• Keine Ergebnisse gefunden

The computational algorithm for the nonisothermal two-phase flow with interface capturing is validated using the case of a liquid jet impinging on a heated surface, by comparing the computed results with the empirical results of Liu et al. [55], the numerical results of Fujimoto et al. [27] and Tong [112], and the experimental results of Stevens [108] reprinted in the latter two references. The coupling at the solid-fluid interface in the case of conjugate heat transfer is validated using a simple case of one-dimensional transient heat transfer.

Liquid Jet Impinging on a Heated Surface

The sketch of the flow configuration is shown in Fig. 4.1. The liquid jet enters the computational domain through the nozzle and impinges onto the heated plate.

After the impingement, the thin liquid sheet is formed flowing over the wall surface and removing heat from it. When the steady state is reached, distributions of the temperature and the heat flux at the wall are established, being characterized by the Nusselt number at the wall. The computational mesh has the shape of a two-dimensional axisymmetric slice, adaptively refined in the regions comprising the falling jet and the liquid film spreading on the solid surface. The properties of liquid and gas are constant and correspond to those of water and air at ambient conditions.

The geometrical parameters are the jet diameter D0 = 4.06 mm, the plate diameter equal to 12D0 and the height representing the distance from the nozzle outlet to the plate of 3.7D0, with the characteristic Reynolds number Re = 10600 based on mean liquid velocity at the nozzle exit. At the wall surface the constant heat flux ˙qw = 1.49·105 W/m2 is applied. As in the aforementioned studies, the flow is assumed to be laminar. Although turbulent flow should be expected at the nozzle exit, its effects were neglected since the radial velocity of the thin liquid sheet spreading over the heated surface is relatively small and therefore it is assumed that the flow relaminarizes and stays laminar after the impingement.

For the constant applied heat flux at the wall surface, the conservation of energy requires the fluxes due to heat conduction and convection to be equal (Incropera and de Witt [42])

˙

qw =h(Tw−Tref) =−kw(∇T)w|, (4.12) where the subscript ref denotes the reference temperature of the undisturbed flow (in the particular case of jet impingement it is the liquid temperature at the nozzle outlet) and the symbol indicates the gradient component normal to the wall.

open boundary, adiabatic

outlet boundary, adiabatic

axis of symmetry

inlet (nozzle)

D0

3.7D0

wall, constant heat flux free surface

jet

liquid film Umean

12D0

y

r

Figure 4.1: The configuration for the case of liquid jet impinging on a heated surface.

The discretization of the convective term in Eq. (4.4) requires the value for the temperature at all boundary cell-faces to be supplied. This is done by first calculating the surface-normal temperature gradient from the prescribed wall heat flux

|(∇T)w| = q˙w

kw, (4.13)

and the temperature at the wall boundary is then obtained by interpolating the cell-center value using the calculated temperature gradient

Tb =TP +d·(∇T)w =TP +|d||(∇T)w|. (4.14) Taking the nozzle diameter as the characteristic length, the Nusselt number char-acterizing the heat transfer from the wall surface is by definition (Incropera and de Witt [42])

Nu = hD0

k , (4.15)

and substituting the overall heat transfer coefficienthfrom Eq. (4.12), the distribu-tion of the Nusselt number at the wall is obtained from the expression

Nuw = D0

kw

˙ qw

Tw−Tref. (4.16)

The simulation was performed using two different velocity profiles at the nozzle outlet, namely the uniform velocity with Uy = −Umean and the 1/7 power-law profile given by the expression Uy =−Umax(1−r/R0)1/7 with Umax =Umean/0.817 and R0 =D0/2. The computationally obtained distributions of the Nusselt number for the two cases are shown in Fig. 4.2 versus the normalized radial coordinate (r = r/D0) and compared to the aforementioned experimental, theoretical and numerical results. As expected and in accordance with the previous findings the best agreement with the experimental results was obtained in the case of the uniform velocity profile. Additionally, in both cases small waves are resolved at the free surface of the spreading liquid film, being more pronounced in the case of the applied power-law velocity profile, which is why the Nusselt number distribution shows small oscillations along the radial coordinate.

0 1 2 3 4 5 6

r 0

50 100 150 200 250 300

Nu

experiment, Stevens [108]

theory, Liu et al. [55]

numerical, Tong [112]

uniform Fujimoto et al. [27]

present simulation

0 1 2 3 4 5 6

r 0

50 100 150 200 250 300

Nu

experiment, Stevens [108]

theory, Liu et al. [55]

numerical, Tong [112]

numerical, Fujimoto et al. [27]

present simulation

Figure 4.2: The distribution of the Nusselt number at the wall for the uniform (left) and the power-law (right) inlet velocity profile at Re = 10600.

Transient Conjugate Heat Transfer

In order to verify the coupling at the solid-fluid interface, the transient heat transfer in the fluid with the accompanying heat conduction in the wall is computed. The computational domain is one-dimensional consisting of three layers placed above each other, namely the solid wall, the liquid layer placed on the solid surface and the gas layer above the liquid surface. As depicted in Fig. 4.3 all three layers have equal heights of 5 mm, the fluids are at rest, the properties of the wall are that of stainless steel, and those of the fluids correspond to water and air. The domain is divided into 100 and 50 cells in the fluid and the solid region, respectively, and in order to obtain better resolution, both the fluid and the solid mesh are graded, with the smallest cells near the solid-fluid interface. The bottom surface of the wall and the top surface of the air are kept at constant temperatures of T1 = 100C and T2 = 25C, respectively, and heat is conducted from the wall through the liquid to the air across the free surface.

The simulation was performed using constant liquid properties evaluated at 25C, as well as using variable properties according to Eqs. (4.7–4.8). The theoretical

heat flux at the steady state and constant thermophysical properties of the materials is determined from the expression

˙

q= T2−T1

3 i=1

∆yi ki

, (4.17)

which for the given geometry equals ˙q= 369.1 W/m2. This exact value is compared to the numerically obtained heat flux.

air, ka, cp,a

water, kw, cp,w

steel, ks, cp,s

T1

T2

y3y2y1

y

q

Figure 4.3: Initial configuration for the case of one-dimensional transient conjugate heat transfer.

The computed temporal distributions and the steady-state solutions of the tem-perature and the heat flux, the latter normalized by the theoretical value, are shown in Fig. 4.4.

-5 0 5 10

y, mm 20

40 60 80 100 120

T , °C

t = 1200 s t = 100 s t = 50 s

variable properties constant properties

-5 0 5 10

y, mm 0

5 10 15 20 25 30

q ·

constant properties variable properties t = 50 s

t = 100 s

t = 1200 s

Figure 4.4: Temporal temperature profiles and heat flux distributions for the case of one-dimensional transient conjugate heat transfer.

The results show smooth obtained profiles of both temperature and heat flux at the fluid-solid and liquid-gas interfaces without discontinuities. The steady-state

temperature profile is linear in all three layers in the absence of heat sources. The relative error in the numerically obtained heat flux in the case of constant liquid properties is less than 0.1%. As expected, in the case of variable properties, the steady-state heat flux is slightly higher due to the increase of the liquid thermal conductivity with temperature.