Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

Accepted Manuscript

Numerical simulation of evaporation of volatile liquids

A.D. Galeev, A.A. Salin, S.I. Ponikarov

PII: S0950-4230(15)30025-5
DOI: 10.1016/j.jlp.2015.08.011
Reference: JLPP 3032

To appear in: Journal of Loss Prevention in the Process Industries

Received Date: 21 November 2014


Revised Date: 27 June 2015
Accepted Date: 28 August 2015

Please cite this article as: Galeev, A.D., Salin, A.A., Ponikarov, S.I., Numerical simulation of evaporation
of volatile liquids, Journal of Loss Prevention in the Process Industries (2015), doi: 10.1016/
j.jlp.2015.08.011.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
1
ACCEPTED MANUSCRIPT
Numerical simulation of evaporation of volatile liquids

A.D. Galeev*, A.A. Salin, S.I. Ponikarov

Kazan National Research Technological University, Department of Mechanical Engineering, 68

Karl Marx Str., 420015 Kazan, Russian Federation

E-mail: galeev_ainur@mail.ru

PT
Abstract. The paper presents the results of the validation of the developed pool

evaporation model using literature and our own experimental data. The proposed model was

RI
used to examine the effect of wind velocity and pool sizes on the evaporation rate of volatile

SC
liquid (hexane). Contrary to the semi-empirical evaporation model widely used in hazard

assessment, stronger dependence of evaporation rate on pool size at low wind speeds is obtained.

U
Keywords: liquid spill, pool evaporation, model validation, numerical simulation
AN
1. Introduction
M

` The correct prediction of the consequences of hypothetical accidental scenarios is an


D

essential component for developing and implementing appropriate protective capabilities and
TE

plan mitigation measures. This requires the use of the reliable mathematical models describing

how toxic chemicals or flammable substances are released and dispersed in the air. A complete
EP

model for an event can roughly be divided into three parts: source modeling, dispersion

modeling and effect modeling (Vik & Pettersson Reif, 2011). There is a lot of work concerning
C

the modeling of hazardous substances dispersion in the environment, whereas little attention has
AC

been given to source term modeling. In emergency spills of stable liquids, the input of hazardous

substances into the environment is the result of evaporation from the pool surface.

Generally, there are two main types of mathematical models to evaluate the evaporation

rate from a pool: semi-empirical and computational fluid dynamics (CFD) models In the area of

hazard analysis of accidental chemical spills, many investigators and hazard reference books

*
Corresponding author. Tel.:+7843 231 42 41; fax: +7843 231 42 41
E-mail address: galeev_ainur@mail.ru (A.D.Galeev)
2
ACCEPTED MANUSCRIPT
have applied Mackay and Matsugu mass transfer equation (Kawamura & Mackay, 1987;

Mackay & Matsugu, 1973). In the model (Kapias & Griffiths, 1999), the mass transfer

coefficient is calculated using Brighton’s theory (Brighton, 1987). This analytical model takes

into account the effects of surface roughness, friction velocity of the airflow, and the effect of

high vapor pressure on the mass transfer process. Khajehnajafi et al. (2011) used Churchill’s

PT
equation (Churchill, 1976) where the Nusselt number is determined for the complete range of

Re and Pr covering laminar, transition and turbulent regimes. This equation is based on

RI
experimental data for forced convection from isothermal flat plates. The Churchill’s equation is

SC
applied to mass transfer calculation substituting the Sh and Sc for the Nu and Pr. The

mathematical model, represented by Habib et al. (2009), based on numerical calculation of

U
differential transient equations for boundary layer together with algebraic turbulence model of
AN
Cebeci-Chang (Cebeci & Chang, 1978), including correction for surface roughness. The above

models have been developed on the assumption that the vapor behaves as a passive contaminant,
M

i.e. it does not affect the flow field. This assumption may be invalid when the molecular weight

of evaporating component differs from the molecular weight of environment air (Desoutter,
D

Habchi, Cuenot & Poinsot, 2009).


TE

By using the CFD, it is possible to overcome the limitations of existing semi-empirical


EP

models. This approach by solving three-dimensional conservation equations for mass,

momentum and energy makes it possible to take into account the complex interaction between
C

the evaporation process and the vapor dispersion. Many works have used CFD to examine the
AC

evaporation rate of water (Raimundo, Gaspar, Oliveira, & Quintela, 2014) or dilute solutions

of ammonia (Rong, Nielsen, & Zhang, 2010; Rong, Elhadidi, Khalifa, Nielsen, & Zhang,

2011; Saha, Wu, Zhang, & Bjerg, 2011). In the work (Rong, Nielsen, & Zhang, 2010) the

effects of airflow and aqueous ammonia solution temperature on ammonia transfer are

investigated and the numerical sensitivity studies are performed by Rong et al. (2011) to

examine the effects of computational geometry and inlet turbulence intensity on ammonia
3
ACCEPTED MANUSCRIPT
emissions. In these studies, the ammonia was a passive contaminant and has no impacts on

airflow. Raimundo et al. (2014) used CFD model to investigate the mass transfer at the free

surface of the water tank. In the abovementioned works, the liquid-air interface is assumed to be

isothermal due to the small liquid volatility. When volatile liquids, such as hexane and acetone,

evaporate, the liquid temperature may significantly change due to heat loss. Furthermore, the

PT
evaporation rate of volatile liquids may be significantly influenced by Stefan flux, which is not

taken into account in above CFD models.

RI
Liu, Olewski, Vechot, Gao & Mannan (2011) developed a methodology using CFD to

SC
simulate the boiling process of liquid nitrogen. The vaporization rate of a boiling liquid is

governed by the heat transfer phenomena including conduction, convection and thermal radiation

U
mechanisms (Véchot, Olewski, Osorio, Basha, Liu, & Mannan, 2013), while the evaporation
AN
rate of non-boiling liquids is controlled by the removal of vapor from the pool surface by airflow

(van den Bosch, 1997). Thus, for boiling and non-boiling liquids the using of CFD code is
M

reasonable to improve the prediction of convective heat and mass transfer, respectively.

The present paper aims to validate the developed pool evaporation CFD model using
D

literature and our own experimental data and to study the effect of the pool sizes on the
TE

evaporation rate from volatile liquid spill using numerical simulation. In the previous work
EP

(Galeev, Starovoytova & Ponikarov, 2013) the pool evaporation model was validated using

only experimental data on liquefied butane evaporation. However, before to use a mathematical
C

model in risk assessment, the model must be carefully tested against experimental data from
AC

different sources. Therefore, in the present work we also compared simulation results both with

the experimental data on evaporation of ethanol and cyclohexane (Habib, Schalau, Acikalin &

Steinbach, 2009) and with our own experimental data on evaporation of acetone. The pool

evaporation rate sensitivity due to pool length has not been fully addressed in the literature. In

the work (Khajehnajafi & Pourdarvish, 2011) the pool size determines the different flow

regimes over the flat surface. Depending on the intensity of turbulence in the main stream, the
4
ACCEPTED MANUSCRIPT
flow near the surface becomes turbulent at some downstream distance (Schlichting, 1968). The

point of transition from laminar to turbulent boundary layer flow may be estimated in terms of

the length Reynolds number, ReL = ULν-1, where U is the main stream air velocity (ms-1), L is

the pool length in the direction of the air flow (m) and ν is the kinematic viscosity of air

(Nielsen, Olsen, & Fredenslund, 1995). For the length Reynolds number, the critical value of

PT
5×105 is often assumed (Incropera, DeWitt, Bergman, & Lavine, 2006). The critical values for

the length Reynolds numbers from 1×105 to 3×105 have been reported (Pasquill, 1943; Bird,

RI
Stewart, & Lightfoot, 1960; Schlichting, 1968; Coulson & Richardson, 1993; Nielsen,

SC
Olsen, & Fredenslund, 1995). When considering accidental spills, the pool sizes are large and

ReL exceeds the critical value and the prevalence of turbulent mechanism of the transport of

U
vapor away from the evaporating pool is expected. In the model (Mackay & Matsugu, 1973;
AN
Kawamura & Mackay, 1987) it is assumed, that evaporation rate Jg,s depends on pool size L as

Jg,s∼L-0,11. The exponent on the pool dimension L was obtained by evaporating water from pools
M

of different sizes (Mackay & Matsugu, 1973). This exponent does not reflect the buoyancy
D

effects because the evaporation rate of water is low. The buoyancy effects are caused by the
TE

density difference between the evaporating component and ambient air and they may be

significant when a pool of volatile liquid has large sizes and wind speed is low. It is evident that
EP

further studies are necessary to bring a better understanding to the role of pool dimensions and

buoyancy effects in evaporation of volatile liquid.


C
AC

2. Pool evaporation model and its validation

The mass flow of vapor from the pool surface Jg,s due to evaporation was calculated on

the basis of the standard wall functions (Fluent, 2006) taking into account the correction

for Stefan flow:

(Y g , s − Y g , P ) ρC µ0.25 k P0.5
J g ,s = K (1)
Y+
5
ACCEPTED MANUSCRIPT
Sc ⋅ y + + +
( y < yC )
Y+ =  ; (2)
Sct ( u + + PC ) ( y + > yC+ )

ρC µ0.25 k P0.5 y P
y+ = ; (3)
µ

u+ =
1
κ
( )
ln Ey + − ∆B ; (4)

PT

0 ( K s+ < 2.25 )

RI

 1  K + − 2.25
∆B =  ln s

{
+ Cs ⋅ K s+  ⋅ sin 0.4258 ⋅ lnK s+ − 0.811 ( )} (2.25 < K s+ < 90 ) ; (5)
 κ  87.75 

SC
1
(
 ln 1 + Cs ⋅ K s+
κ
) ( K s+ > 90 )

U
ρ ⋅ K s ⋅ Cµ0.25k P0.5
K s+ = ; (6)
µ
AN
 Sc 
34

PC = 9.24  − 1[1 + 0.28exp(− 0.007 Sc Sct )] ; (7)
M

 Sct  

ln ((1 − Yg ,P ) (1 − Yg ,s ))
D

K= . (8)
Yg ,s − Yg ,P
TE

It is assumed that the transition between the fully turbulent region and the viscous layer
EP

near the wall occurs at a value yC+ of 11, independent of the concentration of the evaporating

component.
C

In Eqs. (1) – (6) the coefficient Cµ is assumed to be a constant: Cµ=0.09. In the gas
AC

dispersion model, this coefficient is determined by relating Cµ to the mean flow deformation as

adopted in the Realizable k-ε turbulence model (Shih, Liou, Shabbir, & Zhu, 1995).

The change of local liquid temperature is calculated from the equation

∂Tliq qa + qgrd + qs − q p + qar − J g ,s ⋅ ∆H g


= , (9)
∂t CP ,liq ⋅ mliq
6
ACCEPTED MANUSCRIPT
2
where qa is the heat flux from the atmosphere, W/m ; qgrd is the heat flux from the ground to the

liquid, W/m2; qs is the heat flux from solar radiation, W/m2; qp is the heat flux emitted by the

pool due to long-wave radiation, W/m2; qar is the heat flux absorbed by the pool due to long-

wave radiation from the atmosphere, W/m2

The heat flux from the ground to the liquid phase, qgrd = λgrd(∂Tgrd/∂y)y=0 is found from

PT
the numerical solution of three-dimensional nonstationary heat conduction equation for the

substrate as was done in the work (Galeev, Starovoytova, & Ponikarov, 2013).

RI
The heat flux from the atmosphere qa is calculated using wall functions (Fluent, 2006).

SC
The qs, qp, qar were determined from the equations given in the paper (Kawamura &

Mackay, 1987).

U
The change in liquid mass is calculated from the equation:
AN
∂mliq
= − J g ,s . (10)
∂t
M

The equations (9) and (10) were incorporated into the CFD code FLUENT as user-

defined scalar (UDS) transport equations without convection and diffusion terms.
D

The mass flux Jg,s is used as a boundary condition for the dispersion model in the pool
TE

area. The atmospheric dispersion of vapor is calculated by solving the three-dimensional

Reynolds-averaged Navier-Stokes equations and the energy and species equations. The
EP

governing equations are closed using the Realizable k–ε equations for turbulence.
C

To calculate the dispersion of vapor in the atmosphere, it is important to define the


AC

correct profiles for the mean velocity, turbulence kinetic energy and its dissipation rate on the

inlet to the computational domain. These inlet profiles should be maintained throughout a

calculation domain. The fully developed profiles provided by Richards & Hoxey (1993) are

mathematically consistent, i.e. they are a solution of the mathematical models describing a

homogeneous atmospheric boundary layer (Parente, Gorle, van Beeck, & Benocci, 2011).

However, the assumption of constant kinetic energy, k is not consistent with wind-tunnel

measurements (Robins, Castro, Hayden, Steggel, Contini & Hesit, 2001; Yang, Gu, Chen, &
7
ACCEPTED MANUSCRIPT
Jin, 2009) where a variation of k with height is generally observed. Yang et. al. (2009) proposed

a new set of inlet conditions where the k profile is a function of height. Gorle, van Beeck,

Rambaud, & van Tendeloo (2009) proposed formulations for the turbulence model coefficients

to ensure stream-wise homogeneity while using the k profile proposed by Yang et al. (2009). In

the present work the inlet profiles were determined through repeated calculations when the

PT
velocity and turbulence parameters profiles deduced at the outlet section of the computational

domain were set at its inlet boundary and the computation was repeated until the profiles for

RI
outlet and inlet boundaries were close. The plots in Figs. 1-3 show the comparison between

SC
calculated profiles and profiles by Yang et al. (2009). These profiles were obtained at wind

velocities of 1, 2.5 and 5 m/s and under neutral atmospheric stability. The plots show that there is

U
no significant difference in the velocity and ε at the wall while there is the difference in k. The
AN
similar discrepancy between experimental and simulated profiles of k is observed in the work

(Mokhtarzadeh-Dehghan, Aksayoglu, & Robins, 2012). At close velocity profiles, the impact
M

of using different turbulence parameters profiles on the results of vapor dispersion calculation

will be governed by the parameter µt/Sct. Due to the fact that coefficient Cµ=0.028 in correlations
D

of Yang et al. (2009) is much lower than the same coefficient in the used Realizable k-ε model
TE

(is equal to about 0.09-0.14 depending on height from wall), the turbulent viscosities at the wall
EP

practically coincide. By optimizing the value of turbulent Schmidt number, it is possible to reach

a good agreement between the predicted and measured parameters for both approaches. In the
C

work (Galeev, Salin, & Ponikarov, 2013), the results of calculation of neutral gas dispersion
AC

with Sct = 0.5 and calculated profiles, agreed better with experiment data than with Sct=0.7,

whereas in the work (Galeev & Ponikarov, 2014) the good agreement between simulation and

experiment on heavy gas dispersion was obtained at Sct=0.7. In the present work numerical

simulations are performed with calculated profiles for mean velocity and turbulence parameters

and with Sct=0.7


8
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP

Fig. 1. Profiles of (a) velocity U, (b) turbulence kinetic energy k, (c) turbulence kinetic energy

dissipation rate ε and (d) turbulent viscosity µt from numerical calculation (solid line) and from Yang et
C

al. (2009) (dashed line) at the wind velocity of 1 m/s at the height of 10 m
AC
9
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE

Fig. 2. Profiles of (a) velocity U, (b) turbulence kinetic energy k, (c) turbulence kinetic energy
EP

dissipation rate ε and (d) turbulent viscosity µt from numerical calculation (solid line) and from Yang et

al. (2009) (dashed line) at the wind velocity of 2.5 m/s at the height of 10 m
C
AC
10
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP

Fig. 3. Profiles of (a) velocity U, (b) turbulence kinetic energy k, (c) turbulence kinetic energy

dissipation rate ε and (d) turbulent viscosity µt from numerical calculation (solid line) and from Yang et
C

al. (2009) (dashed line) at the wind velocity of 5 m/s at the height of 10 m
AC

To validate the CFD evaporation model, the predicted results were compared with the

experimental data obtained by Habib et. al. (2009). They carried out experiments in the open air

for evaporation of ethanol and cyclohexane at constant temperatures. One facility comprised a

flat terrain without buildings whilst the other was situated in a very rough area. The pool

consisted of a circular basin with a diameter of 0.74 m which was insulated from the ground. The

experimental data for cyclohexane are presented in the work (Khajehnajafi & Pourdarvish.

2011).
11
ACCEPTED MANUSCRIPT
The difference between the pool evaporation model used in the present study and the one

used in our previous works (Galeev, Starovoytova, & Ponikarov, 2013; Galeev, Salin, &

Ponikarov, 2013; Galeev & Ponikarov, 2014) is that the model in the present study takes into

account roughness for very rough area when the roughness height larger than the height of the

centre point of the wall-adjacent cell.

PT
The two parameters of the rough wall model, i.e. the roughness height Ks and roughness

constant Cs, were determined according to the relationship between Ks, Cs and z0 derived

RI
(Blocken, Stathopoulos, & Carmeliet, 2007):

K S = 9.793 ⋅ z 0 C s .

SC
(11)

In the calculation, it is assumed that the aerodynamic roughness of flat terrain z0 is equal

U
to pool surface roughness z0=0.0002 (Ks=0.00195 m) and Cs=1. For the very rough area the
AN
aerodynamic roughness is set equal to 0.04 m (Ks=0.39 m). Although in the work (Troen &

Petersen, 1989) for area with many trees and buildings it is recommended to use the value of
M

z0=0.1 m we assumed a lower value to avoid the divergence problems during iterative process.
D

Ks should be smaller than the height of the centre point of the wall-adjacent cell. The height of
TE

wall-adjacent cells was 0.1 m, therefore, in the simulation of a very rough area, the roughness

height, Ks, was set to the value of 0.05 m and from Eq. (11), the corresponding value of constant
EP

Cs was 7.83. Fluent Manual states that constant Cs should be smaller or equal to 1 and it is not

possible to set a higher value via a user interface. Therefore, a higher value is defined through a
C

User Defined Function. This approach to set roughness parameters in Fluent was used in the gas
AC

dispersion calculation. by Labovsky and Jelemensky (2011)

The computational domain with dimensions 100 m×500 m×100 m in the x, y, z directions

was used. A structured non-uniform grid consisting of parallelepiped cells was generated. The

grid cells were clustered in the region of the pool. The number of grid cells was approximately

96,000.
12
ACCEPTED MANUSCRIPT
The second-order scheme was used for the discretization of the advective terms of the

governing equations. The PRESTO! algorithm (Fluent, 2006) to calculate the pressure and the

SIMPLE algorithm (Patankar, 1980) to correlate the velocity field and pressure were used. The

convergence criterion for dispersed material concentration was 1×10-4, for other variables was

1×10-3. The maximum time step of 2 s was used.

PT
Accuracy of the models was estimated by the value of percent error (%):

δ = 100 ⋅ J g, s, calc − J g, s, expt J g, s, expt . (12)

RI
Tables 1 and 2 show experimental data and prediction values for evaporation of ethanol

SC
and cyclohexane at different temperatures and wind speeds. The computed values of the

Mackay-Matsugu equation are obtained with a difference δ to the measured ones between 27.5 to

U
61.1 %, with an average value of 47.4 %. The computed values of the numerical model are
AN
obtained with a difference δ to the measured ones between 0.6 to 47.4 %, with an average value
M

of 12.7 %.

Table 1
D

Experimental data and predictions for ethanol evaporation


TE

Pool temperature, Wind velocity at 2 Mass flow, g/s


К m height, m/s Experiment Mackay – CFD model
Matsugu Eq.
EP

Flat terrain
310.15 1.4 0.456 0.681 0.393
309.65 1.9 0.533 0.855 0.523
C

325.15 1.7 1.02 1.59 1.06


Very rough terrain
AC

305.65 1.4 0.488 0.663 0.536


310.15 1.8 0.679 0.99 0.9
324.65 1.8 1.32 1.95 1.96

Table 2

Experimental data and predictions for cyclohexane evaporation

Pool temperature, Wind velocity at 2 Mass flow, g/s


К m height, m/s Experiment Mackay – CFD model
Matsugu Eq.
13
ACCEPTED MANUSCRIPT
Flat terrain
303 2.71 1.08 1.74 1.11
310 3.05 1.63 2.49 1.66
317 3.49 2.33 3.63 2.55
Very rough terrain
303 1.71 0.967 1.39 1.15
310 1.62 1.35 1.80 1.46
317 1.4 1.67 2.13 1.68
To validate the evaporation model, our own experimental data for acetone evaporation

PT
were also used. The liquid was poured into the pan with sizes of 0.6 m×0.4 m×0.035 m. The air

velocity was measured at the height of 2 m by an acoustical anemometer with an error. ± 0.2 m/s.

RI
An averaging time of the wind speed was 2 min. To measure the change in mass of the pan with

SC
the liquid, the Mettler-Toledo XP8002S balance was used with an accuracy of 0.01 g. To

measure the temperature of the liquid the thermocouple chromel-copel with an accuracy of 0.1°C

was used.
U
AN
The measurements were carried out at the air temperature of 23˚C and humidity of 40%.

The initial temperature of acetone was equal to 23.5˚C. Wind speed was varied during the
M

experiment (Fig. 4). In the calculation the average wind speed for the considered time period was
D

used, which was equal to 1.05 m/s. The solar flux was determined using the equation given in the
TE

paper (Kawamura and Mackay, 1987), with taking into account the latitude and longitude of

terrain and cloud cover, and its value was 430 W/m2. The heat fluxes due to the long-wave
EP

radiation of the atmosphere and the pool surface have been taken into account through the use of

formulae from the same source. The heat flux from the ground is assumed to be equal to zero
C

because the pan with liquid was placed on the balance. The aerodynamic roughness of the area
AC

around the pool is set equal to 0.01 m (Ks=0.098 m), that corresponds to the area with short grass

(Troen & Petersen, 1989).

The Fig. 5 shows that the model somewhat underestimates the evaporated mass. The

difference between the measured and calculated evaporated mass is 17.5%. The difference

between experimental and predicted dependencies of liquid temperature is explained by the

difference in evaporation rates. It should be noted, that up to about 10 min, the experimental and
14
ACCEPTED MANUSCRIPT
calculated values of the mass and temperature are close, although the wind velocity in

calculation (1.05 m/s) was about 2.5 times greater than the one in the experiment . This could be

explained by the fact that at very low wind speeds and at the small pool size the molecular

mechanism of vapor removal from the evaporation surface could dominate, when the sensitivity

of evaporation to wind velocity is not as high as in the turbulent regime. Conventionally, the

PT
Sherwood number, Sh, (the dimensionless parameter describing the mass transfer) is proportional

to Re-0.5 for laminar flow over a flat surface, while Sh∼Re-0.8 in turbulent flow (Incropera,

RI
DeWitt, Bergman, & Lavine, 2006).

U SC
AN
M
D
TE

Fig. 4. Measured wind velocity


C EP
AC
15
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 5. Measurements (solid line) and predictions (dashed line) of liquid mass and temperature

U
3. Sensitivity study and discussion
AN
This section presents the numerical analysis of the effect of pool sizes on evaporation of

hexane. Hexane has high volatility (at a liquid temperature of 303 К the hexane vapor pressure is
M

equal to 25 kPa) and hexane vapor is 3 times heavier than the ambient air. These factors affect
D

pool evaporation and vapor cloud dispersion processes. The pools with sizes of 10 m×10 m, 30
TE

m×30 m, 60 m×60 m and 100 m×100 m are considered. The initial liquid layer height was equal

to 0.05 m. The initial temperature of ambient air and pool substrate was 303K. The numerical
EP

study included a series of calculations at wind speeds 1, 2.5 and 5 m/s at the height of 10 m. The

solar flux had a range of 0-900 W/m2 (Khajehnajafi & Pourdarvish. 2011), therefore in
C

calculation the average value qs=450 W/m2 was used.


AC

The geometry of the computational domain is shown in Fig. 6. The calculation was

performed for half of the pool because of symmetry of the problem. The computational domain

with dimensions 2000 m×500m×500m in the x-, y-, z-directions was used. A structured

nonuniform grid consisting of parallelepiped cells was generated. The number of grid cells was

approximately 370,000. The height of wall-adjacent cells was equal to 0.2 m. In downwind (x-

axis) and crosswind (z-axis) directions, the grid cell size was 1 m over the pool surface.
16
ACCEPTED MANUSCRIPT
Additionally to the above domain, a domain under the pool containing the ground of 2 m depth

was also included for in the calculations. The profiles of wind velocity, k and ε on the inlet to the

computational domain (plane ABCD) were obtained through calculation and are shown in Figs.

1-3. On the lateral and upper boundaries (ABFE, BCGF and DCGH) of the domain the

symmetry condition was specified, i.e. velocity component normal to the boundary and normal

PT
derivatives of the other variables assumed to be equal to zero. The bottom boundary (ADHE)

was defined as a no-slip wall. At the outlet boundary (EFGH) the gauge pressure was fixed at

RI
zero. The view of used computational grid near the pool is shown in Fig. 7.

U SC
AN
M
D
TE
EP

Fig. 6. Schematic representation of the computational domain


C
AC
17
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 7. The computational grid near the pool

The roughness coefficient value Cs = 1 was applied in this study. The roughness height Ks

U
for the pool surface was given as 0.00195 m and for the adjacent territory was 0.098 m. The
AN
liquid temperature was imposed as a thermal boundary condition on the wall under the pool
M

while the wall around the pool was considered adiabatic.

In the simulation, it was assumed that physical properties of liquid, gas and solid layer
D

did not depend on temperature. The following values have been used: Dm,g=6.9⋅10-6 m2/sec;
TE

µg=6.6⋅10-6 kg/(m⋅sec); µa=1.78⋅10-5 kg/(m⋅sec); CP.g=1680 J/(kg⋅К); CP.a=1006 J/(kg⋅К);

CP.liq=2300 J/(kg⋅К); λg = 0.013 W/(m⋅К); λa = 0.0242 W/(m⋅К); ∆Нg=370000 J/kg; λgrd = 1.28
EP

W/(m⋅К). СP.grd=1130 J/(kg⋅К). ρgrd=2300 kg /m3. Here, the Dm, µ, CP, λ, ρ, denote molecular
C

diffusion coefficient, molecular viscosity, specific heat, thermal conductivity and density,
AC

respectively. The a, g and grd subscripts indicate air, evaporating gas and ground value,

respectively.

The vapor pressure Pg of hexane as a function of liquid temperature Tliq was determined

using the formula (Green & Perry, 2008):

( )
Pg (Tliq ) = exp A + B / Tliq + C ⋅ ln(Tliq ) + D ⋅ TliqE , (13)
18
ACCEPTED MANUSCRIPT
where A=104.65, B=-6995.5, C=-12.702, D=0.000012381, E=2.

In Fig. 8 the plots of specific evaporated mass mvap and volume-averaged liquid

temperature Tliq against time are presented. The plots show that the specific evaporated mass is

higher when the pool sizes are larger, in other conditions being equal. The influence of pool sizes

on the evaporation decreases when the wind velocity increases. As the pool sizes increase the

PT
vapor concentration above the surface of the evaporating liquid and, consequently, the vapor-air

cloud density increases. The negative vertical density gradient appearing above pool surface

RI
causes the suppression of turbulence (stable stratification) that creates a large resistance to the

SC
transport of vapor away from the evaporating pool and results in a slow evaporation rate. In

addition to the buoyancy effect, the increasing of vapor concentration above the pool surface

U
reduces the driving force of the evaporation. The graphs of temperature change are consistent
AN
with the time dependence of mvap. A higher evaporation rate results in a lower pool temperature,

since the heat losses increase with the increasing of evaporation rate. At the wind velocity of 1
M

m/s, the temperature rises throughout the considered period of time because heat flux from the

sun to pool prevails over the heat losses due to evaporation.


D

The regression analysis of numerical results has shown that at the wind speed of 1 m/s the
TE

average specific evaporated mass depends on a pool size as mvap∼L-0.553 with accuracy of
EP

approximation of R2=0.9981, at wind speed of 2.5 m/s as mvap∼L-0.318 (R2=0.9153) and at wind

speed of 5 m/s as mvap∼L-0.0578 (R2=0.9755). Thus, at low wind velocities the stronger
C

dependence of evaporation rate on pool size is observed, than the one (mvap∼L-0.11) adopted in the
AC

Mackay-Matsugu equation commonly used in hazard assessment (Kawamura & Mackay,

1987; Mackay & Matsugu, 1973).

Table 3 presents the values of the physical properties of gases and liquids for the upper

and lower limits of the variation range of liquid temperature. At these temperatures the physical

properties do not differ significantly, that confirms the validity of the assumption of constant
19
ACCEPTED MANUSCRIPT
physical properties. The molecular diffusion coefficient was determined using Gilliland’s

correlation (Gilliland, 1934) and other values were taken from the ChemCAD program database.

Table 3

Physical properties of gases and liquid

Physical property Temperature, K Variation, %


286 308

PT
Dm,g, m2/sec 6.48⋅10-6 7.24E-06 11,7
µg, kg/(m⋅sec) 6.5⋅10-6 6.7⋅10-6 3,1
µa, kg/(m⋅sec) 1.78⋅10-5 1.88⋅10-5 5,6

RI
CP.g, J/(kg⋅К) 1590 1690 6,3
CP.a, J/(kg⋅К) 1005 1008 0,3
CP.liq, J/(kg⋅К) 2215 2310 4,3

SC
λg, W/(m⋅К) 0.0115 0.0135 17,4
λa, W/(m⋅К) 0.024 0.026 8,3
∆Нg, J/kg 380000 362500 4,6

U
AN
M
D
TE
C EP
AC
20
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Fig. 8. Dependencies of specific evaporated mass (left) and average liquid temperature

(right) against time at wind velocities of 1 m/s (a), 2.5 m/s (b) and 5 m/s (c)
21
ACCEPTED MANUSCRIPT
In Fig. 9 the time dependences of the relative turbulent viscosity and the driving force

(∆Yg=Yg,s-Yg,P) are shown. The relative turbulent viscosity is defined as the dynamic coefficient

of average turbulent viscosity above the pool surface (y=0.1 m) divided by the turbulent

viscosity in the absence of the disturbing effect of the source. The relative turbulent viscosity for

an undisturbed flow (when there is no hexane vapor) is equal to 1.0 for each of the wind

PT
velocities. In the initial period of evaporation, a sharp decay of turbulent viscosity occurs due to

the formation of the negative density gradient above pool surface. The importance of buoyancy

RI
effect rises as the air velocity decreases and pool sizes increase. The slight growth of the eddy

SC
viscosity at wind speeds of 2.5 m/s and 5 m/s is attributed to the reduction of liquid temperature,

and, as consequence, the reduction of mass flux of vapor from the pool surface.

U
At low wind velocities, the driving force of evaporation significantly drops as pool sizes
AN
increase due to the increasing of the concentration above the pool surface. Since in the model the

driving force is both in the numerator and the denominator of the formula for the diffusion flux,
M

its effect on the evaporation rate will be determined by the factor ln((1-Yg,P)/(1-Yg,s)). At the wind
D

speed of 1 m/s when the pool sizes increase from 10 m×10 m to 100 m×100 m the average factor
TE

reduces by 11%, at the wind speed of 2.5 m/s – by 6.4% and at the wind speed of 5 м/с – by

3.5%, while average specific evaporated mass mvap reduces by 71.5 %, 53% and 12.8 %,
EP

respectively. Thus, changes in the evaporation rate with increasing of pool sizes are determined

mainly by the effect of turbulence suppression over the pool surface, and not by changes in the
C

driving force.
AC
22
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Рис. 9. Dependencies of average relative turbulent viscosity (left) and driving force

(right) at wind speeds of 1 m/s (a), 2.5 m/s (b) and 5 м/с (c).
23
ACCEPTED MANUSCRIPT

In the paper (Britter, 1989), a criterion that indicates when the plume may be considered

passive, i.e. when the influence of the density difference on the inertia small and may be

neglected, is given. For continuous source this criterion is

B = ( g 0′ ⋅ q 0 / L ) U ≤ 0.15 ,
13
(14)

PT
where g 0′ = g (ρ g − ρ a ) / ρ a ; q0 is the volume flow rate; L is the source dimension and U is the

RI
ambient velocity.

For determining the relative importance of the flow regimes (buoyancy-dominated, stably

SC
stratified and passive dispersion) the Richardson number, Ric, is also used (Havens, 1992):

Ric = g 0′ ⋅ H / u*2 ,

U
AN (15)

where H=q0/UL is the characteristic cloud height and u* is the friction velocity

If Ric ≥30 the flow is negative buoyancy dominated, if 1≤Ric≤30, the shear flow is stably
M

stratified and if Ric≤1 the passive dispersion occurs (Havens, 1992)


D

In Table 4 the values of B и Ric for each of considered cases are given. In the calculation
TE

of these criteria, the average values of the gas flow rate have been substituted in expressions (14)

and (15).
EP

Table 4

Values of B and Ric depending on pool sizes and wind velocity


C

Pool sizes Britter’s criterion B (Richardson number Ric)


AC

1 m/s 2.5 m/s 5 m/s


10 m×10 m 0.35 (12.4) 0.20 (2.2) 0.12 (0.5)
30 m×30 m 0.42 (20.3) 0.27 (5.5) 0.17 (1.3)
60 m×60 m 0.46 (26.7) 0.31 (8.4) 0.21 (2.6)
100 m×100 m 0.51 (35.9) 0.33 (10.3) 0.25 (4.1)
The values of Britter’s criterion are less than 0.15 only when the pool sizes are 10 m×10

m and the wind speed is 5 m/s. This result is confirmed by the graphs (see Fig. 9), which show

that the change of turbulence due to buoyancy effects is small at wind speed of 5 m/s and at the

pool sizes of 10 m×10 m, while in the rest cases the appreciable suppression of turbulence (stable
24
ACCEPTED MANUSCRIPT
stratification) above pool surface takes place. Generally, at the wind speed of 5 m/s the Britter’s

criterion is less, or slightly greater than the critical value, whereas at low wind speeds and

extended pools it significantly exceeds the critical value. The values of Richardson number are

also consistent with the numerical calculations. The Ric in most cases is in the range from 1 to 30

indicating the prevalence of stably stratified shear flow. The Ric < 1 is obtained only for the case

PT
where the pool sizes are 10 m× 10 m and the wind speed is 5 m/s and Ric> 30 is obtained when

the pool sizes are 100 m×100 m and the wind speed is 1 m/s.

RI
4. Conclusion

SC
The validation of the developed pool evaporation CFD model against both literature and

our own experimental data is performed. A good agreement is obtained between the measured

U
and the predicted evaporation rates. Based on the proposed model a sensitivity analysis is
AN
conducted to determine the effect of pool sizes on evaporation of a volatile liquid (hexane). The

numerical analysis has shown that the evaporation rate falls as the pool sizes increase due to the
M

enhancement of buoyancy effects above the pool surface. The importance of the pool sizes effect
D

decreases as the air velocity increases. It was found that in the dependence of evaporation rate on

pool size Jg,s∼Ln the exponent on pool size varies between -0.553 and -0.0578 for wind speeds
TE

between 1 and 5 m/s, while in risk analysis handbooks it is generally accepted that Jg,s is
EP

proportional to L-0.11 regardless of wind speed.


C

Nomenclature:
AC

∆B roughness function;

CP specific heat, J/(kg⋅K);

Cs roughness constant;

Cµ coefficient in the turbulence model;

Dm molecular diffusion coefficient, m2/s;

E empirical constant equal to 9.1;


25
ACCEPTED MANUSCRIPT
∆H heat of vaporization, J/kg;

Jg,s mass flux of vapor from the pool surface due to evaporation, kg/(m2⋅s);

K correction factor for Stefan flow;

Ks roughness height, m;

Ks+ non-dimensional roughness height;

PT
k turbulent kinetic energy, m2/s2;

L characteristic pool size, m;

RI
mliq mass of the liquid per unit surface area of the pool, kg/m2;

SC
Pg(Tliq) vapor pressure of evaporating component at liquid temperature Tliq;

PC function which takes into account the resistance of the diffusion sublayer to mass

U
transfer;
AN
Sc and Sct molecular and turbulent Schmidt numbers, respectively;

T absolute temperature, K;
M

T+ non-dimensional temperature;
D

t time, s;
TE

U wind velocity, m/s;

u+ non-dimensional velocity;
EP

Yg mass fraction of the evaporating component in the gas phase, kg/kg;

Y+ non-dimensional mass-fraction;
C

y+ non-dimensional distance;
AC

yC+ non-dimensional diffusion sublayer thickness;

yP normal distance from the pool surface to the neighboring node of the computational grid;

z0 aerodynamic roughness, m.

Greek symbols:
26
ACCEPTED MANUSCRIPT
ε 2 3
turbulence kinetic energy dissipation rate, m /s ;

κ von Karman constant equal to 0.41;

λ coefficient of thermal conductivity, W/(m⋅K);

µ coefficient of molecular viscosity, kg/(m⋅s);

µt coefficient of turbulent viscosity, kg/(m⋅s);

PT
ρ density, kg/m3.

RI
Subscripts:

SC
no index vapor-air mixture;

a air;

g evaporating component;
U
AN
grd ground;

liq pool;
M

P centroid of the wall-adjacent cell;


D

s surface of pool.
TE

References
EP

Bird, R. B., Stewart, W. E., & Lightfoot, E. N. (1960) Transport Phenomena. Wiley & Sons,

New York.
C

Blocken, B., Stathopoulos, T., & Carmeliet, J. (2007). CFD simulation of the atmospheric
AC

boundary layer: wall function problems. Atmos. Environ., 41, 238-252.

Brighton, P.W.M. (1987). Evaporation from a plane liquid surface into a turbulent boundary

layer. SRD/HSE-report R375, UK Atomic Energy Authority, Safety and Reliability Directorate.

Britter, R. E. (1989). Atmospheric dispersion of dense gases. Annual Rev. Fluid Mech., 21, 317-

344.
27
ACCEPTED MANUSCRIPT
Cebeci, T., & Chang, K.C. (1978). Calculation of incompressible rough-wall boundary-layer

flows. AIAA Journal, 16, 730.

Churchill, S. W. (1976). A comprehensive correlating equation for forced convection from flat

plates. AIChE Journal, 22, 264-268.

Coulson, J. M., & Richardson, J. F. (1993). Chemical Engineering (Fluid Flow, Heat Transfer

PT
and Mass Transjk). (4th ed.). Pergamon Press, Oxford.

Desoutter, G., Habchi, C., Cuenot, B., & Poinsot, T. (2009). DNS and modeling of the turbulent

RI
boundary layer over an evaporating liquid film. Int. J. Heat and Mass Transf., 52, 6028–6041.

SC
Fluent (2006). Fluent 6.3 user’s guide. Lebanon, New Hampshire, USA: Fluent Inc.

Galeev, A. D., Starovoytova, E. V., & Ponikarov, S. I. (2013). Numerical simulation of the

U
consequences of liquefied ammonia instantaneous release using FLUENT software. Process Saf.
AN
Environ. Prot., 91, 191–201.

Galeev, A.D., Salin, A.A., & Ponikarov, S.I. (2013). Consequence analysis of aqueous ammonia
M

spill using computational fluid dynamics. J. Loss Prev. Process Ind., 26, 628–638.

Galeev, A.D., & Ponikarov, S.I. (2014). Numerical analysis of toxic cloud generation and
D

dispersion: A case study of the ethylene oxide spill. Process Saf. Environ. Prot., 92, 702–713.
TE

Gilliland, E. R. (1934). Diffusion Coefficients in Gaseous Systems. Ind. Eng. Chem., 26, 681-
EP

685..

Gorle, C., van Beeck, J., Rambaud, P., & van Tendeloo, G. (2009). CFD modeling of small
C

particle dispersion: The influence of the turbulence kinetic energy in the atmospheric boundary
AC

layer. Atmos. Environ., 43, 673-681.

Green, D.W., & Perry, R.H. (2008). Chemical Engineer’s Handbook. (8th ed.). McGraw-Hill.

Habib, A., Schalau, B., Acikalin, A., & Steinbach, J. (2009). Transient calculation of the

boundary layer flow over spills. Chem. Eng. Technol., 32, 306–311.

Havens, J.A. (1992). Review of dense gas dispersion field experiments. J. Loss Prev. Process

Ind., 5, 28–41.
28
ACCEPTED MANUSCRIPT
Incropera, F.P., DeWitt, D.P., Bergman, T.L., & Lavine, A.S. (2006). Fundamentals of Heat and

Mass Transfer. Wiley.

Kapias, T. & Griffiths, R. F. (1999). Modelling the behaviour of spillages of sulfur trioxide and

oleum. Health and Safety Executive, Contract Research Report 217/1999.

Kawamura, P.I., & Mackay, D, (1987). The evaporation of volatile liquids. J. Hazard. Mater. 15,

PT
343-364.

Khajehnajafi, S., & Pourdarvish, R. (2011). Correlations for mass transfer from a liquid spill:

RI
Comparisons and recommendations. Process Saf. Prog., 30, 178–184.

SC
Labovsky, J., Jelemensky, L. (2011). Verification of CFD pollution dispersion modelling based

on experimental data. J. Loss Prev. Process Ind., 24, 166-177.

U
Liu, Y., Olewski, T., Vechot, L., Gao, X., & Mannan, S. (2011). Modelling of a cryogenic liquid
AN
pool boiling using CFD code. Proc. 14th Annual Symposium, Mary Kay O’Connor Process

Safety Center “Beyond Regulatory Compliance: Making Safety Second Nature”, Texas A&M
M

University, College Station, Texas, USA, 25-27 October 2011.

Mackay, D., & Matsugu, R.S. (1973). Evaporation rates of liquid hydrocarbon spills on land and
D

water. The Can. J. Chem. Eng., 51, 434-439.


TE

Mokhtarzadeh-Dehghan, M.R., Akcayoglu, A., & Robins, A.G. (2012).. Numerical study and
EP

comparison with experiment of dispersion of a heavier-than-air gas in a simulated neutral

atmospheric boundary layer. J. Wind Eng. Ind. Aerodyn., 110, 10-24.


C

Nielsen, F, Olsen, E., & Fredenslund, A. (1995). Prediction of isothermal evaporation rates of
AC

pure volatile organic compounds in occupational environments — A theoretical approach based

on laminar boundary layer theory. The Ann. Occup. Hyg., 39, 497-511

Parente, A., Gorlé, C., van Beeck, J., & Benocci, C. (2011). Improved k–ε model and wall

function formulation for the RANS simulation of ABL flows. J. Wind Eng. Ind. Aerodyn., 99,

267-278
29
ACCEPTED MANUSCRIPT
Pasquill, F. (1943). Evaporation from a plane, free liquid surface into a turbulent air stream.

Proc. R. Soc. Lond. A, 182, 75-95.

Patankar, S. V. (1980). Numerical Heat Transfer and Fluid Flow. Hemisphere, Washington, DC.

Raimundo, A.M., Gaspar, A.R., Oliveira, A.V.M., & Quintela, D.A. (2014). Wind tunnel

measurements and numerical simulations of water evaporation in forced convection airflow. Int.

PT
J. Thermal Sciences, 86, 28–40

Richards, P., & Hoxey,R.(1993).Appropriate boundary conditions for computational wind

RI
engineering models using the k–e turbulence model. J. Wind Eng. Ind. Aerodyn., 46–47,145–

SC
153.

Robins, A., Castro, I.,Hayden, P., Steggel, N., Contini, D., & Hesit, D. (2001).A wind tunnel

U
study of dense gas dispersion in a neutral boundary layer over a rough surface. Atmos. Environ.,
AN
35, 2243–2252.

Rong, L., Nielsen, P.V., & Zhang, G.Q. (2010). Experimental and numerical study on effects of
M

airflow and aqueous ammonium solution temperature on ammonia mass transfer coefficient. J.

Air Waste Manag. Assoc., 60, 419–428.


D

Rong, L., Elhadidi, B., Khalifa, H.E., Nielsen, P.V., & Zhang, G.Q. (2011). Validation of CFD
TE

simulation for ammonia emissions from an aqueous solution. Comput. Electron. Agric., 75, 261–
EP

271.

Saha, C. K., Wu, W., Zhang, G., & Bjerg, B. (2011). Assessing effect of wind tunnel sizes on air
C

velocity and concentration boundary layers and on ammonia emission estimation using
AC

computational fluid dynamics (CFD). Comput. Electron. Agric, 78, 49–60

Schlichting, H. (1968) Boundary-layer Theory. (6th ed.).. McGraw-Hill, New York.

Shih, T.H., Liou, W.W., Shabbir, A., & Zhu, J. (1995). A new k-ε eddy-viscosity model for high

Reynolds number turbulent flows – model development and validation. Comput. Fluids, 24, 227-

238.
30
ACCEPTED MANUSCRIPT
Troen, I., & Petersen, E. L (1989). European Wind Atlas. Risø National Laboratory, Roskilde,

Denmark.

van den Bosch, C.J.H. (1997). Pool evaporation. In C.J.H. van den Bosch, & R.A.P.M.

Weterings (Eds.), CPR 14E, Methods for the calculation of physical effects due to releases of

hazardous materials (liquids and gases), Yellow Book (pp. 3.1 – 3.128). Committee for the

PT
Prevention of Disasters, The Hague.

Véchot, L., Olewski, T., Osorio, C., Basha, O., Liu, Y., & Mannan, M. S. (2013). Laboratory

RI
scale analysis of the influence of different heat transfer mechanisms on liquid nitrogen

SC
vaporization rate. J. Loss Prev. Process Ind., 26, 398–409.

Vik, T., & Pettersson Reif, B.A. (2011). Implementation of a new and improved evaporation

U
model in Fluent, FFI-rapport 2011/00116, Norwegian Defence Research Establishment (FFI).
AN
Yang, Y., Gu, M., Chen, S., & Jin, X. (2009). New inflow boundary conditions for modeling the

neutral equilibrium atmospheric boundary layer in computational wind engineering. J. Wind


M

Eng. Ind. Aerodyn., 97, 88-95.


D
TE
C EP
AC
31
ACCEPTED MANUSCRIPT
Fig. 1. Profiles of (a) velocity U, (b) turbulence kinetic energy k, (c) turbulence kinetic energy dissipation

rate ε and (d) turbulent viscosity µt from numerical calculation (solid line) and from Yang et al. (2009)

(dashed line) at the wind velocity of 1 m/s at the height of 10 m

Fig. 2. Profiles of (a) velocity U, (b) turbulence kinetic energy k, (c) turbulence kinetic energy dissipation

rate ε and (d) turbulent viscosity µt from numerical calculation (solid line) and from Yang et al. (2009)

(dashed line) at the wind velocity of 2.5 m/s at the height of 10 m

PT
Fig. 3. Profiles of (a) velocity U, (b) turbulence kinetic energy k, (c) turbulence kinetic energy dissipation

rate ε and (d) turbulent viscosity µt from numerical calculation (solid line) and from Yang et al. (2009)

RI
(dashed line) at the wind velocity of 5 m/s at the height of 10 m

SC
Fig. 4. Measured wind velocity

Fig. 5. Measurements (solid line) and predictions (dashed line) of liquid mass and temperature

U
Fig. 6. Dependencies of specific evaporated mass (left) and average liquid temperature (right)
AN
against time at wind velocities of 1 m/s (a), 2.5 m/s (b) and 5 m/s (c)

Рис. 7. Dependencies of average relative turbulent viscosity (left) and driving force (right) at
M

wind speeds of 1 m/s (a), 2.5 m/s (b) and 5 м/с (c).
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
Table 1

Experimental data and predictions for ethanol evaporation

Pool temperature, Wind velocity at 2 Mass flow, g/s


К m height, m/s Experiment Mackay – CFD model
Matsugu Eq.
Flat terrain
310.15 1.4 0.456 0.681 0.393
309.65 1.9 0.533 0.855 0.523

PT
325.15 1.7 1.02 1.59 1.06
Very rough terrain
305.65 1.4 0.488 0.663 0.536

RI
310.15 1.8 0.679 0.99 0.9
324.65 1.8 1.32 1.95 1.96

SC
Table 2

U
Experimental data and predictions for cyclohexane evaporation
AN
Pool temperature, Wind velocity at 2 Mass flow, g/s
К m height, m/s Experiment Mackay – CFD model
Matsugu Eq.
M

Flat terrain
303 2.71 1.08 1.74 1.11
310 3.05 1.63 2.49 1.66
D

317 3.49 2.33 3.63 2.55


Very rough terrain
TE

303 1.71 0.967 1.39 1.15


310 1.62 1.35 1.80 1.46
317 1.4 1.67 2.13 1.68
EP

Table 3
C

Physical properties of gases and liquid


AC

Physical property Temperature, K Variation, %


286 308
Dm,g, m2/sec 6.48⋅10 -6
7.24E-06 11,7
µg, kg/(m⋅sec) 6.5⋅10 -6
6.7⋅10-6 3,1
µa, kg/(m⋅sec) 1.78⋅10-5 1.88⋅10-5 5,6
CP.g, J/(kg⋅К) 1590 1690 6,3
CP.a, J/(kg⋅К) 1005 1008 0,3
CP.liq, J/(kg⋅К) 2215 2310 4,3
λg, W/(m⋅К) 0.0115 0.0135 17,4
λa, W/(m⋅К) 0.024 0.026 8,3
∆Нg, J/kg 380000 362500 4,6
ACCEPTED MANUSCRIPT
Table 4

Values of B and Ric depending on pool sizes and wind velocity

Pool sizes Britter’s criterion B (Richardson number Ric)


1 m/s 2.5 m/s 5 m/s
10 m×10 m 0.35 (12.4) 0.20 (2.2) 0.12 (0.5)
30 m×30 m 0.42 (20.3) 0.27 (5.5) 0.17 (1.3)
60 m×60 m 0.46 (26.7) 0.31 (8.4) 0.21 (2.6)
100 m×100 m 0.51 (35.9) 0.33 (10.3) 0.25 (4.1)

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT
• The results of the validation of the developed pool evaporation model are presented;
• A sensitivity analysis is conducted to determine the effect of pool sizes on evaporation of
a volatile liquid (hexane);
• The importance of the pool sizes effect decreases as the air velocity increases.

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like