Sunteți pe pagina 1din 17

Computing (2013) 95 (Suppl 1):S303S318

DOI 10.1007/s00607-013-0292-6

Numerical simulation of transonic flow of wet steam


in nozzles and turbines

Jan Halama Jaroslav Fort

Received: 30 September 2012 / Accepted: 4 January 2013 / Published online: 18 January 2013
Springer-Verlag Wien 2013

Abstract The paper presents several modifications of the flow model published in
tastn and ejna (Proceedings of the 12th international conference of the properties
of water and steam, Begel House, pp 711719, 1995). Modifications related to the
droplet growth model and the equation of state and their implementation into numerical
code are described. The effect of droplet size spectra of incoming wet steam is also
discussed. Numerical results of three-dimensional flow of wet steam in a turbine
cascade show the effect of surface tension correction.

Keywords Finite volume method Nucleation Wet steam Droplet size distribution

Mathematics Subject Classification 76H05 76T10 65M08

1 Introduction

The paper is aimed at the numerical solution of internal aerodynamics problems of


accelerating two-phase flow of steam in nozzles and turbines. The acceleration of such
flow is typically very high. It leads to non-equilibrium condensation, i.e. the vapor
starts to condensate later after certain sub-cooling below the saturation temperature.
We use a flow model based on the conservation laws for mass, momentum and energy

J. Halama J. Fort
Department of Technical Mathematics, FME CTU Prague,
Karlovo nm. 13, 121 35 Prague, Czech Republic
J. Fort
e-mail: Jaroslav.Fort@fs.cvut.cz

J. Halama (B)
Institute of Thermomechanics, AS CR, Dolejkova 5, 182 00 Prague 8, Czech Republic
e-mail: halama@marian.fsik.cvut.cz

123
S304 J. Halama, J. Fort

of the whole mixture and on the conservation laws for the moments of droplet spectra.
Such model is able to distinguish mono- and poly-dispersed mixtures. Our original flow
model (which is further denoted as AVG-P model) is based on the model published
in [1]. The vapor specific heat ratio, considered as a constant in [1], is taken as a
function of vapor temperature. The homogenous nucleation model is supplemented
by the correction term published in [2]. The model is closed by polynomial functions
for the material properties of water taken from [3]. Numerical method based on this
(AVG-P) flow model was verified for the case of flow in a nozzle [4]. The droplet
growth speed considered in the AVG-P model is based on the average droplet size
only. This paper discusses a modification of the droplet growth model which relies
on the apriori given droplet size distribution function reconstructed from the values
of moments according to [5]. This model modification is further denoted as DSDF-P
model. Recent simulation of the unsteady flow in the LP turbine stage addressed two
additional issues, the first one related to the stability of droplet growth model for the
parameters close to the saturation line and the second one related to the numerical
prediction of the inlet flow parameters for wet steam. The stability of the droplet
growth model have been solved by regularization. The inlet boundary condition for
the wet steam requires the complete information about the structure of incoming liquid
phase. Tests for flow in a nozzle show that it is relatively easy to estimate wetness
by numerical simulation of upstream expansion, however it seems that the values
of moments depends strongly on the used flow model and numerical method. It is
therefore questionable to use numerical results of upstream expansion to approximate
the parameters of liquid phase at the inlet boundary. We discuss the sensitivity of
numerical results on different flow model modifications as well as on used numerical
method for the simple case of steady flow in a nozzle with zero wetness at the inlet.
We also discuss the sensitivity of a numerical solution on the differences in values
of moments at the inlet. Our final numerical examples of three-dimensional flow in a
turbine cascade show the effect of the surface tension correction.

2 Model of wet steam flow

Consider the flow of steam with velocity, pressure and temperature within the range
typical for the flow in a low pressure part of a steam turbine. The rapid expansion of
steam leads to non-equilibrium phase change. The nucleation occurs when the steam
temperature drops sufficiently below the saturation temperature. This sub-cooling
is typically 3040 K. Assume the condensed water dispersed in vapor in the form of
high amount of small spherical droplets, which have the same velocity as the vapor.
The flow model is based on the conservation of mass, momentum and energy for the
mixture and the transport equation for the mass fraction of liquid phase. Such kind of
model, known as mixture model, is commonly used, e.g. [68] or [9]. Mass exchange
between vapor and droplets is given by nucleation and droplet growth models. One
of the important parameters for droplet growth model is the droplet size. The whole
spectra of droplet sizes in the elemental volume of mixture is often approximated
by an average radius, therefore the mentioned set of transport equations has to be
complemented at least by the transport equation for the number of droplets, e.g. [6].

123
Numerical simulation of transonic flow S305

The presented model contains three additional transport equations for the moments
according to [10] to be able, e.g., to distinguish between mono-dispersed and poly-
dispersed mixture for the inlet boundary condition or to be able to add some simple
model of droplet size distribution. The moments read:


N 
N
Q0 = N , Q1 = ri , Q2 = ri 2 , (1)
i=1 i=1

ri is the radius of ith


where N denotes the total number of droplets per unit mass and
droplet. The average droplet radius is approximated by r = Q 2 /Q 0 if the wetness
is above a chosen minimum min otherwise it is set r = 0 to avoid numerical errors.
The full system of transport equations reads

W F(W) G(W) H(W)


+ + + = Q, (2)
t x y z

u v
u u 2 + p vu
2
v uv v + p

w uw vw

W=
e , F = (e + p)u , G = (e + p)v ,

u v

Q2 Q2u Q2v

Q1 Q1u Q1v
Q0 Q0u Q0v

0

w 0
wu

wv 0

w 2 + p
0

H=
(e + p)w , Q =

2 ,
w 4
rc Jl + 4 0 r N (r )r(r )dr l
3
3
Q2w

rc J + 2 0 r N (r )r(r )dr
2
Q1w


rc J + 0 N (r )r(r )dr
Q0w
J

where , u, v, w and e denote the density, velocity components and total energy per
unit volume for the mixture, denotes the mass fraction of liquid phase, known also
as wetness. J is the number of new droplets due to homogenous nucleation per unit
volume per second, which have the size of critical radius rc . The variable l denotes
the density of liquid phase, N (r ) denotes the droplet size distribution function. The
integral of N (r ) along the interval  r1 , r2  gives the total number of droplets with
the radius between r1 and r2 per unit mass of mixture. The function r (r ) gives the
droplet growth velocity of droplet with radius r . The integrals in the source term Q are

123
S306 J. Halama, J. Fort

approximated by a simple quadrature rule with appropriately chosen lower and upper
limits. Under the perfect gas assumption the system is closed by the equation

( 1)(1 ) 1
p= e (u 2 + v 2 + w 2 ) + L , (3)
1 + ( 1) 2

where common pressure for both phases is considered and L denotes the latent heat of
condensation/evaporation Further details are provided in [7] or [4]. The specific heat
ratio is taken as a function of temperature

c p (T )
= , (4)
c p (T ) Rv

where Rv is the gas constant for vapor, c p the specific heat at constant pressure c p and
T the vapor temperature. The system (2) models also the single-phase flow of vapor if
= Q 0 = Q 1 = Q 2 = 0. We assume two approximations of droplet growth speed
r (r ). The first one considers that the droplet growth speed depends only on average
radius r . The integrals in the source term Q are then replaced by a simple product

 
r i N (r )r(r )dr = r (r ) r i N (r )dr = r (r )Q i . (5)
0 0

Such model is denoted as AVG-P model. The second approximation considers cer-
tain reconstruction of the droplet size distribution function N (r ) from the values of
moments Q i . The model further denoted as DSDF-P is based on the log-normal dis-
tribution for N (r ), which can be obtained from three moments Q 0 , Q 1 and Q 2 , for
details see [5]

1 ln2 (r/r g )
N (r ) = Q 0 exp , (6)
r ln(g ) 2 2 ln2 (g )

where

r
rg = , g = exp ln(cv2 + 1),
exp(0.5 ln2 (g ))
  (7)
Q2 Q0 Q2
r= , cv = 1.
Q0 Q 21

This paper contain also some numerical results of the flow in a nozzle, which have
been performed by a modification of the presented flow model, where the equation of
state for perfect gas, which is considered in the models [1] and [4], has been replaced
by the virial equation of state

123
Numerical simulation of transonic flow S307

p  
= Rv 1 + Bv + Cv2 , (8)
v T
Rv = 461.52, v = (1 ),
B = 1 (1 + B /)1 + 2 e B (1 e B )5/2 + 3 B ,
1500
B = , = 10000.0, 1 = 0.0015, 2 = 0.000942, 3 = 0.0004882,
T
C = 1 (C 0 )eC + 2 ,
T
C = , 0 = 0.8978, = 11.16, 1 = 1.772, 2 = 0.0000015,
647.286
taken from [9]. The variables T and v denote the temperature and the density of vapor
respectively, the pressure p is considered the same for both phases. Models with this
equation of state are further denoted as the AVG-V and the DSDF-V models.
The number of new droplets due to homogenous nucleation per second and per unit
volume is computed using a classical formula [11]
  
2 v2 4rc2
J= exp , (9)
m 3v l 3k B T

where the surface tension of water is a function of temperature and is the correction
coefficient proposed in [2]

= 1.328 pcor
0.3
, pcor = psaturated (s01 )105 , (10)

where s01 is the entropy corresponding to the inlet total conditions and pcor is con-
sidered in [bar ] unit. This correction is applied for the AVG-P and DSDF-P models
based on the perfect gas assumption in order to match the onset of nucleation. This
correction is not needed for models AVG-V and DSDF-V. The new droplet appearing
due to homogenous nucleation has the initial radius equal to the critical radius

2
rc = , (11)
Ll ln(Ts /T )

where L(T ) is the latent heat of condensation/evaporation and l (T ) denotes the


density of water. The saturation temperature Ts is evaluated according to IAPWS-IF97
formulation. An existing droplet is growing or it can also evaporate depending on the
vapor temperature. We consider the Gyarmathys droplet growth model

v (Ts T ) r rc v 2 Rv T
r (r ) = , Kn = , (12)
Ll (1 + 3.18 K n) r2 4r p

where vapor thermal conductivity v and vapor viscosity v are functions of tem-
perature and r is the droplet radius. The formula for the droplet growth speed r (r )
should not be directly implemented into numerical codes, since rc goes to infinity for
vapor temperature approaching the saturation temperature. This growth is compen-
sated by the term (Ts T ), which is going to zero at the same time, so r (r ) remains

123
S308 J. Halama, J. Fort

finite, however direct implementation of the Eq. (12) is dangerous, since the product
of almost infinity with almost zero has unpredictable behavior within numerical
code. Therefore we implement the formula (13)

v (Ts T ) r rc
r (r ) =
Ll (1 + 3.18 K n) r 2
 
v Ts T 2 Ts T
= , (13)
Ll (1 + 3.18 K n) r Ll r 2 ln Ts
T

Ts T
where the term (zero divided by zero for T going to Ts ) is approximated
ln TTs
using Taylor expansion

Ts
Ts T 1 1 1 1 19 4 3 5
Ts
=T T
Ts
= 1 + 2 + 3 + ,
ln ln 2 12 24 720 160
T T
Ts
= 1. (14)
T

This implementation of droplet growth formula yields a stable numerical algorithm


also for cases with steam parameters close to saturation line.

3 Numerical solution

The numerical method has to cover very different time scales of convection, nucleation
and droplet growth. Our numerical method is therefore based on the symmetrical
splitting method used in [12], which allows separate solution of each phenomena by
individual numerical method. The original problem is approximated by the solution
of three successive sub-problems

W
= Q(W ) (15)
t
W F(W ) G(W )
= (16)
t x y
W
= Q(W ), (17)
t

where the Eq. (15) is solved with initial data W (t) = W(t). The advance t/2 in
time for the solution of (15) yields the initial data for the Eq. (16), i.e. W (t) =
W (t + t/2). Then the shift t in time for the solution of (16) yields the initial data
for the Eq. (17), i.e. W (t) = W (t + t). The time shift t/2 for the solution of
(17), i.e. W (t +t/2), corresponds to the time shift t for the solution W(t +t) of
the original un-split system (2). The single step of Lax-Wendroff finite volume method
is applied for (16) and several steps of Runge-Kutta method evaluate the source term
contribution in (15) and (17). The algorithm of the splitting method reads

123
Numerical simulation of transonic flow S309

 
( j+1) ( j) t
WK = RK W K , , j = 0, . . . , m 1
2m
(m+1) (m)
WK = FV(W K , t) (18)
 
( j+1) ( j) t
WK = RK W K , , j = m + 1, . . . , 2m
2m

where W(0) K = WK , WK
n n+1
= W(2m+1)
K , FV(WnK , t) denotes one step of Lax
Wendroff method with initial data W K and time step t (subscript K denotes the
n

K -th cell and superscript n the n-th time level). Similarly RK(WnK , t) denotes one
step of the Runge-Kutta method. The local number of sub-steps is m = t/ , where
is the time scale of condensation and t is the time step of the finite volume method
for the Eq. (16).
Due to the stability problems of the fully explicit method in three-dimensional case,
the explicit method for the parts (15) and (17) has been replaced by the implicit Euler
method. The nonlinear equation is solved by iterative method.

(0)
W K = WnK
( j+1) (0) t ( j)
WK = WK + Q(W K ), j = 0, . . . , M 1
2
(M+1) (M)
WK = FV(W K , t) (19)
( j+1) (M+1) t ( j)
WK = WK + Q(W K ), j = M + 1, . . . , L
2
(L+1)
Wn+1
K = WK .

4 Wet steam flow in a nozzle

Consider three cases of transonic flow (subsonic velocity at the inlet and the supersonic
velocity at the outlet) of steam in the Barschdorff nozzle, see the Fig. 1. All cases are
defined by the Dirichlet boundary conditions at the inlet, see the Table 1. The outlet
static pressure corresponds to the design conditions of the nozzle, therefore it is a
result of computation. The results below have been obtained by the 1D version of the
fully explicit code, taking into account the varying cross-sectional area, for transport
equations see the Appendix. In our experience results of 1D code are in this case in a
good agreement with results of 2D simulation. The main reason of choosing the 1D
code is to focus on the flow model and to avoid additional numerical errors caused by
multi-D domain discretization. The stopping criterion for the time marching method
is the value of time derivative of W, which goes to machine zero for the 1D code.
The domain has been discretized using 200 cells along the axis, the grid spacing in
the convergent-divergent part of the domain is equal to 1.4 103 and the time step
has been set to t = 1.2 107 .
The case BAR is the case published in [13]. There it is dry steam at the inlet and
the nucleation starts downstream the throat. Figure 2 shows the numerical results
achieved by all models. The correct position of nucleation start for models based on

123
S310 J. Halama, J. Fort

Fig. 1 The domain of solution for Barschdorff nozzle [13], i denotes the inlet, o the outlet and w the
wall boundary

Table 1 Inlet boundary conditions for considered nozzle flow cases

Case ptotal,inlet Ttotal,inlet Q0 Q1 Q2 inlet

BAR 78390 373.1 0 0 0 0


MONO 21061 334.3 1.26 1017 5.08 109 2.06 102 0.035
POLY 21061 334.3 5.55 1016 1.59 109 9.09 101 0.035

the perfect gas equation of state (the models AVG-P and DSDF-P) is achieved thanks
to the correction (10). This correction is not applied for the models based on the virial
equation of state (8), i.e. for the models AVG-V and DSDF-V. All models yield nearly
the same distribution of pressure and wetness, see the Fig. 2. The best agreement
with experimental data is achieved by the model DSDF-V. The models AVG-P and
DSDF-P do not reach the local maximum of pressure at x = 0.04 m. Nucleation starts
for smaller x-values for the AVG-V model.
Significant difference between results achieved by all models can be observed in
the average droplet size prediction, where the DSDF-V model yields the smallest and
the AVG-V the biggest droplets. The results show relation between the average droplet
size and the size of the pressure rise in the nucleation zone. Bigger pressure rise is most
probably linked with higher nucleation rate J at the start of the nucleation, i.e. the
mixture contains higher amount of smaller droplets. The average droplet size depends
also on the position of the nucleation start, later start means higher subcooling of the
vapor and therefore it is linked to smaller droplets. These observations are also consis-
tent with numerical results published in [14], where earlier nucleation start, compared
to our results, yields the outlet average droplet radius around 7 108 m. Similar
effect has the amount of numerical dissipation introduced by numerical method FV
used for the convection step (16). More dissipative methods yield smaller magnitude
of pressure jump and the mixture therefore contains smaller amount of bigger droplets
(i.e. bigger average radius), for details see [4], where two different methods for the
AVG-P model produced the outlet average droplet radius around 2.5 108 m and
3.5108 m. Also an interesting observation is as follows. Additional grid refinement,
which reduces numerical dissipation, leads to smaller droplets and has practically no
effect on wetness. The wetness is relatively insensitive to all presented modifications
of the flow model as well as of the numerical method, i.e. numerical simulation can
provide relatively reliable prediction for wetness. It seems the droplet size prediction
unfortunately strongly depends on the used model, method, grid etc.

123
Numerical simulation of transonic flow S311

50000

40000
pressure [Pa]

30000 model AVG-P


model AVG-V
model DSDF-P
model DSDF-V
experiment

20000
0 0.02 0.04 0.06 0.08 0.1 0.12
x [m]
(a)

0.06

0.05
model AVG-P
model AVG-V
0.04 model DSDF-P
model DSDF-V
wetness

0.03

0.02

0.01

0
-0.2 -0.1 0 0.1
x [m]
(b)

3e-08

model AVG-P
model AVG-V
average radius [m]

model DSDF-P
2e-08 model DSDF-V

1e-08

0
-0.2 -0.1 0 0.1
x [m]
(c)
Fig. 2 The results for the case BAR achieved by all presented models, experimental data from [13]

123
S312 J. Halama, J. Fort

N(r)

0 5e-08 1e-07 1.5e-07 2e-07


droplet radius [m]

Fig. 3 The droplet size distribution at the inlet for the case POLY

The other two cases MONO and POLY have the same non-zero wetness and
the same average radius 4 108 m at the inlet. The case MONO considers the
mono-dispersed mixture. The droplet size distribution at the inlet for the case POLY
(poly-dispersed mixture at the inlet) has been artificially created by the log-normal
distribution, see the Fig. 3.
Figure 4 presents the results achieved by the AVG-P as well as DSDF-P mod-
els. There can be seen almost no difference for the pressure distribution for three
case/model combinations. The differences in wetness are below 0.005 at the outlet.
The most significant effect is clearly visible for droplet size distribution. If we com-
pare the results for both MONO and POLY cases achieved by AVG-P model (results
of computations which differ only in values of Q 0 , Q 1 and Q 2 at the inlet), we see,
that the POLY case yields bigger droplets in any location along nozzle. The droplet
size for the case POLY obtained from DSDF-P model gives the smallest droplets,
we even see some decrease of average droplet size in the front part of the nozzle. It
is most probably caused by the decrease of surface for droplets smaller than critical
size, which dominates over the surface increase for droplets bigger than critical size.
Decreasing total surface Q 2 together with constant number of droplets Q 0 yields the
average size drop. Presented results refer to the sensitivity of wetness and droplet
size on the structure of liquid phase at the inlet (compare MONO and POLY cases
computed by AVG-P model). The distribution of pressure seems to be insensitive.

5 Wet steam flow in a cascade

The results of three-dimensional two-phase flow in a stator cascade have been obtained
by the numerical method with the implicit Euler method for the integration of source
term and the AVG-P flow model. The computational grid consists of the single block
of structured hexahedral grid with 91x25x18 points in downstream, tangential and
span-wise directions, see the Fig. 5. We consider the constant inlet total temperature

123
Numerical simulation of transonic flow S313

20000

15000
pressure [Pa]

10000
case MONO, model AVG-P
case POLY, model AVG-P
case POLY, model DSDF-P

5000

-0.2 -0.1 0 0.1


x [m]
(a)

0.09

0.08

case MONO, model AVG-P


0.07 case POLY, model AVG-P
case POLY, model DSDF-P
wetness [-]

0.06

0.05

0.04

0.03
-0.2 -0.1 0 0.1
x [m]
(b)
6.5e-08

6e-08

case MONO, model AVG-P


average radius [m]

5.5e-08 case POLY, model AVG-P


case POLY, model DSDF-P

5e-08

4.5e-08

4e-08
-0.2 -0.1 0 0.1
x [m]
(c)
Fig. 4 The results for the cases MONO and POLY achieved by the AVG-P and DSDF-P models

123
S314 J. Halama, J. Fort

Fig. 5 Computational grid for 3D stator cascade, several blade channels are plotted, computation is
performed using only one blade channel

T01 = 336 K, the constant inlet total pressure p01 = 37534 Pa, zero inlet pitch angle,
linear distribution of yaw angle between hub and tip casings and dry meta-stable steam
at the inlet, i.e. zero inlet wetness and zero Hills moments Q i at the inlet. The outlet
static pressure is a function of radial coordinate and is constant in circumferential
direction. The effect of surface tension correction (10) is noticeable from compari-
son between the results of the original method (without correction) and the modified
method (with correction). The results achieved without this correction show more
pronounced sub-cooling of steam before start of nucleation (Figs. 6, 7).

6 Conclusions

Numerical method has been successfully extended for the cases with non-zero wet-
ness of incoming steam. The droplet growth model has been regularized. Numerical
simulation of the case BAR for the nozzle yields the values of wetness practically inde-
pendent of the used flow model, while the computed droplet sizes and thus moments
differ from model to model. Another comparison between the mono-dispersed MONO
and poly-dispersed POLY cases (which both have the same average size of droplet
at the inlet and the same inlet wetness) shows non-negligible differences in wetness
and average droplet size, even for AVG-P flow model (i.e. difference in results caused
only by difference in values of Q 0 , Q 1 and Q 2 at the inlet). This demonstrates the
importance of transport equations for the moments even in the AVG-P model, because

123
Numerical simulation of transonic flow S315

Fig. 6 Wetness contours

123
S316 J. Halama, J. Fort

Fig. 7 Contours of sub-cooling (vapor temperature minus saturation temperature)

123
Numerical simulation of transonic flow S317

for the same inlet wetness and the same average droplet size at the inlet we get different
results.
The results for three-dimensional flow in a stator cascade show the effect of surface
tension correction.

Acknowledgments This work has been supported by the grants No. 101/11/1593 and No. 201/08/0012
of the Grant Agency of the Czech Republic.

Appendix

The system of transport equations for the one-dimensional case with the variable
cross-sectional area A(x) reads

(A(x)W) (A(x)F)
+ = P + A(x)Q, (20)
t
x
u 0
u u 2 + p p A (x)

e (e + p)u 0

W=
, F = u , P = 0 ,

Q2 Q2u 0

Q1 Q1u 0
Q0 Q0u 0

0

0

0


4 3 2
Q = 3 rc Jl + 4 0 r N (r )r(r )dr l .


rc2 J + 2 0 r N (r )r(r )dr



rc J + 0 N (r )r(r )dr
J

References

1. tastn M, ejna M (1995) Condensation effects in transonic flow through turbine cascade. In: Pro-
ceedings of the 12th international conference of the properties of water and steam, Begel House,
pp 711719
2. Petr V, Kolovratnk M (2001) Heterogenous effects in the droplet nucleation process in LP steam
turbines. In: 4th European conference on turbomachinery, Firenze
3. Heiler M (1999) Instationre Phnomene in homogen/heterogen kondensierenden Dsen- und Tur-
binenstrmungen, University Karlsruhe, PhD thesis
4. Halama J, Benkhaldoun F, Fort J (2011) Flux schemes based finite volume method for internal transonic
flow with condensation. Int J Numer Methods Fluids 65(8):953968
5. John V, Angelov I, Oncul A, Thevenin D (2007) Techniques for the reconstruction of a distribution
from a finite number of its moments. J Chem Eng Sci 62(11):28902904
6. Dykas S, Goodheart K, Schnerr GH (2003) Numerical study of accurate and efficient modeling for
simulation of condensing flow in transonic steam turbines. In: 5th European conference on Turboma-
chinery, Prague, pp 751760

123
S318 J. Halama, J. Fort

7. ejna M, Lain J (1994) Numerical modelling of wet steam flow with homogenous condensation on
unstructured triangular meshes. J ZAMM 74(5):375378
8. Gerber AG, Kermani MJ (2004) A pressure based Eulerian-Eulerian multi-phase model for non-
equilibrium condensation in transonic steam flow. Int J Heat Mass Transf 47:22172231
9. Sun L, Zheng Q, Liu S (2007) 2D-simulation of wet steam flow in a steam turbine with spontaneous
condensation. J Mar Sci Appl 6(2):5963
10. Hill PG (1966) Condensation of water vapor during supersonic expansion in nozzles, part 3. J Fluid
Mech 3:593620
11. Becker R, Doering W (1935) Kinetische Behandlung der Keimbildung in bersttingten Dmpfen.
J Ann Phys 24(8):719
12. Strang G (1968) On the construction and comparison of difference schemes. SIAM J Numer Anal
5:506517
13. Barschdorff D (1971) Verlauf der Zustandgroesen und gasdynamische Zuammenhaenge der spontanen
Kondensation reinen Wasserdampfes in Lavalduesen, Forsch. Ing.-Wes., vol 37, no 5
14. Dykas S (2001) Numerical calculation of the steam condensing flow. Task Q 4:519535

123
Copyright of Computing is the property of Springer Science & Business Media B.V. and its content may not be
copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written
permission. However, users may print, download, or email articles for individual use.

S-ar putea să vă placă și