is
very close to the j
HK
provided by the HertzKnudsen formula
jHK ¼nsffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
kBT=2m
p,[5]
which gives the exact flux of atoms in vapor in equilibrium with
a condensed phase, where a Maxwell VDF can be formally
divided by two semi-Maxwellian functions at υx¼0 for posi-
tive and negative fluxes, respectively.
Thus, the net evaporation coefficient defined here as
~
α¼j=jHK 1.02 is even slightly higher than unity, which we at-
tribute to uncertainties in the determination of surface temper-
ature and corresponding equilibrium vapor density estimated
from MD data profiles and the fit of vapor density presented
in Fig. 2.
We show above that the interatomic collisions produce a
backward flux in vapor, which increases with approach to the
evaporation surface. To study the effect of the vapor gap length
on vapor flow, we performed several MD simulations with differ-
ent positions of the absorbing boundary (see Fig. 7). The higher
longitudinal temperatures T
x
observed in vapor for larger gap
lengths are associated with the wider VDFs, which have larger
wings of negative velocities produced by collisions and trans-
ported by the backward flux to an observation point from outer
vapor layers. In contrast to this, the T
y
profiles for the different gap
lengths remain almost identical because there is no transverse flux
and the energy transfer from transverse degrees of freedom to
longitudinal ones depends on a local density, which does not
change much. Decreasing flow velocity u
x
with the gap length is
also explained by the widening of the negative wings of VDFs,
similarly to T
x
.
Fig. 7 demonstrates that the averaged potential energy of
atoms Edecreases about twice in the inner part of the in-
terphase and drops almost to zero at the end of the outer
nonequilibrium part of the interface layer. Thus, the atoms re-
leasing from the edge of interphase have a near-zero binding
energy, which leads to an almost semi-Maxwellian VDF of
evaporated atoms, as already demonstrated by Figs. 5 and 6.
Such semi-Maxwellian VDF with TThgives the flux jjHK,
which corresponds to α1. Thus, the suggested evaporation
mechanism with a near-zero binding energy yields the evapo-
ration coefficient close to unity.
The net evaporation coefficient calculated as a function of the
vapor gap length is shown in Fig. 8. It is defined as a ratio
~
αðlgapÞ¼j=jHK . The evaporation coefficient as a parameter αfor a
numerical solution of BKE can be obtained as a limit of ~
αðlgapÞα
at lgap 0, at which the backward flux vanishes. It can be esti-
mated as α0.97 ±0.01. However, this formal procedure may
have no physical meaning for an l
gap
smaller than the interaction
cutoff distance of 0.8125nm or the diameter 0.55nm of collision
cross-section. Also, it is worth noting that another definition of
temperature of evaporation surface can lead to the different
vapor density n
s
, which gives the different evaporation co-
efficient, as an example ~
α0.8 obtained in MD simulation of LJ
evaporation (25).
Because the vapor fluxes calculated for temperatures in the
range of 65 to 80 K are all close to j
HK
, and due to uncertainty in
the definition of the surface temperature and equilibrium vapor
density, we assume that the evaporation coefficient α¼1 for use
in the numerical solution of BKE.
-20 -19 -18 -17 -16 -15
position x(nm)
0.1
1
10
0.2
0.3
0.5
2
3
5
20
30
0.05
number density n (atoms/nm3)
70
72
74
76
78
80
82
84
temperatures Tx and Ty (K)
Tx
Ty
n
Tc
Fig. 9. Number density and temperatures profiles near a
condensation surface on the cold side of liquid film shown in Fig. 1.
The inner boundary of the nonequilibrium part of the interphase layer
(the right dashed line) is determined by a point of divergence
Tc¼72.4K between longitudinal T
x
and transverse T
y
temperatures.
The outer boundary corresponds to a point where the density profile
becomes flat. Circles indicate positions where velocity data for VDFs
are gathered during 1.5 ns. Evolution of instantaneous (not time-
averaged) profiles is presented in Movie S1.
Fig. 10. VDFs obtained in MD simulation and from BKE solutions with
the different condensation coefficients. VDFs are taken at the hot
and cold sides of liquid film presented in Figs. 1 and 9. While the VDFs
at the hot surface obtained from BKE are almost insensitive to choice
of β, the VDFs at the cold surface depend strongly on this choice.
Zhakhovsky et al. PNAS Latest Articles
|
7of9
Condensation Coefficient from EvaporationCondensation
Calculated by MD and BKE Methods
For the numerical solution of BKE in a vapor flow between the
evaporation and condensation surfaces, the positions and tem-
peratures of which are determined from MD simulation, the
boundary conditions have to be imposed on VDFs on these sur-
faces. Assuming that the hot surface with temperature T
h
is lo-
cated on the left end of the gap, the positive flux j+
hfrom the
surface is represented by a sum of evaporated and reflected atom
fluxes. It is supposed that an atom coming to the surface with a
negative flux j
may be absorbed with probability β, which de-
termines a condensation coefficient. Thus, the reflected atoms
form the reflected flux ð1βÞjjj, which together with the evap-
orated atom flux expressed as αjHK using Eq. 5results in the
positive flux j+
h¼αjHK +ð1βÞjjj. The required boundary VDF
must yield this flux.
As we demonstrated in Evaporation Coefficient from MD
Simulation, the VDF of evaporated atoms is represented by the
semi-Maxwellian VDF fHK ðThÞfor T
h
, which yields the Hertz
Knudsen flux. The VDF of reflected atoms can also be represented
by a similar semi-Maxwellian function fHK , because the reflection
is assumed to be diffuse with perfect accommodation to the
surface temperature. To produce the reflected flux, the VDF f
HK
must be normalized by the j
HK
and multiplied by the ð1βÞjjj.
Therefore, the left boundary condition applied for a positive wing
of VDF at the surface is given by
f+
h¼ðα+ð1βÞjjjjHK ÞfHK ðThÞ¼~
αfHK ðThÞ.[6]
The negative wing βfis eliminated on the left boundary.
Similarly, the right boundary condition imposed on the cold
surface at T
c
is given by
f
c¼ðα+ð1βÞjj+jjHK ÞfHK ðTcÞ,[7]
where we assume that the αis equal to that on the hot surface.
As indicated in Evaporation Coefficient from MD Simulation,
the evaporation coefficient is almost independent of temperature,
and α1inMDsimulationofevaporation;hence,α¼1isap-
plied on both boundaries in BKE calculations. The unknown con-
densation coefficient βis also taken to be independent of
surface temperature.
Definition of position and temperature of the cold surface is
similar to that for the hot surface introduced in the discussion of
Fig. 4. The density and temperature profiles with a point of di-
vergence between the longitudinal T
x
and transverse T
y
temper-
atures on the cold side of the film presented in Fig. 1 in the vicinity
of the interphase layer are shown in Fig. 9. Thus, the position and
temperature of the right boundary used in BKE calculations are
determined at the end of the equilibrium part of the cold in-
terphase layer formed in MD simulation of steady evaporation
condensation.
By varying the β, governing both boundary conditions in Eqs. 6
and 7, the VDFs obtained by MD and BKE methods can be
compared with the aim to find an optimal condensation co-
efficient leading to a good agreement between those VDFs. We
find numerical solutions of BKE with β¼0.8; 0.9; 1 for two evap-
orationcondensation systems with ThjTc¼80.4j72.4K in the va-
por gap Lgap ¼364.3nm and ThjTc¼79.4j60.5K in the vapor gap
Lgap ¼366.7nm. See details of simulations in SI Appendix. Fig. 10
shows the VDFs taken at hot and cold surfaces of liquid film for
β¼0.9; 1. The VDFs for β¼0.8 are not shown because they give
the largest deviation from the VDFs obtained in MD. One can see
that the VDFs from BKE are weakly dependent on βat the hot
surface, while BKE solution at the cold surface is very sensitive to
the variation of β.
Similar VDFs from BKE solutions of evaporationcondensation
between the liquid and solid sides of the same film at
ThjTc¼79.4j60.5K, respectively, are shown in Fig. 11. The cor-
responding flow profiles are presented in Fig. S2. The steady re-
gime of vapor flow in the corresponding MD simulation is
illustrated in Movies S1 and S2, where the instantaneous (not
time-averaged) profiles of density and temperatures across the
cold and hot interphase layers are shown, respectively. The den-
sity profile is almost static, and only temperatures in the low-dense
vapor experience the natural fluctuations in those videos. It is
readily seen that the point of temperature divergence appears at
very low number density within the interphase layer, which implies
the fast energy exchange between atoms there, as discussed
in Evaporation Coefficient from MD Simulation. Fig. 11 shows
again that the VDF on the hot surface is almost insensitive to the
choice of β.
Thus, βas a fitting parameter is determined primarily by the
VDF at the cold surface. Lower βproduces a larger vapor flux from
the condensation surface, which must be compensated by a
positive flux to the surface since the total flux in the steady
evaporationcondensation is fixed. As a result, the number den-
sity at the cold surface becomes higher than this in MD simulation.
The above comparison of VDFs from MD simulations and so-
lutions of BKE indicates that the best condensation coefficient is
β¼1 in the considered temperature range. The best agreement
between the flow profiles obtained in MD and BKE calculations is
also achieved at β¼1, as illustrated in Figs. S3 and S4.
SI Appendix provides the detailed description of the simula-
tion technique and comparison of MD and BKE temperature
profiles. Movies S1 and S2 show snapshots of density and tem-
perature profiles across the interphase layer on the condensation
side for TcjTh¼72.4j80.4K and on the evaporation side for
ThjTc¼79.4j60.5K, respectively.
Fig. 11. VDFs obtained in MD simulation and from BKE solutions with
the different condensation coefficients in evaporationcondensation
between the liquid and solid sides of the same film at ThjTc¼
79.4j60.5K, respectively. Lower βresults in larger negative flux from
the cold surface, which leads to larger deviation from VDF provided
by MD simulation.
8of9
|
www.pnas.org/cgi/doi/10.1073/pnas.1714503115 Zhakhovsky et al.
Conclusions
Consistent application of MD and BKE methods to the problem
of evaporationcondensation bridges the gap between the
atomistic representation of complex atom motion and probabi-
listic evolution of velocity distributions between two surfaces of
the condensed material. Using multimillion atom simulation, we
demonstrate that, contrary to intuition, atoms are released from a
condensed phase without the use of their kinetic energy but with
the support from collisions with other atoms in the interphase
layer. This effectively means that the evaporated atoms have a
near-zero binding energy, which leads to the virtually semi-
Maxwellian VDF of evaporated atoms and evaporation coeffi-
cient close to unity.
We also find that the best agreement between the steady flow
profile and VDFs obtained by MD and BKE methods is achieved if
both evaporation and condensation coefficients are close to unity
in the considered conditions.
We think that the evaporationcondensation of liquid metals,
water, and polyatomic molecular liquids should be studied next to
verify the applicability of our results to more complex materials.
Such studies may provide more appropriate boundary conditions
for use in continuum mechanics approaches to the evaporation
condensation problem.
Acknowledgments
This work was supported by Russian Foundation for Basic Research Grant 17-08-
00805. S.I.A. was supported by Russian Science Foundation Grant 14-19-01599.
1Kozyrev AV, Sitnikov AG (2001) Evaporation of a spherical droplet in a moderate-pressure gas. Phys Usp 44:725733.
2Maxwell JC (1890) The Scientific Papers of James Clerk Maxwell, ed Niven WD (Cambridge Univ Press, Cambridge, UK), Vol 2.
3Langmuir I (1915) The dissociation of hydrogen into atoms. [Part II] Calculation of the degree of dissociation and the heat of formation. J Am Chem Soc
37:417458.
4Hertz H (1882) Ueber die verdunstung der Flüssigkeiten, insbesondere des quecksilbers, im luftleeren raume. Ann Phys 253:177193.
5Knudsen M (1915) Die maximale verdampfungsgeschwindigkeit des quecksilbers. Ann Phys 352:697708.
6Kucherov RY, Rikenglaz LE (1960) On hydrodynamic boundary conditions for evaporation and condensation. Sov Phys JETP 37:8889.
7Labuntsov DA (1967) An analysis of the processes of evaporation and condensation. High Temp 5:579647.
8Muratova TM, Labuntsov DA (1969) Kinetic analysis of the processes of evaporation and condensation. High Temp 7:959967.
9Kogan MN, Makashev NK (1971) Role of the Knudsen layer in the theory of heterogeneous reactions and in flows with surface reactions. Fluid Dyn 6:913920.
10 Pao Y-P (1971) Application of kinetic theory to the problem of evaporation and condensation. Phys Fluids 14:306312.
11 Yen SM (1973) Numerical solutions of non-linear kinetic equations for a one-dimensional evaporation-condensation problem. Comput Fluids 1:367377.
12 Fischer J (1976) Distribution of pure vapor between two parallel plates under the influence of strong evaporation and condensation. Phys Fluids 19:13051311.
13 Aoki K, Cercignani C (1983) Evaporation and condensation on two parallel plates at finite Reynolds numbers. Phys Fluids 26:11631164.
14 Cercignani C, Fiszdon W, Frezzotti A (1985) The paradox of the inverted temperature profiles between an evaporating and a condensing surface. Phys Fluids
28:32373240.
15 Hermans LJF, Beenakker JJM (1986) The temperature paradox in the kinetic theory of evaporation. Phys Fluids 29:42314232.
16 Koffman LD, Plesset MS, Lees L (1984) Theory of evaporation and condensation. Phys Fluids 27:876880.
17 Anisimov SI (1968) Vaporization of metal absorbing laser radiation. Sov Phys JETP 27:182183.
18 Anisimov SI, Imas YA, Romanov GS, Khodyko YV (1971) Effects of High-Power Radiation on Metals (Joint Publications Research Service, Washington, DC).
19 Anisimov SI, Rakhmatulina AK (1973) The dynamics of the expansion of a vapor when evaporated into a vacuum. Sov Phys JETP 37:441444.
20 Tcheremissine FG (2006) Solution to the Boltzmann kinetic equation for high-speed flows. Comput Math Math Phys 46:315329.
21 Aristov VV (2001) Direct Methods for Solving the Boltzmann Equation and Study of Nonequilibrium Flows. Fluid Mechanics and its Applications, ed Moreau R
(Springer, Dordrecht, The Netherlands), Vol 60.
22 Kryukov A, Levashov V, Shishkova I (2001) Numerical analysis of strong evaporationcondensation through the porous matter. Int J Heat Mass Transfer
44:41194125.
23 Kryukov AP, et al. (2006) Selective water vapor cryopumping through argon. J Vac Sci Technol A 24:15921596.
24 Kryukov A, Levashov V, Shishkova I (2009) Evaporation in mixture of vapor and gas mixture. Int J Heat Mass Transfer 52:55855590.
25 Zhakhovskii V, Anisimov S (1997) Molecular-dynamics simulation of evaporation of a liquid. J Exp Theor Phys 84:734745.
26 Marek R, Straub J (2001) Analysis of the evaporation coefficient and the condensation coefficient of water. Int J Heat Mass Transfer 44:3953.
27 Ishiyama T, Fujikawa S, Kurz T, Lauterborn W (2013) Nonequilibrium kinetic boundary condition at the vapor-liquid interface of argon. Phys Rev E 88:042406.
28 Kon M, Kobayashi K, Watanabe M (2016) Liquid temperature dependence of kinetic boundary condition at vaporliquid interface. Int J Heat Mass Transfer
99:317326.
29 Meland R, Ytrehus T (2007) Molecular dynamics simulation of evaporation of two-component liquid. Proceedings of 25th International Symposium on Rarefied
Gas Dynamics, eds Ivanov MS, Rebrov AK (Publ. House of the Siberian Branch of the Russian Acad. of Sciences, Novosibirsk), pp 12291232.
30 Yang T, Pan C (2005) Molecular dynamics simulation of a thin water layer evaporation and evaporation coefficient. Int J Heat Mass Transfer 48:35163526.
31 Cheng S, Lechman JB, Plimpton SJ, Grest GS (2011) Evaporation of Lennard-Jones fluids. J Chem Phys 134:224704.
32 Kon M, Kobayashi K, Watanabe M (2014) Method of determining kinetic boundary conditions in net evaporation/condensation. Phys Fluids 26:072003.
33 Kon M, Kobayashi K, Watanabe M (2017) Kinetic boundary condition in vapor-liquid two-phase system during unsteady net evaporation/condensation. Eur J
Mech B Fluids 64:8192.
34 Kobayashi K, Sasaki K, Kon M, Fujii H, Watanabe M (2017) Kinetic boundary conditions for vaporgas binary mixture. Microfluid Nanofluid 21:53.
35 Kobayashi K, Hori K, Kon M, Sasaki K, Watanabe M (2016) Molecular dynamics study on evaporation and reflection of monatomic molecules to construct kinetic
boundary condition in vaporliquid equilibria. Heat Mass Transfer 52:18511859.
36 Kogan MN (1969) Rarefield Gas Dynamics (Plenum, New York).
37 Heerman DW (1986) Computer Simulation Methods in Theoretical Physics (Springer, Berlin).
38 Haynes WM, ed (2017) CRC Handbook of Chemistry and Physics (CRC Press, Boca Raton, FL), 97th Ed, pp 6-128 and 15-10.
39 Zhakhovskii VV, Zybin SV, Nishihara K, Anisimov SI (1999) Shock wave structure in Lennard-Jones crystal via molecular dynamics. Phys Rev Lett 83:11751178.
40 Shishkova IN, Sazhin SS (2006) A numerical algorithm for kinetic modelling of evaporation processes. J Comput Phys 218:635653.
41 Gerasimov DN, Yurin EI (2014) The atom velocity distribution function in the process of liquid evaporation. High Temp 52:366373.
Zhakhovsky et al. PNAS Latest Articles
|
9of9
... Rayleigh-Taylor instability and Rayleigh-Taylor mixing are a subject of active research, on the side of fundamentals and in the application problems . Rayleigh-Taylor instability and Rayleigh-Taylor mixing govern a wide range of processes in nature and technology and are particularly important in high-energy density plasmas [5][6][7][8][9][30][31][32][33][34][35][36][37][38][39][40][41][42][43]. Examples comprise the abundance of chemical elements in supernova remnants, the coronal mass ejections in the solar flares, the formation of hot spots in inertial confinement fusion, and the efficiency of plasma thrusters [5][6][7][8][30][31][32][33][34][35][36][37][38][39][40][41][42][43]. ...
... Rayleigh-Taylor instability and Rayleigh-Taylor mixing govern a wide range of processes in nature and technology and are particularly important in high-energy density plasmas [5][6][7][8][9][30][31][32][33][34][35][36][37][38][39][40][41][42][43]. Examples comprise the abundance of chemical elements in supernova remnants, the coronal mass ejections in the solar flares, the formation of hot spots in inertial confinement fusion, and the efficiency of plasma thrusters [5][6][7][8][30][31][32][33][34][35][36][37][38][39][40][41][42][43]. In this work, by exploring data on fluctuations spectra obtained in the state-of-the-art fine-resolution experiments [9], we find the characteristics of Rayleigh-Taylor mixing and the physics properties of matter in high-energy density plasmas, including the data-based value of the kinematic viscosity in high-energy density plasmas. ...
... The properties of matter in high-energy density plasma are a subject to active research [8,[34][35][36][41][42][43][44][45][46][47][48][49]. The transport coefficients are a long-standing puzzle [27][28][29]. ...
Article
Full-text available
We explore properties of matter and characteristics of Rayleigh–Taylor mixing by analyzing data gathered in the state-of-the-art fine-resolution experiments in high-energy density plasmas. The eminent quality data represent fluctuations spectra of the X-ray imagery intensity versus spatial frequency. We find, by using the rigorous statistical method, that the fluctuations spectra are accurately captured by a compound function, being a product of a power law and an exponential and describing, respectively, self-similar and scale-dependent spectral parts. From the self-similar part, we find that Rayleigh–Taylor mixing has steep spectra and strong correlations. From the scale-dependent part, we derive the first data-based value of the kinematic viscosity in high-energy density plasmas. Our results explain the experiments, agree with the group theory and other experiments, and carve the path for better understanding Rayleigh–Taylor mixing in nature and technology.
... 45,85 Numerical simulations thoroughly quantified Rayleigh-Taylor/ Richtmyer-Meshkov unstable flows, including the dynamics of the bulk and the structures of the interface. 16,[38][39][40][41][42][46][47][48]90 They employed the Lagrangian methods (e.g., molecular dynamics, smoothed particle hydrodynamics) and the Eulerian methods (e.g., front tracking, level set, large eddy simulations, direct numerical simulations). Large-scale computations modeled the effect of turbulence (presuming it develops) on diagnostics parameters of the mixing zone in Rayleigh-Taylor/ Richtmyer-Meshkov flows. ...
... Large-scale computations modeled the effect of turbulence (presuming it develops) on diagnostics parameters of the mixing zone in Rayleigh-Taylor/ Richtmyer-Meshkov flows. 16,[38][39][40][41][42][46][47][48]90 In theory, remarkable success was achieved in understanding Rayleigh-Taylor/Richtmyer-Meshkov dynamics. [3][4][5][7][8][9][10]27,28,[33][34][35][36][37][56][57][58][68][69][70][71][72][86][87][88][89][90] This includes the linear and weakly nonlinear theories focusing on the early time evolution, [3][4][5][86][87][88][89] the empirical models of the late time mixing process, 27,28,[56][57][58] and the group theory approached the problem of unstable dynamics from the first principles. ...
... Potential applications of our work include a broad range of processes, to which Rayleigh-Taylor/Richtmyer-Meshkov mixing and canonical turbulence are relevant, and a development of universal theoretical models of realistic complexity by using the group theory and representations theory methods. [7][8][9][10][11][12][13][14][15][16][17][18][19][20][21][22][23][24][25][26][33][34][35][36][37][62][63][64][65][66][67] The novelty of our work is in the direct link between the properties of Rayleigh-Taylor/Richtmyer-Meshkov mixing with variable acceleration and the properties of canonical Kolmogorov turbulence, by using the group theory and by taking into account the invariant forms of these processes (see Secs. II A-II D). ...
Article
Full-text available
Canonical turbulence and Rayleigh–Taylor/Richtmyer–Meshkov mixing with variable acceleration are paradigmatic complexities in science, mathematics, and engineering, with broadly ranging applications in nature, technology, and industry. We employ scaling symmetries and invariant forms to represent these challenging processes and to assess their very different properties. We directly link—for the first time to our knowledge—the attributes of Rayleigh–Taylor/Richtmyer–Meshkov interfacial mixing with variable acceleration to those of canonical turbulence, including scaling laws, spectral shapes, and characteristic scales. We explore the role of control dimensional parameters in quantifying these processes. The theory results compare well with available observations, the chart perspectives for future experiments and simulations, and for better understanding realistic complexity.
... Along their phase transition trajectories, molecules experience changes in their interaction with the liquid potential energy, creating an energy barrier that must be overcome to transition between phases. 40 This energy barrier introduces asymmetry into the evaporation and condensation processes as it decelerates evaporating molecules, reducing their probability of breaking free of the liquid, while accelerating condensing molecules as they approach the liquid. In this work, the average change in interaction with the liquid potential energy along the transition trajectories was estimated to be DE p;l =k B ¼ 6:29 K. Thus, evaporation and condensation coefficients should not be assumed equal, even under equilibrium conditions, as is often done for simplicity in phase-change rate theories or other MD simulations. ...
Article
Full-text available
The understanding of the liquid–vapor interface is of great importance in various fields of science and technology; however, it remains an unresolved issue from a microscopic perspective. In this paper, we propose a new approach to defining the liquid–vapor interface, enabling the tracking of phase-transitioning molecules as they travel from the densely packed liquid phase to the freely moving vapor and vice versa. This approach was applied to study evaporating, condensing, and reflecting molecules in molecular dynamics simulations of argon liquid–vapor equilibrium at a temperature of 90 K. The results showed that evaporation positions are distributed over a wide range of surface-normal coordinates due to the non-flat and non-stationary nature of the liquid-phase surface. Additionally, the evaporation coefficient was found to be slightly lower than the condensation coefficient, indicating that these processes are not symmetrical due to the energy barrier at the interface, even under equilibrium conditions. Furthermore, both evaporation and condensation probabilities were observed to increase with the surface-normal velocity component prior to the event. However, evaporation probability tended to decrease as the bonding energy between evaporating molecules and the liquid-phase molecules increased at the beginning of evaporation trajectory. The analysis of the absolute velocity distributions revealed that the velocity distribution along the condensation trajectory changes from Maxwellian distribution to accelerated Maxwellian distribution due to the energy barrier at the interface. On the other hand, the evaporating molecules start their trajectories with the accelerated Maxwellian distribution, which is decelerated to the Maxwellian distribution before the molecules escape the interface.
... Researches in this direction are carried out using both computational, theoretical, and experimental methods. At present, theoretical investigations of the evaporation processe are performed using the methods of continuum mechanics, molecular-kinetic theory, and molecular-dynamic modeling [11][12][13][14][15]. The problem of material evaporation from a surface with a given temperature was considered in [16] using the Monte Carlo direct numerical simulation and the solution of Euler equations for the discontinuity breakdown problem. ...
... As pointed out by Polikarpov et al. (2019), since the Knudsen layer typically extends over a length 1-2 mean-free paths, these jumps can only be theoretically predicted because they occur on a spatial scale that cannot be currently resolved by measurement techniques. On this account, the application of the kinetic theory of gases is essential to examine the vapor behavior near the interface either by performing molecular dynamics simulations Zhakhovsky et al., 2018) or by direct solution of the Enskog-Vlasov equation using the direct simulation Monte Carlo (DSMC) method (Kon et al., 2014(Kon et al., , 2016. These methods have the merit to disclose many details of the molecular gas dynamics, such as the fraction of molecules emitted, reflected, and impinging onto the droplet surface. ...
Article
The dynamics of near-critical single droplets allow to investigate the transition from two-phase to single-phase mixing under well-defined conditions, devoid of the additional complications due to drop-drop interactions and combustion. Recently, an empirical regime map was proposed to predict the evolution of microscopic transcritical droplets. The experiments show that classical evaporation remains the controlling mechanism over a wide range of supercritical ambient pressures and temperatures with respect to the critical point of the evaporating fluid. Moreover, the onset ambient pressure for the transition to single-phase mixing varies inversely with temperature. To explain this trend, the behavior of a single droplet at near-critical conditions is investigated theoretically by means of a Langmuir-type evaporation model, originally proposed by Young. The model incorporates a modified boundary condition due to the inclusion of gas kinetic effects close to the vapor−liquid interface. This advanced evaporation model is employed to reproduce analytically the above-mentioned regime map, showing a good agreement with experimental findings. The analysis also revealed that the onset of the single-phase mixing regime is associated with the quenching of the evaporation process. The latter is caused by the decrease of the evaporation coefficient, which control the mass-transfer rate across the Knudsen layer. The resulting reduction in evaporative cooling leads to the rapid heating of the liquid droplet and to the disintegration of the material interface at the critical temperature.
... The comparison of the calculation results with the experimental data is shown in Figs. 3 and 4. As can be seen from the comparison results the model of condensation process used in this paper can quite accurately predict the position of the condensation zone along the nozzle, temperature changes and distribution of vapour concentration, as well as the average radius of the formed droplets. It should be noted the satisfactory agreement of our calculation results with experimental data [17,60] was obtained at value of condensation coefficient equal to 1 which corresponds to the recommendation [61]. ...
Article
Phase transition usually consumes or releases energy to produce cooling or heating within different materials, providing a generalized framework for temperature regulation in practical applications. Because of the strong coupling between the enthalpy change in thermodynamics and heat-mass transfer kinetics, unveiling the mechanism of temperature regulation via the phase transition remains a great challenge. Here, we develop a new theoretical method by establishing a connection of enthalpy change from thermodynamics to phase transition dynamics to study evaporation-induced cooling as an example. Our new approach can spontaneously generate evaporative cooling at interfaces, and the predicted results are consistent with recent experiments. The evaporation-induced steady vapor is dictated by an anomalous cold-to-hot mass transfer through temperature-dependent chemical potentials, which enables temperature regulation inside liquids via a thermodynamic-kinetic interplay. Moreover, we show that a simple prohibition of heat exchange between liquids and reservoir can greatly enhance the cooling magnitude by a factor of 2∼4, which is highly dependent on the thermodynamics and kinetic coefficients of liquids. Our new method paves the way for exploration of cooling or heating induced by different phase transitions, such as evaporation, sublimation, or condensation, in a unified framework, which can significantly promote the development of temperature regulation by phase transitions.
Article
Full-text available
Using molecular dynamics simulations, the present study investigated the precise characteristics of the binary mixture of condensable gas (vapor) and non-condensable gas (NC gas) molecules creating kinetic boundary conditions (KBCs) at a gas–liquid interface in equilibrium. We counted the molecules utilizing the improved two-boundary method proposed in previous studies by Kobayashi et al. (Heat Mass Trans 52:1851–1859, 2016. doi:10.1007/s00231-015-1700-6). In this study, we employed Ar for the vapor molecules, and Ne for the NC gas molecules. The present method allowed us to count easily the evaporating, condensing, degassing, dissolving, and reflecting molecules in order to investigate the detailed motion of the molecules, and also to evaluate the velocity distribution function of the KBCs at the interface. Our results showed that the evaporation and condensation coefficients for vapor and NC gas molecules decrease with the increase in the molar fraction of the NC gas molecules in the liquid. We also found that the KBCs can be specified as a function of the molar fraction and liquid temperature. Furthermore, we discussed the method to construct the KBCs of vapor and NC gas molecules.
Article
Full-text available
Using molecular dynamics simulations, the present study investigates the precise characteristics of evaporating and reflecting monatomic molecules (argon) composing a kinetic boundary condition (KBC) in a vapor–liquid equilibria. We counted the evaporating and reflecting molecules utilizing two boundaries (vapor and liquid boundaries) proposed by the previous studies (Meland et al. in Phys Fluids 16:223–243, 2004; Gu et al. in Fluid Phase Equilib 297:77–89, 2010). In the present study, we improved the method using the two boundaries incorporating the concept of the spontaneously evaporating molecular mass flux. The present method allows us to count the evaporating and reflecting molecules easily, to investigate the detail motion of the evaporating and reflecting molecules, and also to evaluate the velocity distribution function of the KBC at the vapor–liquid interface, appropriately. From the results, we confirm that the evaporating and reflecting molecules in the normal direction to the interface have slightly faster and significantly slower average velocities than that of the Maxwell distribution at the liquid temperature, respectively. Also, the stall time of the reflecting molecules at the interphase that is the region in the vicinity of the vapor–liquid interface is much shorter than those of the evaporating molecules. Furthermore, we discuss our method for constructing the KBC that incorporates condensation and evaporation coefficients. Based on these results, we suggest that the proposed method is appropriate for investigating KBC in various nonequilibrium states or multi-component systems.
Article
Full-text available
The aim of the present study is to develop the method of determining the kinetic boundary condition (KBC) at a vapor-liquid interface in net evaporation/condensation. We proposed a novel method for determining the KBC by combining the numerical simulations of the mean field kinetic theory and the molecular gas dynamics. The method was evaluated on steady vapor flow between two liquid slabs at different temperatures. A uniform net mass flux in the vapor phase induced by net evaporation and condensation is obtained from the numerical simulation of the mean field kinetic theory for both vapor and liquid phases. The KBC was specified by using the uniform net mass flux, and the numerical simulation of the molecular gas dynamics was conducted for the vapor phase. Comparing the macroscopic variables in the vapor phase obtained from both numerical simulations, we can validate the KBC whether the appropriate solutions are obtained. Moreover, the evaporation and condensation coefficients were estimated uniquely. The results showed that the condensation and evaporation coefficients were identical and constant in net evaporation. On the other hand, in net condensation, the condensation coefficient increased with the collision molecular mass flux. We also presented the applicable limit of the KBC which is assumed to be the isotropic Gaussian distribution at the liquid temperature. From these results, the KBCs in net evaporation and condensation, which enable the exact macroscopic variables to be determined, were proposed.
Article
Heat and mass transfer caused by nonequilibrium phase change (net evaporation/condensation) plays an important role in a vapor–liquid two-phase flow. In general, liquid temperature changes with time because of the heat and mass transfer between the vapor and liquid phases; however, a precise investigation of the transport phenomena related to this temporal evolution of liquid temperature is still lacking. The aim of this study is to examine a kinetic boundary condition, which depends on liquid temperature, for the Boltzmann equation in a vapor–liquid two-phase system with unsteady net evaporation/condensation. In this study, we confirmed whether the kinetic boundary condition follows the temporal evolution of liquid temperature attributed to unsteady net evaporation/condensation by using the molecular simulation based on mean-field kinetic theory, and then we validated the accuracy of the kinetic boundary condition by solving the initial boundary value problem of the Boltzmann equation in unsteady net evaporation/condensation. These results showed that the kinetic boundary condition follows the temporal evolution of liquid temperature in the simulation setting of this study. Furthermore, we concluded that the kinetic boundary condition that depends on liquid temperature is guaranteed to be accurate even in unsteady net evaporation/condensation by considering the temporal evolution of liquid temperature.
Article
For the accurate description of heat and mass transfer through a vapor-liquid interface, the appropriate modeling of the interface during nonequilibrium phase change (net evaporation/condensation) is a crucial issue. The aim of this study is to propose a microscopic interfacial model which should be imposed at the interface as the kinetic boundary condition for the Boltzmann equation. In this study, we constructed the kinetic boundary condition for monoatomic molecules over a wide range of liquid temperature based on mean field kinetic theory, and we validated the accuracy of the constructed kinetic boundary condition by solving the boundary value problem of the Boltzmann equation. These results showed that we can impose the kinetic boundary condition at the interface by simply specifying liquid temperature and simulate the complex vapor-liquid two-phase flow induced by net evaporation/condensation. Furthermore, we applied the constructed kinetic boundary condition to the boundary condition for the fluid-dynamic-type equations. This application enables us to deal with a large spatio-temporal scale of the interfacial dynamics in the vapor-liquid two-phase system with net evaporation/condensation.
Article
Analytical expressions determining the velocity distribution function of evaporating particles were obtained for evaporation into vacuum. The formulas are compared with numerical simulation results; the coincidence can be treated as sufficiently good. The particle distribution function over the longitudinal, i.e., directed from evaporation surface, velocity turns out to be significantly non-Maxwellian. The distribution function over the transverse projections of the velocity is Maxwellian but with a distribution modulus different from the evaporating liquid temperature T. The mean energy of a particle escaping from the surface is 2kT.
Article
The evaporation and condensation coefficients of water are extensively analyzed considering also data hitherto not taken into account. From the performed evaluation, a decline of both coefficients with increasing temperature and pressure is derived. For water, the condensation coefficients is generally higher than the evaporation coefficient. Evaporation and condensation coefficients exceed 0.1 for dynamically renewing water surfaces, while the analysis reveals coefficients below 0.1 for stagnant surfaces. The influence of impurities and surface active substances, as well as the effect of the dynamic surface tension is discussed.