You are on page 1of 8

International Journal of Heat and Mass Transfer 136 (2019) 449–456

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Kinetic simulation of the non-equilibrium effects at the liquid-vapor


interface
A.Ph. Polikarpov a,⇑, I.A. Graur b, E.Ya. Gatapova c, O.A. Kabov c
a
Ural Federal University, 51 str. Lenina, 620000 Ekaterinbourg, Russia
b
Aix-Marseille Université, CNRS, IUSTI UMR 7343, 13013 Marseille, France
c
Kutateladze Institute of Thermophysics SB RAS, Novosibirsk 630090, Russia

a r t i c l e i n f o a b s t r a c t

Article history: Phase change phenomena at microscale is important for novel cooling microsystems with intensive evap-
Received 22 November 2018 oration, so the development of reliable models and simulations are challenging. The vapor behaviors near
Received in revised form 26 February 2019 its condensed phase are simulated using the non-linear S-model kinetic equation. The pressure and tem-
Accepted 28 February 2019
perature jumps obtained numerically are in good agreement with the analytical expressions derived from
the appropriate Onsager-Casimir reciprocity relations. The results of the evaporation flux are close to
those given by the Hertz-Knudsen-Schrage formula, only when the values of the pressure and tempera-
Keywords:
ture at the upper boundary of the Knudsen layer are used. Comparison with recently measured temper-
Liquid-vapor interface
Evaporation rate
ature jumps are provided and disagreement with some experiments are discussed.
Knudsen layer Ó 2019 Elsevier Ltd. All rights reserved.
Molecular mean free path
Non-equilibrium state
Temperature and pressure jumps

1. Introduction the kinetic theory of gases. Only recently, the new series of mea-
surements [5,6] have appeared, where the temperature jump was
Understanding of heat and mass transfer mechanisms at liquid- found of the same order as that predicted by the kinetic theory.
vapor interface is important not only from the fundamental point However, still in recent papers [6,7] the temperature in vapor near
of view, but also for various applications, such as for the design interface was measured higher compared to the interface temper-
and optimization of the cooling microsystems. During the evapora- ature. The positive values of the temperature difference between
tion process a thin layer, the Knudsen layer, forms near the liquid liquid and vapor temperatures at interface (vapor temperature is
interface at the vapor side. Inside this layer, which thickness is of lower than the interface temperature) were measured only by
the order of several mean free paths, the vapor is in equilibrium the authors of Ref. [5].
state only when the flux of the evaporation molecules is equal to To go forward in the understanding of the flow behavior at
the flux of the condensed molecules. When a net evaporation (or liquid-vapor interface the gas flow evaporating from its condensed
condensation) flux exists a vapor near the interface is in non- phase is investigated on the basis of the kinetic approach. The non-
equilibrium state and the continuity of the thermodynamic vari- linear S-model kinetic equation [8] is solved numerically by the
ables, like pressure and temperature, cannot be ensured anymore. Discrete Velocity Method (DVM) [9]. The structure of the Knudsen
This non-equilibrium behavior of a vapor cannot be described by layer is analyzed and the macroscopic temperature and pressure
the continuum equations and other approaches, as the gas kinetic jumps, obtained from the numerical simulations, are compared
theory and molecular dynamics have to be implemented. with the analytical expressions derived by the authors of Ref.
From a two decades different authors [1–4] have measured the [10] from the kinetic theory of gases and the thermodynamics of
liquid and vapor properties namely the temperature profiles and irreversible processes. The experimental data of Refs. [4,6] are used
the temperature jump at the liquid-vapor interfaces. In all these as input parameters for the numerical analysis.
experiments the temperature jump measured on the interface
was found surprisingly large, much larger than that predicted by 2. Problem statement

We consider a plane condensed phase at rest occupying the half


⇑ Corresponding author. space (y0 < 0), and the gas (vapor) evaporating from this infinite
E-mail address: alexey.polikarpov@gmail.com (A.Ph. Polikarpov). planar surface kept at constant and uniform temperature T s . The

https://doi.org/10.1016/j.ijheatmasstransfer.2019.02.100
0017-9310/Ó 2019 Elsevier Ltd. All rights reserved.
450 A.Ph. Polikarpov et al. / International Journal of Heat and Mass Transfer 136 (2019) 449–456

R R
interface is located at y0 ¼ 0, where y0 is the variable normal to the n0 ðy0 Þ ¼ f ðy0 ; v Þdv ; u0y ðy0 Þ ¼ n10 f ðy0 ; v Þv y dv ;
R R ð6Þ
condensed phase surface, see Fig. 1. The steady one-dimensional
T 0 ðy0 Þ ¼ 3kmB n0 f ðy0 ; v ÞV 2 dv ; q0y ðy0 Þ ¼ m2 f ðy0 ; v ÞV 2 ðv y  u0y Þdv :
flow is considered.
When a gas is near a surface (liquid or solid) a thin layer, the
Knudsen layer, forms in the vicinity of the surface. The thickness
of this layer is usually of the order of several molecular mean free The evaporation flow rate, expressed in the number of
paths. To estimate the thickness of this layer we use the equivalent molecules evaporating per seconds and per unit area, J 0n , and the
mean free path defined as [11]: evaporation mass flow rate, expressed in mass of vapor (in
kilogram) per second evaporating from a unit area, J 0m , are defined
ls v s
‘s ¼ ; ð1Þ as following:
ps Z Z
using the reference parameters with subscript s, corresponding to J 0n ¼ v y f ðy0 ; v Þdv; J 0m ¼ m v y f ðy0 ; v Þdv : ð7Þ
the vapor characteristics at the condensed phase surface. In Eq.
(1) ls ¼ lðT s Þ is the dynamic viscosity of the vapor phase The second definition of the evaporation mass flow rate is
sffiffiffiffiffi usually provided from the experiments. It is clear that previous
T0 relations represent the conservation of the number of particles
lðT Þ ¼ ls
0
ð2Þ and the mass conservation. Additionally, the y momentum and
Ts
energy conservation are written as
at temperature T s ; v s is the most probable molecular speed, Z Z
pffiffiffiffiffiffiffiffiffiffiffiffi J 0v y ¼ v 2y f ðy0 ; vÞdv ; J 0E ¼ m v y v 2 f ðy0 ; v Þdv: ð8Þ
v ðT 0 Þ ¼ 2RT 0 ; ð3Þ
The constancy of J 0m ; J 0v y , and J 0E will be used for the accuracy test
calculated also at the temperature T s : v s ¼ v ðT s Þ; R ¼ kB =m is the
specific gas constant, kB is the Boltzmann constant, m is the molec- of the applied numerical method.
ular mass.
The upper boundary of the computational domain is far from 4. Boundary conditions
the evaporation surface, at the distance H, see Fig. 1. Different val-
ues of H are tested and finally H ¼ 25‘s is retained to do all the The distribution function of evaporating molecules is assumed
simulations. to be a half-range Maxwellian:
 3=2
m  
3. S-model kinetic equation f ðt;0; v Þ ¼ ðrns þ ð1  rÞnr Þ exp mv 2 =ð2kB T s Þ ; v y > 0;
2pkB T s
ð9Þ
To model the evaporation process of a monoatomic gas from its
condensed phase the S-model kinetic equation [8] is used. The were
evaporation phenomenon is considered here as one dimensional sffiffiffiffiffiffiffiffiffiffiffi Z
2pm
in physical space, so the S-model kinetic equation is written as
nr ¼  f v dv : ð10Þ
  kB T s v y <0 y
@f @f
0 þ vy ¼ t0 f  f ;
S
ð4Þ
@t @y 0 Here ns is the number density, calculated from the saturated
surface temperature and pressure as ns ¼ ps =ðkB T s Þ. The coefficient
where f ðt ; y ; v Þ is the one particle velocity distribution function, t
0 0 0
r that is a part of the incident molecules evaporating immediately
is the time, v ¼ ðv x ; v y ; v z Þ is the molecular velocity vector, t0 is the
from the condensed surface, while ð1  rÞ part of molecules is
collision frequency, t0 ¼ p0 =l0 ; p0 is the gas pressure. In the frame of
assumed to be reflected diffusively from the interface.
S
the S-model the equilibrium distribution function f in Eq. (4) is The uniform equilibrium vapor state (subscript 1) is described
defined as following by the equilibrium Maxwellian distribution function
" !#  3=2
2mVq0 mV2 5 p m  
f ðt ; y ; v Þ ¼ f f ðt; H; v Þ ¼ 1 exp mðv  u1 Þ2 =ð2kB T 1 Þ ; v y < 0;
S 0 0 M
1þ  ; ð5Þ
15n0 ðy0 ÞðkB T 0 ðy0 ÞÞ
2
2kB T 0 ðy0 Þ 2 kB T 1 2pkB T 1
ð11Þ
here T 0 ðy0 Þ is a gas temperature, n0 ðy0 Þ is a gas number density,
u0 ¼ ð0; u0y ; 0Þ is a bulk velocity vector, V ¼ v  u0 is the peculiar where u1 ¼ ð0; uy1 ; 0Þ. As it was discussed in Refs. [13–15], in the
M case of evaporation, a solution of the boundary value problem exists
velocity vector, q0 ¼ ð0; q0y ; 0Þ is a heat flux vector, f is the Maxwel-
only when some relations between the parameters are satisfied. In
lian distribution function [12]. The macroscopic parameters are the case of evaporation these relations are given by [14]
defined as follows:
p1 T1
¼ h1 ðMay1 Þ; ¼ h2 ðMay1 Þ: ð12Þ
ps Ts

The functions h1 and h2 are obtained numerically and their tab-


ulated values can be found in Ref. [14]. In the case of weak evapo-
ration conditions, that means that the variation from the uniform
equilibrium state at reste with pressure and temperature is small
(or where the evaporation is weak, i.e. evaporation Mach number
is small compared to 1), the relations between three parameters
become
p1 uy1 T1  uy1
¼ 1 þ C 4 pffiffiffiffiffiffiffiffiffiffiffiffi ; ¼ 1 þ d4 pffiffiffiffiffiffiffiffiffiffiffiffi ; ð13Þ
ps 2RT s Ts 2RT s
Fig. 1. Problem configuration.
A.Ph. Polikarpov et al. / International Journal of Heat and Mass Transfer 136 (2019) 449–456 451


where C 4 ¼ 2:13204 and d4 ¼ 0:44675, obtained with p1  ps T1  Ts
XP ¼ ; XT ¼ : ð19Þ
Boltzmann-Krook-Welander (BKW) model in Ref. [16]. The previous ps Ts
relations give the boundary conditions for the Euler equations.
We assume then that the deviations between the temperature
However, for the Navier-Stokes equations more complete boundary
of the condensed surface and that far from it and the correspond-
conditions have to be used on the liquid-vapor interface, which are
ing pressures are small: X P  1 and X T  1. For a given gas the
discussed in Section 6.
pressure and temperature differences are coupled by the relation
The number and the nature of conditions (12) are different for
evaporation and condensation flows [14,15]. ps  p1 ¼ bðT s  T 1 Þ; ð20Þ
where b is a positive constant corresponding to the slop of the Clau-
5. Dimensionless form sius–Clapeyron curve at T s , so X P and X T are not independent quan-
tities. However, here we will consider two forces separately, to see
For further derivation we introduce the following dimension- clearly the impact of each force on the evaporation process.
less quantities: Following [19] we introduce the ‘‘fluxes” corresponding to the
y0 v u0 vs n0 T0 q0 driving ‘‘forces” as:
y¼ ; c ¼ ; u ¼ ; t ¼ t0 ; n ¼ ; T ¼ ; q ¼ :
‘s vs vs ‘s ns Ts ps v s
J 0P ¼ ns u0y ; J 0T ¼ 
1 0
q; ð21Þ
ð14Þ kB T s y

Now the dimensionless S-model kinetic equation can be written where u0y and q0y do not depend on y. The thermodynamic fluxes are
in the form: related to the thermodynamic forces in the matrix form:
" # " #

@f @f pffiffiffi S  J 0P K0PP K0PT XP


þ vy ¼n T f f : ð15Þ ¼  : ð22Þ
@t @y J 0T K0TP K0TT XT
The dimensionless boundary conditions for the distribution
The Onsager-Casimir relation K0PT ¼ K0TP in this case yields the
function of the reflected molecules at the liquid-vapor interface
coupling between the mass flux caused by temperature drop and
can be written as
the thermal flux caused by the pressure drop [19].
y ¼ 0; t > 0; cy > 0; Previous equation allows to express the thermodynamic fluxes
  ð16Þ in function of the thermodynamic forces. In this way the expres-
f ðt; 0; cÞ ¼ ðrns þ ð1  rÞnr Þf s ;
M M
f s ¼ p3=2
1
exp c2 :
sions analogous to the Hertz-Knudsen equation are obtained in
the end of this Section. However, first we are interested to express
the thermodynamic forces in function of fluxes, so we can write
The number density nr can be calculated from the impermeabil-

" #1 " #
" #
ity condition on the condensed surface: XP K0PP K0PT J 0P a011 a012 J 0P
Z ¼  ¼  : ð23Þ
pffiffiffiffi XT K0TP K0TT J 0T a021 a022 J 0T
nr ¼ 2 p cy f dc: ð17Þ
The elements a0ij ¼ aij =ðns v s Þ in previous relation are obtained in
cy <0

As it was mentioned in previous section far from the condensed [10,20], from gas kinetic theory for the case of the diffuse reflection
surface the gas is supposed in equilibrium steady-state, so for the of the molecules from a surface. The coefficients aij have the fol-
molecules coming from infinity two parameters from three in the lowing values:
Maxwellian depend on the third one. Here we fix the macroscopic  
pffiffiffiffi 1 pffiffiffiffi pffiffiffiffi
flow velocity in the Maxwellian distribution function as a11 ¼ 2 p  b11 ; a12 ¼ a21 ¼ 2 pb12 ; a22 ¼ 2 pb22 ;
r
M
y ¼ H;t > 0; cy < 0; f1 ð24Þ
 5=2   where
1 p Ts Ts
¼ 3=2 1 exp  ðuy1  cÞ2 : ð18Þ
p ps T 1 T1 1 23 1 1 1 13
b11 ¼  ; b12 ¼ þ ; b22 ¼ þ : ð25Þ
At the upper boundary initially, at t ¼ 0, all three parameters,
p 32 16 5p 8 25p
p1 ; T 1 and uy1 are fixed, and the distribution function for the The numerical values of aij coefficients for r ¼ 1 are the
incoming molecules is calculated from Eq. (18), then only the following
macroscopic velocity is still kept constant, but other two parame- a11 ¼ 2:125; a12 ¼ a21 ¼ 0:447; a22 ¼ 1:030: ð26Þ
ters are obtained from previous time step.
To minimize the computational efforts, the cz variable is elimi- In Ref. [10] a particular approximation method was used to
nated by introducing the reduced distribution functions as in Ref. evaluate the numerical values of aij coefficients. Other approxima-
[17]. The Discrete Velocity Method [9] was used to solve Eq. (15) tion methods have also been used and give slightly different val-
with the boundary conditions Eqs. (16)–(18). The details of the ues, see Ref. [14].
numerical realization can be found in [18]. Finally the pressure and temperature jumps can be expressed as
following:
6. Jump boundary conditions ps  p1 J0 q0
¼ a11 m þ a12 ; ð27Þ
ps mns v s ps v s
In this Section we present the jump boundary conditions by fol-
lowing the approach based on the Onsager-Casimir reciprocity Ts  T1 J0 q0
¼ a21 m þ a22 : ð28Þ
relations, as it was presented in Ref. [19] for the case of evapora-
Ts mns v s ps v s
tion and condensation of a gas between two parallel condensed
phases. In the case of evaporation from a plate liquid surface we In previous expressions J 0m and q0 are the evaporation mass flux
can introduce, by analogy with [19], the thermodynamic ‘‘forces” (7) and the heat flux (6) outside from the Knudsen layer in the con-
as following tinuum part of the flow. As it is clear from the previous relations
452 A.Ph. Polikarpov et al. / International Journal of Heat and Mass Transfer 136 (2019) 449–456

that the intensity of pressure and temperature jumps is propor- iments [4] for measured large value of the temperature jump. We
tional to both mass and heat fluxes. It is worth to note that this show in the next Section that this expression works well only
form of jumps expression is similar to Eq. (13), but in the present when the p1 and T 1 are taken in the upper boundary of the Knud-
form the heat exchange is also considered. sen layer, where it is very difficult to make the measurements
As it was pointed out in [21] if one uses as the surface temper- because of the very thin thickness of this layer.
ature the temperature of the adjacent liquid, the results found
using non-equilibrium thermodynamics and the results obtained 6.2. Comments on jumps
from the kinetic theory are in perfect agreement with each other.
We can also express the fluxes in the function of forces from Eq. It is worth to discuss first the definition of the temperature
(22) as jump as it is used in the kinetic theory. This jump is defined as a
" # " #

J 0P K0PP K0PT XP a0022 a0012 XP difference between the solid (or liquid) surface temperature, T s ,
¼  ¼  ; ð29Þ and the gas temperature near the surface, Tjy¼0 . It is well known
J 0T K0TP K0TT XT a0021 a0011 XT
[27,14] that the near a surface a very thin layer, the Knudsen layer,
where a00ij ¼ aij ns v s =D and D ¼ a11 a22  a12 a21 . From (29) we have exists, which thickness is of the order of several molecular mean
 
J 0m 1 p  p1 Ts  T1 free path. Inside this layer the continuum approach does not valid
¼ a22 s  a12 ; ð30Þ
mns v s D ps Ts any more. Therefore the temperature jump boundary condition is
used for the Navier–Stokes (NS) equations:
 
q0 1 p  p1 Ts  T1 dT
¼ a21 s þ a11 : ð31Þ Tjy¼0  T s ¼ nT ‘ ; ð37Þ
ps v s D ps Ts dy
Previous relations are analogous to that obtained from non-
where nT is the temperature jump coefficient [12,11]. This condition
equilibrium thermodynamics [21,22]. We can provide the explicit
assures that the solution of the NS equation with the jump condi-
expressions for the surface resistivities
tion coincide with the solution of the Boltzmann equation (or of
a22 r r
¼ pffiffiffiffi ¼ pffiffiffiffi ; other kinetic equations) on the upper boundary of the Knudsen
D 2 pð1 þ rðb11  b12 =b22 ÞÞ 2 pð1  0:455rÞ
2
layer. It is clear from Eq. (37) that in the case of the gas - solid inter-
a12 rb12 0:434r face the temperature jump ðTjy¼0  T s Þ is proportional to the molec-
¼ pffiffiffiffi ¼ pffiffiffiffi ; ð32Þ
D 2 pb22 ð1 þ rðb11  b12 =b22 ÞÞ 2 pð1  0:455rÞ ular mean free path. Therefore this jump becomes negligible under
2

atmospheric conditions where the molecular mean free path is


a11 1 þ rb11 1  0:4r small, of the order of a micron. This temperature jump has to be
¼ pffiffiffiffi ¼ pffiffiffiffi :
D 2 pb22 ð1 þ rðb11  b12 =b22 ÞÞ
2
2 p 0:291ð1  0:455rÞ taken into account either under reduced pressure conditions or in
the microsystem applications, when the characteristic length-
scale of a flow is of order of tens hundred microns.
The numerical values of the coefficients provided above for When the liquid-gas interface is considered, this difference
evaporation coefficient equal to 1 are: between the gas temperature and surface temperature, T s  Tjy¼0 ,
a22 a12 a11 exists also. However, historically, the difference between the tem-
¼ 0:517; ¼ 0:225; ¼ 1:068: ð33Þ perature at the upper boundary of the Knudsen layer and the sur-
D D D
face temperature, T s  Tjy¼H is called the temperature jump. As in
the case of the gas-solid interface the NS equations do not valid
6.1. Hertz-Knudsen-Schrage formula
inside the Knudsen layer. Therefore, the boundary conditions,
Eqs. (27), (28), are proposed to use for the Navier–Stokes equations
More than one hundred year ago Hertz and Knudsen [23], con-
[28] to take into account the Knudsen layer influence. The imple-
sidering only the fluxes balance near the liquid interface, proposed
mentation of these conditions ensure that both solutions: the solu-
the equation which relate the evaporation flux to the liquid tem-
tion of the NS equations with temperature and pressure jump
perature (and pressure) and to the parameters on the upper
boundary conditions and the solution of the kinetic equation coin-
boundary of the Knudsen layer. The flux of the particles evaporated
cide on the upper boundary of the Knudsen layer.
from a surface was estimated from the gas kinetic theory as
rffiffiffiffiffiffiffiffiffiffiffiffi Furthermore, contrarily to the gas-solid interface, in the case of
ns v ms 8kB T s the gas-liquid interface one more condition for the pressure jump
J 0n ¼ ; where v ms ¼ ; ð34Þ
4 pm
where v ms is the average molecular velocity at the interface temper-
ature. The same molecular flux comes to the interface from the
Knudsen layer with the parameters n1 and T 1 . The balance of the
fluxes allows to derive the Hertz-Knudsen formula
rffiffiffiffiffiffiffiffiffiffiffi 
m ps p1
J 0m ¼ pffiffiffiffi
ffi  pffiffiffiffiffiffi
ffi : ð35Þ
2pkB Ts T1
This expression was improved by Kucherov and Rikenglas
[24,25] and by Schrage [26] by taking into account the macroscopic
vapor velocity and by introducing the evaporation coefficient as:
rffiffiffiffiffiffiffiffiffiffiffi 
2r m ps p1
J 0m ¼ pffiffiffiffi
ffi  pffiffiffiffiffiffi
ffi : ð36Þ
2r 2pkB Ts T1
Later, many various modifications of this expression were pro-
posed to much it with the measurements. However, this formula Fig. 2. Temperature profile and temperature jump definition at liquid-vapor
provides the evaporation flux much larger that one found in exper- interface.
A.Ph. Polikarpov et al. / International Journal of Heat and Mass Transfer 136 (2019) 449–456 453

exists. Both pressure and temperature jumps are proportional to Table 1


the mass and heat fluxes, and so depend on their intensity. Experimental data from Ref. [4]. The saturation pressure p1s is calculated from the
saturation temperature by using the expression provided in Ref. [29].
Fig. 2 schematically demonstrates the temperature profile nor-
mal to the liquid-vapor interface, located at y ¼ 0, as it can be Operating conditions [4]
obtained from the solution of a kinetic equation, see also next Sec- Experimental data 30  C 50  C 80  C
tion with the numerical results. From Fig. 2 it is clear that the 
T s ð CÞ 2:65 4:66 9:76
strong temperature gradient is observed inside the Knudsen layer, ps ðPaÞ1 738:8 851:2 291:9
then the temperature reaches asymptotically its value far from the T v ð CÞ 6:64 10:91 4:69
Knudsen layer in the continuum region. In adopted here simula- pv ðPaÞ 736:0 847:9 288:1
tions we assume the absence of the macroscopic parameter gradi- J  104 ½kg=ðm2 sÞ 5:78 7:66 11:9
ent outside of the Knudsen layer. q  104 ½W=m2  231:45 396:63 650:56

7. Results

We present here a first step of application of non-linear S-model Table 2


kinetic equation for modeling of evaporation process in the case, Experimental data from Ref. [6]. The saturation pressure p1s is calculated from the
saturation temperature by using the expression provided in Ref. [29]. The values of
where the calculations have been made under assumption of the the evaporation flux were additionally provided by the authors of Ref. [6].
constant temperature in the vapor continuum region, so the heat
flux in vapor phase is negligible. However, the model can be Operating conditions [6]

adopted to the situation, when the heat flux is important in the Experimental data Case 1 Case 4 Case 7
vapor continuum region. T s ð CÞ 10:82  0:05 4:52  0:05 4:08  0:05
ps ðPaÞ 265:7  1:3 435:7  2:2 815:5  4:1
7.1. Comparison with experiments ps ðPaÞ1 268:43 437:17 817:28
J  104 ½kg=ðm2 sÞ 3:6350 3:1967 1:9532

Recently several experiments are carried out to measure the


temperature discontinuities on the liquid-gas interfaces in the case
tal data of Ref. [4]. For each Figure the maximum value on the y
of pure substance evaporation [3,4,6] and in the case of presence of
axis corresponds to the value of saturation pressure (temperature)
non-condensable gas [5]. We analyse here the experimental
of the liquid layer. Both temperature and pressure jumps are visi-
results, provided from Refs. [4,6], where evaporation process of
ble on Figures and they are associated to the difference between
pure substance (water) is considered. To compare with the mea-
the saturated values on the interface and on the upper boundary
surements of the water evaporation the expressions provided in
of the Knudsen layer.
[29] is used to calculate the saturation pressure value from the
For two surface temperatures, T s ¼ 2:65  C and T s ¼ 4:66  C and
measured liquid water temperature:
corresponding saturation pressures, see Table 1, the molecular
psat ðTÞ ¼ k1 expðk2  k3 =T þ k4 T  k5 T 2 þ k6 T 3  k7 T 4  k8 lnðTÞÞ; mean free path, estimated using Eq. (1), is equal to 6.31 lm and
5.60 lm, respectively, so the Knudsen layer thickness for both
k1 ¼ 611:2; k2 ¼ 1045:8511577; k3 ¼ 21394:6662629;
cases is of the order of 2 mean free paths. It is worth to note that
k4 ¼ 1:0969044; k5 ¼ 1:3003741  103 ; k6 ¼ 7:747298  107 ; here we use the so-called equivalent mean free path, Eq. (1), while
k7 ¼ 2:1649005  1012 ; k8 ¼ 211:3896559: ð38Þ various other definitions exist in the literature, which take into
account different molecular interaction models. However, all these
expressions provide similar order of magnitude of the mean free
It is worth to underline that the specific temperature range of path. We would like also to underline that the calculated numeri-
the liquid water was used in the experiments, namely, the water cally Knudsen layer thickness is much smaller than thermocouple
was maintained at the liquid state for the temperatures below bed size, used in experiments [4], which was referred to be of 25
0  C, i.e. below its triple point. As it was mentioned in [29], the lm.
water is metastable in this temperature range and the measure- The profiles of the heat flux, in ½W=m2 , are also presented on
ments are impacted by the possibility of ice formation. Fig. 3 (e) and (f). The heat flux changes its sign through the Knud-
The S-model allows us to calculate the evaporation properties of sen layer: it is positive near the liquid surface and become negative
the monoatomic gas. However, the numerical results are compared outside of this layer.
with the experiments made with water evaporation. To do this, From Fig. 3 it is clear that all parameters have the gradients
first, all the numerical results were obtained in dimensionless inside the Knudsen layer, which thickness is around of 10 lm for
form, then, to provide the dimensional values of the parameters two cases. Outside the Knudsen layer all parameters are quasi con-
of interest the water vapor properties are implemented. Besides, stant. It is worth to note that only one dimensional problem is con-
it was shown in Refs. [30,15], that for the small evaporation rate sidered here, therefore the constancy of the mass, momentum and
the influence of the internal degree of freedom of a molecule on full energy fluxes have to be conserved. However, the numerical
the temperature and pressure jumps is still negligible, which justi- values of the mass, momentum and full energy fluxes, Eqs. (7)
fies here the implementation of the monoatomic gas model. and (8), which should theoretically be constant, show small varia-
Three sets of experimental data from Ref. [4] and three sets tions over 0 < y < 1. To quantify these variations we introduce the
from Ref. [6] were used as initial conditions for the numerical cal- deviation, devðJÞ, of the numerical values of a flux from its average,
culations and all these data are provided in Tables 1 and 2. The [31] i.e.,
indications 30  C; 50  C and 80  C are used to refer to the operating
1 X 1 X
N N
conditions from Ref. [4]. All simulations have been made with the
y
jJðyj Þ  J av j y

devðJÞ ¼ ; J av ¼ Jðy Þ; ð39Þ


evaporation coefficient equal to 1. N y j¼1 jJ av j Ny j¼1 j
Fig. 3(a)-(d) show the profiles of the pressure in [Pa] and tem-
perature [°C] as a function of a distance (in [lm]) from the liquid here Ny is the number of the computational points between 0 and 1
surface for two cases, heating 30  C and 50  C, from the experimen- in y direction. These deviations for the evaporation (mass) and
454 A.Ph. Polikarpov et al. / International Journal of Heat and Mass Transfer 136 (2019) 449–456

Fig. 3. (a) and (b) pressure profiles, (c) and (d) temperature profiles, (e) and (f) heat flux profiles. All the profiles are obtained numerically using the experimental data [4],
provided in Table 1, which correspond to heating 30  C and 50  C, for (a), (c), (e) and (b), (d), (f), respectively. For figures (a)-(d) the maximal value on y axis corresponds to the
saturation pressure (a), (b) and temperature (c), (d) of the interface.

energy fluxes are provided in Table 3 and they are used as the accu- umn, HKS1, Hertz-Knudsen-Schrage formula (36) with experiment
racy test of the numerical computations. data [4], with T s (ps ) and T 1 ¼ T v ; p1 ¼ pv given in Table 1; fourth
As it is clear from Fig. 3 the gradients of all macroscopic param- column HKS2, Hertz-Knudsen-Schrage formula (36) with T 1 and
eters exist in the Knudsen layer. As the continuum approach does p1 obtained from numerical solution of the S-model kinetic equa-
not allow to simulate the flow behaviors inside the Knudsen layer tion; fifth column presents the numerical S-model results, sixth
the values obtained from the numerical solution of the S-model column contains the experimental data [4].
kinetic equation, i.e. the value on the upper boundary of the Knud- Analysing the results presented in Table 4 we can conclude that
sen layer, must be used as the boundary conditions, when the con- the numerical solution of the S-model kinetic equation provides
tinuum approach is applied with the Navier–Stokes equation in the results on the evaporation mass flow rate which are very close
order to describe correctly the interface behaviors. to the measured values. Futher, expression (30), obtained from
Table 4 gives the values of the average over distance evapora-
tion rate J 0m , obtained by different ways. Second column presents
the results derived from the Onsager-Casimir theory, Eq. (30), with Table 4
the pressure and temperature values, p1 and T 1 , obtained numer- Mass flow rate J 0m  104 in kg=ðm2 sÞ, obtained from: Eq. (30), second column; HKS1,
ically from the solution of the S-model kinetic equation; third col- Hertz-Knudsen-Schrage formula (36) with experiment data T s (ps ) and
T 1 ¼ T v ; p1 ¼ pv from [4], third column; HKS2, Hertz-Knudsen-Schrage formula
(36) with experimental data T s (ps ), T 1 and p1 are taken from numerical solution of
the S-model kinetic equation, fourth column; numerical S-model results, fifth
Table 3
column; experimental data [4], sixth column.
Deviations, Eq. (39) for the evaporation (mass) and full energy fluxes for different
experimental conditions from Ref. [4]. J0m  104 ½kg=ðm2 sÞ
Operating conditions [4] 5 5
devðJ m Þ  10 devðJE Þ  10 Operating conditions [4] Eq. (30) HKS1 HKS2 S-model [4]
30  C 0:203 0:170 30  C 5:17 179:8 5:49 5:88 5:78
50  C 0:237 0:171 50  C 8:73 281:7 9:39 7:86 7:66
80  C 0:303 0:179 80  C 12:80 261:1 13:74 12:70 11:9
A.Ph. Polikarpov et al. / International Journal of Heat and Mass Transfer 136 (2019) 449–456 455

Onsager-Casimir theory gives also very similar values of the evap- Table 7
oration flow rate, when the values p1 and T 1 are taken from the Evaporation flux J0m in ½kg=ðm2 sÞ, comparison with measured values from Ref. [6].

numerical results. The Hertz-Knudsen-Schrage expression for mass J0m  104 ½kg=ðm2 sÞ
flow rate, Eq. (36), calculated also with the numerical S-model
Operating conditions [6] [6] S-model
results gives also similar values. However, when the same Hertz-
Case 1 3:6350 3:6329
Knudsen-Schrage expression, Eq. (36), is used but with p1 ¼ pv
Case 4 3:1967 3:1957
and T 1 ¼ T v (see Table 1) much larger values for the evaporation Case 7 1:9532 1:9530
mass flow rate is obtained. It means that the temperature T v and
pressure pv , measured in Ref. [4], do not correspond to the upper
boundary of a Knudsen layer. Therefore, Eq. (36) overestimates
Table 8
the evaporation rate, when temperature and pressure of vapor Pressure jump ps  p1 in ½Pa, obtained from: Eq. (27), where numerical values of J0m
are taken at some distance outside the Knudsen layer. and q0 from numerical solution using S-model equation are used, second column;
One additional comment related to the extraction of the evapo- numerical S-model results, third column.
ration coefficient can be done. Usually, to extract the values of this ps  p1 ½Pa
coefficient, the Hertz-Knudsen-Schrage expression for evaporation
Operating conditions [4] Eq. (27) S-model
rate, Eq. (36), is implemented. But usually the experimental values
of the vapor temperature and pressure far from the liquid interface 30  C 0:32 0:29
50  C 0:43 0:46
are used. Therefore, in order to much the measured evaporation
80  C 0:69 0:67
rate given by Eq. (36) very small evaporation coefficient have to
be used. In our opinion the large amount of the experimental data
on the evaporation coefficient, where its value was found very
small, can be related to this error.
could be partially explained by the formulation of the boundary
The values of the temperature jump, i.e. the difference between
condition on the upper boundary of the Knudsen layer. In both
the interface temperature, T s , and the vapor temperature at the
experiments the negative (directed to the liquid interface) heat
upper boundary of the Knudsen layer, T 1 , are given in Table 5. Sec-
flux exists in the vapor phase outside the Knudsen layer, while in
ond column provides the measured values of this jump; third col-
the present form of the boundary condition at the upper boundary
umn gives the values, calculated from Onsager-Casemir relation,
of the Knudsen layer the constant vapor temperature is assumed.
Eq. (28), but using the measured in [4] evaporation rate and heat
The values of the pressure jump, i.e. ps  p1 are provided in
flux; forth column presents the value obtained from the same
Table 8. Second column gives the jump values calculated from
Onsager-Casemir relation, Eq. (28), but with evaporation rate and
Eq. (27) using numerical values of the evaporation rate and heat
heat flux, obtained numerically from the solution of the S-model
flux; third column provides the values obtained directly from the
kinetic equation; forth column gives the temperature jump
numerical solution of the S-model kinetic equation. The values of
obtained numerically. One can see that the experimental values
the pressure jump obtained from both approaches are very similar.
are very large compared to the values obtained numerically for
Finally, the validation and improvement of the presented
very similar evaporation rate.
kinetic approache should be done by the detailed comparison with
Table 6 provides the values of the temperature jump obtained
the precise measurements for different operating conditions
for three sets of the experimental conditions from Ref. [6] (see also
requiring different rarefaction regimes. Such experimental data
Table 2). As for the experimental data from Ref. [4] the calculated
are practically missing. Such kind of data can be obtained by the
temperature jump is notably smaller than the measured one,
contact methods which use the microthermocouples as well as
despite the fact that the evaporation flux is reproduced numeri-
by the more difficult in the realization non-contact methods. As
cally with very good accuracy, see Table 7. Further, for two exper-
it was pointed out above, the pressure and temperature disconti-
imental data Refs. [4,6], the calculated vapor temperature near the
nuities on the vapor-liquid interface are proportional to the evap-
liquid interface is found lower than that measured one. This fact
oration rate and the heat flux through the interface. Therefore, for
the future experiments this point should be taken into account to
develop a new measurement system.
Table 5
Temperature jump T s  T 1 in  C, obtained from: experiments [4], second column; Eq.
(28)1, with experimental values of J0m and q0 from [4], third column; Eq. (28)2, where
numerical values of J 0m and q0 from numerical solution using S-model equation are 8. Conclusions
used, fourth column; numerical S-model results, fifth column.

T s  T 1 ½ C The kinetic approach is developed for numerical simulation of


Operating conditions [4] [4] Eq. (28)1 Eq. (28)2 S-model the evaporation process from a liquid surface. This approach allows
the detailed simulations of the vapor flow behaviors above its con-
30  C 3:99 0:15 0:025 0:036
50  C 6:25 0:23 0:029 0:028 densed phase. The temperature jumps obtained numerically for
80  C 14:44 1:09 0:131 0:124 different experimental conditions were found of the same order
as that measured recently and presented in Refs. [6,5], but much
smaller than that found previously in Ref. [4]. The comparison with
Table 6
the experimental data from Ref. [4] shows that the vapor parame-
Temperature jump, T s  T 1 , in ½ C, comparison with the measured values from Ref. ters are measured in [4] very far from the upper boundary of the
[6]. Knudsen layer and therefore the application of the Knudsen-
T s  T 1 ½ C
Hertz-Schrage formula predicts much higher mass flow rate as it
was really measured. Furthermore, if the values of the pressure
Operating conditions [6] [6] S-model
and temperature at the upper boundary of the Knudsen layer,
Case 1 0:36 0:039 obtained from the numerical solution of the S-model kinetic equa-
Case 4 0:24 0:022
tion, are implemented in the Knudsen-Hertz-Schrage formula the
Case 7 0:14 0:0075
evaporation rate is in excellent agreement with the measured
456 A.Ph. Polikarpov et al. / International Journal of Heat and Mass Transfer 136 (2019) 449–456

one. The proposed approach could be used for the measurements [5] E.Ya. Gatapova, I. Graur, O.A. Kabov, V.A. Aniskin, M.A. Filipenko, F. Sharipov,
The temperature jump at water-air interface during evaporation, Int. J. Heat
of the evaporation coefficient.
Mass Transfer (2017).
Besides, the measured and calculated evaporation fluxes are [6] M.A. Kazemi, D.S. Nobes, J.A.W. Elliott, Experimental and numerical study of
very close each other. However, the calculated vapor temperature the evaporation of water at low pressures, Langmuir 33 (18) (2017) 4578–
is found to be lower than that of the liquid interface, while in the 4591.
[7] P. Jafari, A. Masoudi, P. Irajizad, M. Nazari, V. Kashyap, B. Eslami, H. Ghasemi,
analyzed experiments [4,6] the measured near the liquid surface Evaporation mass flux: A predictive model and experiments, Langmuir 34
vapor temperature is higher than that of the liquid phase. This fact (2018) 11676–11684.
could be partially explained by the assumption of the constant [8] E.M. Shakhov, Generalization of the Krook kinetic relaxation equation, Fluid
Dyn. 3 (5) (1968) 95–96.
vapor temperature made in the numerical simulations. To improve [9] J.E. Broadwell, Shock structure in a simple discrete velocity gas, Phys. Fluids 7
the simulations the boundary conditions could be modified to take (8) (1964) 1243–1247.
into account the presence the heat flux in the vapor phase. In addi- [10] J.W. Cipolla Jr., H. Lang, S.K. Loyalka, Kinetic theory of condensation and
evaporation. II, J. Chem. Phys. 61 (1974) 69–77.
tion, to account more precisely the heat and mass exchanges [11] F. Sharipov, Rarefied Gas Dynamics. Fundamentals for Research and Practice,
between two phases the coupling between the continuum and WILEY-VCH Verlag GmBH & Co. KGaA., Weinheim, 2016.
kinetic approaches could be also realized. [12] M.N. Kogan, Rarefied Gas Dynamics, Plenum Press, New York, 1969.
[13] Y. Sone, H. Sugimoto, Strong evaporation from a plane condensed phase, in:
In practice very often the evaporation of one substance in the Adiabatic Waves in Liquid-Vapor Systems, Springer, New York, 1990, pp. 293–
presence of a non-condensable takes place, as in many cooling 304.
devices. Therefore, the next step will be the simulation of the fluid [14] Y. Sone, Kinetic Theory and Fluid Mechanics, Birkhäuser, Boston, 2002.
[15] A. Frezzotti, Bondary conditions at vapor-liquid interface, Phys. Fluids 23
evaporation into a mixture of the evaporated fluid and non-
(2011) 030609.
condensable gas by using the kinetic approach. The numerical [16] Y. Sone, Y. Onishi, Kinetic theory of evaporation and condensation -
results will be compared to the available experimental data [5]. hydrodynamic equation and slip boundary condition, J. Phys. Soc. Japan 44
(1978) 1981–1994.
[17] I.A. Graur, A. Polikarpov, Comparison of different kinetic models for the heat
Conflict of Interest transfer problem, Heat Mass Transf. 46 (2009) 237–244.
[18] M.T. Ho, L. Wu, I. Graur, Y. Zhang, J.M. Reese, Comparative study of the
Boltzmann and McCormack equations for Couette and Fourier flows of binary
The authors declared that there is no conflict of interest. gaseous mixtures, Int. J. Heat Mass Transfer 96 (2016) 29–41.
[19] F. Sharipov, Onsager-casimir reciprocity relations for open gaseous systems at
Acknowledgements arbitrary rarefaction. II Application of the theory for single gas, Phys. A 203
(1994) 457–485.
[20] H. Lang, Evaporation and condensation for general gas-liquid surface
The author (I. Graur) thank Mohammad Amin Kazemi and Janet scattering, J. Chem. Phys. 62 (3) (1975) 858–863.
Elliott for the provided measurements of the evaporation rate. [21] D. Bedeaux, L.J.F. Hermans, T. Ytrehus, Slow evaporation and condensation,
Phys. A 169 (1990) 263–280.
Problem statement and analysis were carried out under state con-
[22] D. Bedeaux, S. Kjelstrup, Transfer coefficients for evaporation, Phys. A 270
tract with IT SB RAS (AAAA-A17-117022850022-0). The numerical (1999) 413–426.
simulations were financially supported by the European Union net- [23] M. Knudsen, Die maximale verdampfungsgeschwindigkeit des que-cksilbers,
work program H2020, MIGRATE project under Grant Agreement Ann. Phys. Chem. 47 (1915) 697–708.
[24] R.Y. Kucherov, L.E. Rikenglas, Slipping and temperature discontinuity at the
No. 643095 (I. Graur) and by the RFBR according to the research boundary of a gas mixture, Zh. Eksp. Teor. Fiz. 36 (6) (1959) 125.
project No. 18-31-00194 (A. Polikarpov). A. Polikarpov also thanks [25] R.Y. Kucherov, L.E. Rikenglas, On hydrodynamic boundary conditions for
the Act 211 Government of the Russian Federation, contract 02. evaporation and condebsation, Sov. Phys. JETP 10 (37) (1960) 88–89.
[26] A. Schrage, A Theoretical Study of Interphase Mass Transfer, Columbia
A03.21.0006, for incentive pay. University Press, New York, 1953.
[27] C. Cercignani, Theory and Application of the Boltzmann Equation, Scottish
References Academic Press, Edinburgh, 1975.
[28] E. Ya Gatapova, I. Graur, F. Sharipov, O.A. Kabov, The temperature and pressure
jumps at the vapor-liquid interface: Application to a two-phase cooling
[1] P.N. Shankar, M.D. Deshpande, On the temperature distribution in liquid-vapor
system, Int. J. Heat Mass Transfer 83 (2015) 235–243.
phase change between plane liquid surfaces, Phys. Fluids A 2 (1990) 1030–
[29] F. Duan, I. Thompson, C.A. Ward, Statistical rate theory determination of water
1038.
properties below the triple point, J. Chem. Phys. 112 (2008) 8605–8613.
[2] C.A. Ward, G. Fang, Expression for predicting liquid evaporation flux: Statistical
[30] A. Frezzotti, A numerical investigation of the steady evaporation of a
rate theory approach, Phys. Rev. E 59 (1999) 429–440.
polyatomic gas, Eur. J. Mech. Ser. B Fluids 26 (1) (2007) 93–104.
[3] C.A. Ward, D. Stanga, Interfacial conditions during evaporation or condensation
[31] K. Aoki, N. Masukawa, Gas flows caused by evaporation and condensation on
of water, Phys. Rev. E 64 (2001) 051509.
two parallel condensed phases and the negative temperature gradient:
[4] V.K. Badam, V. Kumar, F. Durst, K. Danov, Experimental and theoretical
numerical analysis by using a nonlinear kinetic equation, Phys. Fluids 6 (3)
invesigations on interfacial temerature jumps during evaporation, Exp.
(1994) 1379–1395.
Thermal Fluid Sci. 32 (2007) 276–292.

You might also like