Вы находитесь на странице: 1из 11

03663199(94)001154

hr. J. Hydrogen Energy, Vol. 20, No. I I, pp. 881-891, 1995 Copyright @ International Association for Hydrogen Energy Else&r Science Ltd Printed in Great Britain. All rights reserved 0361X3199/95 $9.50 + 0.00

STUDY

OF TWO-DIMENSIONAL HEAT AND MASS TRANSFER DESORPTION IN A METAL-HYDROGEN REACTOR


A. JEMNI and S. BEN NASRALLAH Ecole Nationale dIngtnieurs de Monastir, Route de Kairouan, 5000 Monastir, Tunisia

DURING

(Received for publication

16 November 1994)

Abstract-A study of two-dimensional dynamic heat and mass transfer in a metal-hydrogen


is presented in this paper. A mathematical

reactor during desorption model has been established snd solved numerically by the method of

finite domains. The numerical simulation is used to present the time-space evolution of the temperature, the pressure and the hydride density in the reactor and to determine the sensitivity to some parameters (reactor geometry, outlet pressure,temperature of heating fluid and heat conductivity). This simulation allows us to study the effect of neglecting the term related to heat transport by convection in the model.

NOMENCLATURE

P
V

CP 4 Ed
HO

H f&s H/M k m M P P, % R

Re S
T t V

Specific heat (J kg- K-l) Particle diameter (m) Activation energy (J mol- ) Conductance between hydride bed and heating fluid (W mm2 K-l) Conductance between outlet face of the reactor and exterior medium (W m- K-l) Reactor height (m) Heat coefficient exchange between solid and gas (Wm- K-l) Hydrogen-to-metal atomic ratio Permeability (m) Hydrogen mass desorbed (kg mm3 s-l) Molecular weight (kg mol.- ) Pressure (Pa) Prandtl number Universal gas constant (J mol- Km ) Reactor ray (m) Reynolds number Solid-gas exchange area (m mm3) Temperature (K) Time (s) Gas velocity (m s-)

Density (kg mm3) Kinetic viscosity (m2 s- ) Aver.aging volume (m3)

Subscripts effective e

eq f g ge m, n se
S

ss

equilibrium heating fluid gas gas effective spatial index solid effective solid saturated

Superscripts gaseous phase g i time index


S

solid phase average volume fluctuation INTRODUCTION

Greek letters

AH Ar At
AZ

& A P

Reaction heat of formation (J kg- 1 Thickness of control volume (m) Time increment (s) Thickness of control volume (m) Porosity Thermal conductivity (W m-l K- -11 Dynamic viscosity (kg m- 1 s- )
881

The metal-hydrogen reactor can be used in many installations (hydrogen accumulator, heat accumulator, compressor, heat pumps and refrigerators, heat engines, etc). A knowledge of heat and mass transfer in a metalhydrogen reactor during the absorption and desorption of hydrogen is very important for their reactor optimization. Several attempts have been made to analyze the hydride behavior during the sorption phenomena [l-13]. Among these studies there are a some studies which have investigated the desorption phenomena [ll-131. The

882

HEAT AND MASS TRANSFER

where 4i is a quantity associated with the i phase. We also define the intrinsic average over a phase i as:

Fig. 1. Metal-hydrogen

one-dimensional model used by Mayer et al. [l l] and the two-dimensional model presented by Sun and Deng [12] and Pons [13] assume that gas and solid temperatures are equal and that gas pressure is constant in the reactor. In this paper we report on a theoretical study of the two-dimensional unsteady heat and mass transfer during the hydrogen desorption within a metal-hydrogen reactor. The model used in the present paper takes into account the effect of the difference between the solid and gas temperatures, as well as that of the variation of the gas pressure. We first present the set of equation which govern heat and mass transfer in the reactor during the desorption. The system of equations was resolved numerically by the finite domain method. The numerical simulation gives the time-space evolution of different state variables (temperature, pressure, hydride density) in the reactor and determines heat and mass transfer sensitivity to some of these parameters (outlet pressure, temperature of heating fluid, effective thermal conductivity of the solid and reactor geometry).

where wi is the volume occupied by phase i in the total averaging volume w. are The macroscopic differential equations obtained by taking the average of microscopic equations over the averaging volume w and using closing assumptions. The microscopic equations are the mass, the energy and the momentum equations balance in each phase and at the interface. These equations are obtained by using thermodynamic and mechanic laws of continuous media. Several simplifying assumptions are made in order to obtain a closed set of governing equations at the macroscopic scale: (1) the viscous dissipation and compression work and negligible; (2) the gas phase is ideal from the thermodynamic view point; (3) the dispersion term and the tortuousity term can be modeled as diffusive fluxes. The equations governing heat and mass transfer in a metal-hydrogen reactor are [lo]: Energy equations When the transfers are two-dimensional and depend on the r-and z-axes, the energy equations become: for the fluid

FORMULATION OF HEAT AND MASS TRANSFER The cylindrical reactor considered in this paper exchanges heat through lateral and base areas at a constant temperature and a constant flow rate heating fluid (Fig.1). The reactor is composed of a solid phase (metal-hydride) and a gaseous phase (hydrogen), so it is a discontinuous medium. The equations which govern heat and mass transfer in the reactor media are obtained by changing the scale of description in space [14]. We pass from a microscopic view, in which the averaging volume w is small compared with the pores, to a macroscopic view, in which the averaging volume is large with respect to the pores. This scale changing permits conversion from the real discontinuous medium to a fictious continuous equivalent medium. Each macroscopic term is obtained by averaging the microscopic one. We define the average of some microscopic function 4 as :

- H,,( T; - T:)S + mc,,( T; - Tz) (I)


for the solid

+ H8J T; - T:) x S - m(AH + CpsT; - C,,T;). (2)

A. JEMNI

and S. BEN NASRALLAH

883

Momentum equation The gas velocity can be expressed by Darcys law:

for the fluid Pi w a(T:) ~ (z, R) = h(T; - T,); ar (H, r) = h(Ti - T,),

Hydrogen mass balance For the gas: the mass conservation hydrogen is: F:at equation of the for the solid

a aZ -%e

XJ:) +

div(pi V,) := -m.

-)

(4)
-),

~ se ar
a(T) se x

a(T)

(z, R) = h(Tz - T,);

Assuming that the hydrogen is an ideal gas and considering Darcys law, the mass conservation equation of the hydrogen becomes 1

W,

h(TS

T),

(10)

k 1 a vg r 3 For the solid: the mass conservation equation of the solid becomes:

(6)
Initial conditions Initially, the temperature, the pressure and the hydride density in the reactor are assumed to be constants:

where h is the conductance between hydride bed and heating fluid and Tf is the temperature of the heating fluid. The heat exchange at the outlet reactor and the flow existing above a porous surface are more complex because, on the one hand, the outlet face of the reactor in contact with the gas can release a natural convection movement which is analogous to the one observed on a horizontal plate-the metallic edge at the outlet of the reactor influences this movement; on the other hand, the convective fl,~x of the gas at the reactor outlet disturbs the effect of l.he natural convection movement. In order to resolve this problem, we introduce a heat coefficient exchange h, between the outlet face of the reactor and the extl:rior medium; so at the outlet, the gas temperature and the solid temperature are related by the equations: i,, 2 (0, r) = h,(Ti - 7;.);

T; = T; = To, Pi = PO, p* = po.


Boundary conditions Taking into account the symmetry about the z-axis, we deduce that T (2, 0) = 0; 2 (2, 0) = 0; F (2, 0) = 0. (7)

4,
Reaction kinetic

(0, r) = h,(Tz - Tf),

The expression of the hydrogen mass desorbed per unit time and unit volume, m, given for LaNi,-hydrogen desorption in Ref. [ 111, is: m = C, exp

The wall is impervious and therefore:

(z, fo = 0; 2

(H, r) = 0.

(8)

The outlet reactor pressure is assumed to be constant Qz = 0, r) = PO. The heat flux continuity through the lateral area and base area (r = R and z = H) allow us to write the following equations:

where E, is the activation energy for the diffusion of the hydrogen atoms through the hydride phase and C, is a constant. Based on the work of Suda and Kobayas [l], E, and C, were equal to 16,473 J mol- and 9.57 s-l respectively. The expression for the hydrogen equilibrium pressure P,, was deduced by smoothing of the experimental results obtained by Uchida et al. [15]:

884

HEAT AND MASS TRANSFER

N-l

(2) the accumulated terms and the source terms can be approximated by their averages on the control domain contructed around P,,,. In order to ensure the stability of the numerical model, we used an implicit scheme and we supposed that the values ofconvected quantities along the face of the control domain are equal to their values at the grid point situated in the upstream (upwind scheme). The first derivatives, which are evaluated on the control domain faces, are approximated by: w = ccl,, az 0 n+(l/ZLm 1 2 m-l m m+l Fig. 2. Numerical grid M-l M
i+l

n+l n n-l

2
1

- 43 AZ

Considering these assumptions, the form of the resulting algebraic equations becomes:

+ A,&+,:,,

+ A,&+:,,

+ A,.

(14)

-1193.2852@+

479.9432(;~}

The resulting system of algebraic equations is solved numerically by the iterative line-by-line method scanning. The choice of this method is justified by its rapidity of convergence compared with the point-by-point method. Scanning along the z-axis, we set: (13) where Z = A,~~~~,, 1 + A,$:,_ 1 + A,. This resolution method consists in the evaluation, using the predicted solution, of A,, A,, A, and Z coefficents. Then, the tridiagonal resulting system of equations is solved by the standard Gaussian elimination method. If the difference between the calculated and estimated solutions is small, the convergence to the solution is achieved, or else we repeat the procedure of evaluation of the coefficients using the solutions which have already been calculated until convergence. RESULTS AND DISCUSSION

x ,x,(-3323.884(+-&))

x 105.

NUMERICAL

RESOLUTION

The system of equations presented is solved numerically by the method of finite differences based on the notion of control domain as described by Patankar [ 161. The advantage of this method is to ensure the flux conservation, and thus to avoid the generation of parasitical sources. The method consists of defining a grid of points P,,, within the calculated domain and then builds around each point a control domain (n, m). In our case, we have considered a regular mesh within the integrated domain and we have used 35 x 25 nodes. Figure 2 shows the mesh used in the numerical resolution. The point P,,, is located in the center of the control domain. The value of the physical scalar 4 at P,,,, and at time t + At will be denoted &,,. The equations are integrated on this control domaib and on the interval of time [t, t + At]. At the boundary limits of the reactor, the equations are made discrete by integrating over half of the control domain by taking into account the boundary conditions. At the corner we have used the quarter of control domain. In order to bring the resulting integral equations back to algebraic equations tying together the solution values to the nodes on the grid, we make the following hypotheses: (1) the fluxes are constant on the face of the control doman that is perpendicular to them;

During the theoretical simulation, the volume of the considered reactor is 235.6 cm3, and it is heated by a heating fluid at constant temperature Tf = 313 K. The outlet hydrogen pressure is P, = 1.5 x 10 Pa. The reactor is filled with LaNi,. Initially the solid temperature is 290 K. The permeability, the porosity, the effective thermal conductivity of solid and the effective thermal conductivity of the gas are, respectively, lo- rn, 0.5, 1.2 W m-l K- and 0.12 W mm Km. The heat exchange coefficient between solid and hydrogen is given by [ 171:

H,, = 2 (2 + l.lP,3 Re0.6),


P

(16)

A. JEMNl and S. BEN NASRAL.LAH


t = 300 t= s

885

60s

Fig. 3. Time-space evolution of temperature (R = 0.05 m, H = 0.03 m, T = 313 K, T, = 290 K, P, = 1.5 x lo5 Pa, K = lo-* m2, E = 0.5, i,se= I.2 W m- Km. 4, = 0.12 W mm K-l).

HEAT AND MASS TRANSFER

5400

t = 1500 5:

3 8, i -1 t = 9000
Y

Fig. 4. Time-space evolution

of solid and gas temperature dilTerence(R = 0.05 m, H = 0.03m, T = 3 13 K, T, = 290 K, P, = 1.5 x lo5 Pa, K = 10-s m*, E = 0.5, I.,, = 1.2 W m- K-l, I.,, = 0.12 W m- K-l).

A. JEMNI

and S. BEN NASRALLAH

887

t =

IOROO s

Fig, 5. Time-space

evolution

of hydride metal density CR = 0.05 my H = 0.03 m, Tf = 313 K, To = 290 K, P, = 1.5 X 10 Pa, 1.2~m~ K-,i.,, =0.12 W mm K-). K = 10-s mZ,e=o.5,;.,,=

HEAT AND MASS TRANSFER

O.10 t =
9000 s
t = 10800 s

Fig.

6. Time-space evolution of pressure (P - PO) (R = 0.05 m, H = 0.03m, T, = 313 K, To = 290 K, f, = 1.5 x 10 Pa, K = 1~
m2, 6 = 0.5, A,, = 1.2 W m-l K-l, I,, = 0.12 W m- K-l).

A. JEMNI and S. BEN NASRALLAH where d, is the particle diameter, P, is the Prandtl number and Re is the Reynolds number. The results of numerical simulation are presented as a three-dimension curve giving the time-space evolution during the desorption of the temperature, the hydride bed density and the pressure. Taking into account the symmetry about the z-axis, the figures are plotted for only half of the cylinder Figure 3 shows the temperature distribution of fluid within the reactor after 60, 300, 1800, 5400, 9000 and 10,800 s. Keeping in mind that the dehydriding reaction was endothermic, the temperature inside the reactor first decreases, then increases. This is because of the decrease in reaction velocity. Against the wall, the temperature is greater than in the interior of the reactor, due to the external heating fluid. After a substantial period of time, the remaining quantity of hydrogen in the reactor becomes too small. Consequently, the heat needed for the hydrogen dissocation tends to zero, so the problem comes down to one of stationary heat conduction inside an inert porous medium. The time-space evolution of the solid temperature within the reactor is similar to the time-space evolution of gas temperature; however, the difference between the solid and the gas temperature is important for a short time, then decreases with time (Fig. 4). This difference is more important near the impervious wall, because the thermal characteristics of the two phases are very different. The results show that there is no thermal equilibrium in the whole reactor. Fig 5 shows that the mass desorbed is higher near the wall where the temperature is high. This is because the dissociation reaction velocity increases with temperature. When time is long enough, the hydride density inside a reactor tends to a constant.
15

889

Tf=313K Tf=343K Tf=373K

Fig. 8. Influence of the heating fluid temperature on the total mass desorbed.

Figure 6 shows the over-pressure inside the reactor which is due to the hydrogen desorption. This overpressure is more noticeable near the heated partition since the desorption reaction velocity is high. When time is long, pressure increases within the reactor and tends to the outlet pressure. The results of the numerical simulation, for the selected dimensions (R = 5 cm and H = 3 cm), show that the heat and the mass transfers depend on r and z, so neglecting the two-dim.ensional effect can generate an important error Sensitivity to the outlet pressure From Fig. 7 the effect of the outlet hydrogen pressure on the hydrogen mass desorbed can be seen. We notice that the desorption reaction is faster when the outlet pressure is small. Sensitivity to the temperature of the fluid heating The evolution of the total reactor hydrogen mass desorbed with time for different temperatures of the fluid heating Tr ~Fig. 8) shows that the reaction velocity increases with Tr.

---

PO= 0.25 x 10e5 Pa PO = 0.50 x 10.Pa PO= 1.50 x 10m5 Pa

Sensitivity to the effective thermal conductivity of the solid The heat needed for hydrogen desorption is transferred from the heal.ed fluid to the inside of the reactor essentially by a conducion process. Thus it is of interest to explore the effect of thermal conductivity. Figure 9 shows that the time required to release hydrogen stored is small for high values of the effective thermal conductivity. According to this, increasing the effective conductivity of the solid permits an acceleration of the desorption reaction

Fig. 7. Influence of the hydrogen outlet pressure on the total mass desorbed.

890

HEAT AND MASS TRANSFER

15r

t=3600s

hse

= 1.2 W / (Kin) Ase=SW/(Km) hse=lOW/(Km)


0 I 5 I 10 I 15 I 20

Time (s) Fig. 9. Influence of the effective thermal conductivity of the solid on the total mass desorbed.

h/r Fig. 10. Influence of the height to the radius ratio of the reactor on the total hydrogen mass desorbed.

velocity. Several methods have been suggested for increasing this conductivity; for example, addition of wire or metal powder of high thermal diffusivity. Choi and Mills [S] have studied this effect on hydrogen absorption. They showed that increasing the thermal conductivity up to a value of about 4 W K- m-i gives a substantial improvement in the rate of hydrogen absorption, whereas an increase above a value of about 5 W K- m-l yields little further improvement. In the case of desorption, we note that this result remains valid.

Efict

of the term of heat transport by convection

In order to simplify the theoretical model, the numerical simulation during the desorption was realized with and without taking into account the term of heat transport by convection (Pi I/,C,, T:). The results for different temperatures of heating fluid, pressures and H/R values show that the change of solid and gas temperature is less than 1 %, so the convection heat transport term can be neglected.

Sensitivity to the reactor geometry The total hydrogen mass desorbed is plotted for each reactor dimension as a function of time in Fig. 10. The reactor volume, the physical characteristcs and the boundary conditions were kept constant during the simulation. Examining this plot, we note that the total hydrogen mass desorbed is a minimum at a ratio of the reactor height(H) to the reactor radius (R) equal to unity, so for low values of H/R, the heat and mass transfer in the reactor are one-dimensional and depend only on z. Under these conditions, the resistance to the transfers along the z-direction increases with H/R values, hence the total hydrogen mass desorbed decreases. For large values of H/R, the transfers are also one-dimensional and depend only on r. When H/R rises the resistance to the transfers according to r decreases and the total mass desorbed increases. For intermediate values of H/R, the two-dimensional effects are not negligeable. When H/R increases the resistance to the transfers along the zdirection increases and that along the radial direction decreases. These competing effects explain the existence of the minimum. This result also applies in the absorption case [lo]. CONCLUSIONS A mathematical two-dimensional model describing the stationary heat and mass transfer processes within a metal-hydrogen reactor has been developed and solved for hydrogen desorption. The model takes into account the effect of the difference between the solid and gas temperatures, as well as that of the variation of gas pressure. The calculated time-space evolution of pressure, temperature and mass shows the significant influence of two-dimensional effects for the reactor dimensions used. According to the numerical simulation results the choice of reactor geometry, the outlet pressure and the temperature of the heating fluid are very important. These results also show that the thermal equilibrium assumption is not valid in the whole reactor, that the term of heat transport by convection can be neglected and that an increase in the effective thermal conductivity of the solid leads to significant improvement of performance for the reactor. Therefore, the model can be used as a helpful tool in the optimization of the metal-hydrogen reactor designs and performance.

A. JEMNI

and S. BEN NASRA LLAH

891

REFERENCES
1. S. Suda and N. Kobayashi, J. Less-Common Metals 73, 119 (1980). 2. P. D. Goodell, G. D. Sandrock and E. L. Huston, J. km-Common Metals 73, 135 (1980). 3. W. Supper, M. Groll and U. Mayer, J. Less-Common Met& 104, 279 (1984). 4. M. Y. Song and J. Y. Lee, Int. J. Hydrogen Energy 8, 363 (1983). 5. M. Kawamura, S. Ono and Y. Mizuno, J. Less-Common Metals 89, 365 (1983). 6. A. Jemni, S. Ben Nasrallah, J. Lamloumi and A. Percheron Guegan, In!. Symp. on Metal-Hydrogen Systems, Uppsala, Sweden, 8812 June (1992). 7. M. Ram Gopal and S. Srinivasa Murthy, Int. J. Hydrogen Energy 17, 795 (1992). 8. H. Choi and A. F. Mills. Int. J. Heat Mass Transfer 33. 1281 ( 1990)

9. S. Wakao, M. Sekine. H. Endo, T. Ito and H. Kanazawa, J. Less-Common Metals 89, 341 (1983). 10. A. Jemni and S. 9. Nasrallah, Int. J. Hydrogen Energy -. 20, 43-52 (1995). 11. U. Mayer, M. Groll and W. Supper, J. Less-Common Met& 131, 235 (1987). 12. D. W. Sun and S. J. Deng, J. Less-Common Metals 155, 271 (1989). 13. M. Pons, Transferts de chaleur dans la poudre de LaNi, et ieur ccuplage avec la reaction, dhydruration, Thesis, Univerrite Paris (1991). 14. S. Whitaker, Advances in Heat Transfer, pp. 119-203. Academic Press, New York (1977). 15. H. Uchida, K. Temao and Y. C. Huang, Int. Symp. on Metul-jYydrogen Systems, Stuttgart, Germany, 449 September (1988). 16. S. V. Patankar, Numericul Heat Trunsier Fhid Flow. Hemisphere/MacGraw-Hill, New York (1980). 17. C. Dang Vu and B. Delcambre, Rec. Phys. Appl. 22, 487 (1987).

Вам также может понравиться