Influence of the Flow Direction on the Mass Transport of Vapors Through Membranes Consisting of Several Layers

Similar documents
Accepted Manuscript. Title: Linearized description of the non-isothermal flow of a saturated vapor through a micro-porous membrane

3.10. Capillary Condensation and Adsorption Hysteresis

2. Modeling of shrinkage during first drying period

I. Borsi. EMS SCHOOL ON INDUSTRIAL MATHEMATICS Bedlewo, October 11 18, 2010

DEVELOPING A MICRO-SCALE MODEL OF SOIL FREEZING

Liquids and solids are essentially incompressible substances and the variation of their density with pressure is usually negligible.

PHYSICS OF FLUID SPREADING ON ROUGH SURFACES

Diffusion and Adsorption in porous media. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

Boundary Conditions in Fluid Mechanics

Capillary surfaces and complex analysis: new opportunities to study menisci singularities. Mars Alimov, Kazan Federal University, Russia

Influence of water on the total heat transfer in evacuated insulations

Reaction at the Interfaces

MOLECULAR SIMULATION OF THE MICROREGION

Clausius Clapeyron Equation

A drop forms when liquid is forced out of a small tube. The shape of the drop is determined by a balance of pressure, gravity, and surface tension

Four Verification Cases for PORODRY

Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces.

Colloidal Particles at Liquid Interfaces: An Introduction

5.2 Surface Tension Capillary Pressure: The Young-Laplace Equation. Figure 5.1 Origin of surface tension at liquid-vapor interface.

CHAPTER 2. SOIL-WATER POTENTIAL: CONCEPTS AND MEASUREMENT

Chapter 10: Boiling and Condensation 1. Based on lecture by Yoav Peles, Mech. Aero. Nuc. Eng., RPI.

Evaporation rates from square capillaries limited by corner flow

Chemical Reaction Engineering Prof. Jayant Modak Department of Chemical Engineering Indian Institute of Science, Bangalore

Dimensionless Numbers

Hartmann Flow in a Rotating System in the Presence of Inclined Magnetic Field with Hall Effects

PERFORMANCE CHARACTERISTICS OF A LONG HEAT PIPE

Membrane processes selective hydromechanical diffusion-based porous nonporous

8.2 Surface phenomenon of liquid. Out-class reading: Levine p Curved interfaces

Lecturer, Department t of Mechanical Engineering, SVMIT, Bharuch

Summary of Dimensionless Numbers of Fluid Mechanics and Heat Transfer

Numerical Studies of Droplet Deformation and Break-up

MOLECULAR DYNAMICS SIMULATION ON TOTAL THERMAL RESISTANCE OF NANO TRIANGULAR PIPE

Accurate Determination of Pore Size Distributions

Multiphase Flow and Heat Transfer

This is an author-deposited version published in : Eprints ID : 16043

dynamics of f luids in porous media

Shape of the Interfaces

A FLAMELET MODEL FOR DIFFUSION FLAMES IN POROUS MEDIA. Max Akira Endo Kokubun. Fernando Filho Fachini

Peter Buhler. NanothermodynamicS

Complexity of Two-Phase Flow in Porous Media

Fluid Mechanics Theory I

THEORETICAL AND EXPERIMENTAL STUDY OF NH 3 ABSORPTION INTO AMMONIA/WATER SOLUTION USING POLYMERIC POROUS MEMBRANES.

Supporting Information: On Localized Vapor Pressure Gradients Governing Condensation and Frost Phenomena

Absorption of gas by a falling liquid film

Outline. Definition and mechanism Theory of diffusion Molecular diffusion in gases Molecular diffusion in liquid Mass transfer

Supplementary Figures

Lecture 9 Laminar Diffusion Flame Configurations

APPLICATION OF DIFFERENTIAL SCANNING CALORIMETRY TO CORE ANALYSIS

A Mathematical Model for Analyzing the Thermal Characteristics of a Flat Micro Heat Pipe with a Grooved Wick

Biological Process Engineering An Analogical Approach to Fluid Flow, Heat Transfer, and Mass Transfer Applied to Biological Systems

CONTENTS. Introduction LHP Library Examples Future Improvements CARMEN GREGORI DE LA MALLA EAI. ESTEC, October 2004

P = 1 3 (σ xx + σ yy + σ zz ) = F A. It is created by the bombardment of the surface by molecules of fluid.

COMPARISON OF WETTABILITY AND CAPILLARY EFFECT EVALUATED BY DIFFERENT CHARACTERIZING METHODS

1 One-dimensional analysis

Thickness and Shape of Films Driven by a Marangoni Flow

Pore scale mechanisms for enhanced vapor transport through partially saturated porous media

Thermodynamics of Fluid Phase Equilibria Dr. Jayant K. Singh Department of Chemical Engineering Indian Institute of Technology, Kanpur

On the displacement of two immiscible Stokes fluids in a 3D Hele-Shaw cell

THE PHYSICS OF FOAM. Boulder School for Condensed Matter and Materials Physics. July 1-26, 2002: Physics of Soft Condensed Matter. 1.

Chapter Seven. For ideal gases, the ideal gas law provides a precise relationship between density and pressure:

PHASE CHANGES EVAPORATION EVAPORATION PHYSICAL PROPERTIES OF LIQUID PHYSICAL PROPERTIES OF LIQUID SOME PROPERTIES OF LIQUIDS 2014/08/08

In-Plane Liquid Distribution In Nonwoven Fabrics: Part 2 Simulation

Part I.

Aalborg Universitet. Transport phenomena in gas-selective silica membranes Boffa, Vittorio. Creative Commons License Unspecified

Investigations Of Heat Transfer And Components Efficiencies In Two-Phase Isobutane Injector

CFD SIMULATIONS OF FLOW, HEAT AND MASS TRANSFER IN THIN-FILM EVAPORATOR

8.2 Surface phenomena of liquid. Out-class reading: Levine p Curved interfaces

Project: Drying of a sausage. Report

Ultrafast water harvesting and transport in hierarchical microchannels

FLAT PLATE HEAT PIPES: FROM OBSERVATIONS TO THE MODELING OF THE CAPILLARY STRUCTURE

LIQUID FILM THICKNESS OF OSCILLATING FLOW IN A MICRO TUBE

Chapter 3 PROPERTIES OF PURE SUBSTANCES. Thermodynamics: An Engineering Approach, 6 th Edition Yunus A. Cengel, Michael A. Boles McGraw-Hill, 2008

CHARCOAL SORPTION STUDIES I. THE PORE DISTRIBUTION IN ACTIVATED CHARCOALS1

Chapter 2 Mass Transfer Coefficient

APPLICATION OF A MODEL FOR SOLUTION OF SHOCK WAVE PARAMETERS IN STEAM TO EVALUATION OF VALUE OF SPEED OF SOUND

Mass Transfer Operations

Pore-Level Bénard Marangoni Convection in Microgravity

Liquid-vapour oscillations of water in hydrophobic nanopores

Effect of pressure on the rate of evaporation from capillaries: statistical rate theory approach

Effect of Sorption/Curved Interface Thermodynamics on Pressure transient

Binary Fluid Mixture and Thermocapillary Effects on the Wetting Characteristics of a Heated Curved Meniscus

Evaporation-driven soil salinization

Heat and Mass Transfer Unit-1 Conduction

The Effects of Friction Factors on Capillary Tube Length

Scale-up problems are often perceived as difficult. Here the reaction calorimetry has proven to be

Steady waves in compressible flow

IHTC DRAFT MEASUREMENT OF LIQUID FILM THICKNESS IN MICRO TUBE ANNULAR FLOW

ENGR Heat Transfer II

Unsaturated Flow (brief lecture)

MHD and Thermal Dispersion-Radiation Effects on Non-Newtonian Fluid Saturated Non-Darcy Mixed Convective Flow with Melting Effect

Diffusional Growth of Liquid Phase Hydrometeros.

Heat and Mass Transfer Prof. S.P. Sukhatme Department of Mechanical Engineering Indian Institute of Technology, Bombay

CHAPTER 3: SURFACE AND INTERFACIAL TENSION. To measure surface tension using both, the DuNouy ring and the capillary tube methods.

Mechanical Engineering. Postal Correspondence Course HEAT TRANSFER. GATE, IES & PSUs

COMBINED EFFECTS OF RADIATION AND JOULE HEATING WITH VISCOUS DISSIPATION ON MAGNETOHYDRODYNAMIC FREE CONVECTION FLOW AROUND A SPHERE

Transport processes. 7. Semester Chemical Engineering Civil Engineering

Convective Mass Transfer

Adsorption Processes. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

Modeling as a tool for understanding the MEA. Henrik Ekström Utö Summer School, June 22 nd 2010

A droplet of colloidal solution is left to evaporate on a superhydrophobic surface. Avijit Baidya

Transcription:

Influence of the Flow Direction on the Mass Transport of Vapors Through Membranes Consisting of Several Layers Thomas Loimer 1,a and Petr Uchytil 2,b 1 Institute of Fluid Mechanics and Heat Transfer, Vienna University of Technology, Resselgasse 3, 14 Wien, Austria 2 Institute of Chemical Process Fundamentals, Czech Academy of Sciences, Rozvojová 135, 165 2 Praha, Czech Republic a thomas.loimer@tuwien.ac.at, b Uchytil@icpf.cas.cz Keywords: Porous media, phase change, vapor flow, inorganic membranes Abstract. The flow of isobutane through a porous ceramic membrane consisting of three different layers is theoretically investigated. The individual layers have different thicknesses with pore sizes of 1 nm for the separation layer, 1 nm for the middle layer and 6 µm for the support layer. The mass flow is calculated for values of the upstream relative pressure, i.e., the ratio of the upstream pressure to the upstream saturation pressure, varying between.7 and 1. For small relative upstream pressures, the fluid does not condense within the porous membrane, and the mass flux is a few percent larger for the flow direction from the support to the separation layer than for the other flow direction. For larger relative upstream pressures, the fluid may condense and liquid or a two-phase mixture may be transported through parts of the membrane. The mass flow can become about 7 times larger for one flow direction than for the opposite flow direction. The flow is modeled accounting for capillary condensation, for the transport of energy, and for the temperature variation due to the Joule-Thomson effect. Introduction For the pressure-driven flow of a vapor through a porous medium, there are several reasons why the fluid may condense, flow as a liquid or a two-phase mixture through parts of the porous medium and eventually re-evaporate. Shall the situation be such that the fluid is in a state of a vapor near saturation at the upstream side of the membrane. The permeability of the membrane shall be small, hence the pressure difference is large. At the downstream side, because the pressure is smaller than at the upstream side, the vapor is in a state farther away from saturation than at the upstream side. Nevertheless, although the state of the fluid moves from a state close to saturation to a state farther away from saturation, condensation may occur. Condensation may occur due to the Joule-Thomson effect [1 4], due to capillary condensation [5, 6], or due to absorption [7]. If the flow is adiabatic, due to the Joule-Thomson effect the fluid is colder at the downstream side of the membrane than at the upstream side. The temperature gradient causes the transport of heat in downstream direction. However, if the vapor at the upstream side of the membrane is close enough to saturation, the vapor may have to condense partially or fully to release a sufficient amount of heat. Due to the Joule-Thomson effect, condensation may occur partially or fully, a two-phase mixture or a liquid may flow through a part or the entire porous membrane. Moreover, the wetting properties of the fluid-solid combination, i.e., the wetting angle, do not play a role. Accounting for the Joule-Thomson effect, but not for capillary condensation, the flow of a vapor through a single, homogeneous porous membrane was analyzed in Refs. [1] and [2]. Capillary condensation denotes the phenomenon that at a curved interface the gaseous phase of a fluid is in equilibrium with its liquid phase at a pressure different from the saturation pressure. If the interface is concave, as happens in pores where the solid material is wetted by the liquid phase, the equilibrium pressure is decreased with respect to the saturation pressure and a vapor may condense

in the pores of a solid material, whereas it stays in the gaseous state in free space. Note that, if only capillary condensation is considered, the fluid may either flow in its liquid or in its gaseous phase through parts of the porous membrane, but not as a two-phase mixture. Ceramic membranes usually consist of several layers which differ in pore size. The separation layers, which has the smallest pores, is made as thin as possible to reduce the flow resistance. For mechanical stability, the separation layer must be supported by thicker layers. These ceramic membranes are asymmetric with respect to the flow direction. For the flow of a gas, it has been shown that the mass flux through an asymmetric porous membrane can differ by a few percent if the fluid is not only transported by viscous, but also by free molecular flow [8 12]. For a specific choice of the membrane layers, differences of up to 6% were reported. In a related problem, the evaporation of a liquid in a slender, porous cylinder, it was shown that a non-homogeneous distribution of the permeability in flow direction may favourably affect the evaporation process [13]. Here, the flow of vapors of isobutane through an asymmetric porous membrane is described. The mass flow is given for conditions where the vapor stays in its gaseous phase and for conditions where condensation occurs. The flow is described accounting for the Joule-Thomson effect and for capillary condensation, but not for adsorption or surface flow. Transport occurs due to viscous and free molecular flow. The individual layers of the ceramic membrane are modeled as bundles of effective capillaries, i.e., the layers are described by their pore size, by the tortuosity and by the void fraction. As a consequence, the pressure difference at a front of phase change within a layer is everywhere the same for the same temperature. Also, for a given temperature, the curvature of a liquid-vapor interface is everywhere the same within one layer of the membrane. Theory The flow is described assuming one-dimensional, adiabatic flow through a porous structure consisting of several layers. The liquid phase of the fluid shall ideally wet the solid matrix of the membrane. The boundaries between two membrane layers are modeled as a jump in the pore diameter. Hence, the pore has a corner at the locations of boundaries. Since here we have a wetting system, a liquid meniscus stays attached to the corner. The radius of curvature of the meniscus may change between the value of the radius of the smaller pore and the value of the radius of the larger pore. Within a pore, the radius of curvature is fixed and equal to the pore radius. At curved menisci, the pressure difference is given by the Young-Laplace equation, and the pressure of the gaseous phase is given by Kelvin s equation. The thermic and caloric properties of the fluid are accurately modeled using engineering correlations, hence the temperature difference due to the Joule-Thomson effect is recovered. The flow is governed by the balances of mass, momentum and energy, ṁ = const, ṁ = κ dp ν dz, ṁh + q = const., where ṁ refers to the mass flux, κ is the permeability of the membrane, which is given in units of length squared, the apparent kinematic viscosity of the fluid is given by ν app and p refers to the pressure. The spatial coordinate is given by z, the specific enthalpy of the fluid is denoted by h and q is the heat flux. The apparent kinematic viscosity of the gaseous phase of the fluid is expressed by ν app = ν g (1 + βkn) 1, where ν g is the kinematic viscosity of the gaseous phase of the fluid, β is a factor taken here to be equal to 8.1 and Kn stands for the Knudsen-number. The heat flux is given by Fourier s law of heat conduction, q = k dt dz, (5) (1) (2) (3) (4)

Table 1: Properties of the three layers of the ceramic membrane. Layer Pore radius Thickness void fraction tortuosity 1 1 nm 2 µm.6 3 2 1 nm 15 µm.6 3 3 6 µm 2 mm.6 3 where k is the effective thermal conductivity of the fluid-filled membrane and T is the temperature. At fronts of phase change within the membrane, the pressure difference between the liquid and the gaseous phase of the fluid is given by the Young-Laplace equation, p = 2σ/r, (6) where the surface tension is given by σ and the pore radius is denoted by r. At the interface, the pressure of the gaseous phase of the fluid, p K, is given by Kelvin s equation, ( ) p K 2σ v liq = exp. (7) p sat r RT Here, p sat refers to the saturation pressure, v liq to the specific volume of the liquid phase of the fluid, and R to the specific gas constant. The governing equations are solved subject to the boundary conditions T = T 1, p = p 1, q = at z, (8) p = p 2 at z = z 2. (9) The membrane extends from z = to z = z 2. Correlations for the fluid properties are taken from literature, see Table A3 in Ref. [4]. Results The mass flow is calculated for the flow of isobutane through a ceramic membrane. The solid material has a thermal conductivity of 36 W/(mK). The properties of the three layers of the membrane are given in Table 1. The upstream temperature T 1 is 293.15 K, the pressure difference between the upstream and the downstream side of the membrane is fixed to.5 bar. Fig. 1a shows the mass flow through the membrane, made dimensionless with the mass flow that would result if the gaseous phase would be transported isothermally and only by Knudsen flow through the membrane, ṁ gas,knudsen. The isothermal transport due to Knudsen, or free molecular flow is independent of the orientation of the membrane. (The mass flow due to isothermal viscous flow alone is also independent of the orientation of the membrane. The mass flow due to the combination of free molecular and Knudsen flow depends on the membrane orientation.) In Fig. 1a, the mass flow of isobutane is plotted for the flow in direction from the separation to the support layer, ṁ A, and for the reverse direction, ṁ B. Figure 1b shows the ratio ṁ A /ṁ B. For small values of the relative pressure, i.e., for a vapor far from saturation, the mass flux is slightly larger in direction B, the flow direction from the support to the separation layer, than in direction A. However, at about p 1 /p sat.8, the mass flux increases strongly for flow direction A, because condensation occurs in the separation layer. For flow direction B, the separation layer is located at the downstream side, it experiences smaller pressures and capillary condensation occurs at larger values of p 1 /p sat. This is the case for p 1 /p sat.84. The mass flux ṁ A can become about 3 times larger than ṁ B. For p 1 = p sat ṁ B is about 7 times larger than ṁ A.

ṁ/ṁgas,knudsen 5 4 3 2 1 ṁ A (separation layer upstream) ṁ B (separation layer downstr.) ṁ A ṁ B 4 3 2 1 a).7.8 ➁.9 ➃ ➄ 1 p 1 /p sat b).7.8 ➁.9 ➃ ➄ 1 p 1 /p sat Figure 1: Mass flow (a) and mass flow ratio (b). The mass flow for the two flow directions, ṁ A and ṁ B, is made dimensionless with the isothermal purely free molecular flow of the gaseous phase, ṁ gas,knudsen. The ratio ṁ A /ṁ B is shown in (b). Numbers in circles mark values of p 1 /p sat referred to in Fig. 2. flow direction A p 1 /p sat.8.84.9.95.97 ➁ ➃ ➄ flow direction B 1. L 1 L 2 L 3 L 3 L 2 L 1 Phases: liquid, gaseous, two-phase. Figure 2: Phases within the membrane. Some values of p 1 /p sat are also denoted by numbers in circles, cf. Fig. 1. For p 1 = p sat, the mass flux of ṁ B 5ṁ gas,knudsen corresponds to ṁ B 4 g/(m 2 s). For this value, the Eckert-number in the free space downstream of the membrane is about 1.5 1 6 and the Mach-number, also in the free space, is about 2 1 4. The large changes in the mass flux are a consequence of condensation and the formation of liquid menisci in the pore space. The menisci have a large curvature, therefore there is a large pressure difference across the menisci which dominates the flow. The distribution of phases within the membrane is depicted in Fig. 2 for several values of p 1 /p sat. A comparison with Figs. 1a and 1b shows that large changes in the mass flux occur if condensation occurs in an additional layer. For flow direction A, for a vapor close to saturation also a liquid film may form in front of the membrane. The temperature and pressure distributions for p 1 /p sat =.95 for flow direction A are plotted in Figs. 3a and 3b. The temperature distribution, Fig. 3a, is plotted versus the spatial coordinate made dimensionless with the total thickness of the membrane, L 1 + L 2 + L 3. The extent of the second layer is indicated with dashed lines. The separation layer is too thin to be indicated by a line. The temperature distribution shows that there is a temperature boundary layer in front of the membrane. Nearly the entire decrease of the temperature to its downstream value takes place within this boundary layer, while the membrane everywhere assumes a temperature very close to its downstream value T 2. The large differences of the temperature variation are caused by the large differences of the thermal

(T T2)/(T1 T2) 1.8.6.4.2 a) p [bar] p 1 p 2 3 2 1 1 2 3.5.5 1.5 1.5 1 z/(l 1 + L 2 + L 3 ) z/l 1.5 1 z/l 3 b) z/l 2 Figure 3: Temperature (a) and pressure distribution for p 1 /p sat =.95, flow direction A. conductivities of the ceramic material and the isobutane vapor, viz., 36 W/(mK) versus.16 W/(mK). The pressure distribution is plotted in Fig. 3b. The z-coordinate is scaled by the individual thicknesses of the three layers. Under the conditions depicted here, the entire first layer is filled with liquid. Interfaces form at the upstream and the downstream front of the separation layer. The pressure difference across these interfaces is large, indeed so large that the fluid within the pore stays under tension. This is possible in a liquid that is confined in a small space and stays under a meniscus with a large curvature [14]. The pressure difference across the first layer is also large, therefore the mass flux is much larger than for the flow of the gaseous phase of the fluid through the same membrane. Conclusions Large mass flow differences may occur for opposite directions of the flow of a vapor close to saturation through an asymmetric membrane. Based on theoretical considerations, the mass flows of a vapor far from saturation through a membrane with a smallest pore size of 1 nm differs by a few percent. The difference occurs if the fluid is transported in a combination of Knudsen and viscous flow. The mass flows of a vapor close to saturation may differ by a factor up to seven for the two different orientations of the membrane. These large differences are caused by condensation and the flow of liquid through different parts of the membrane. Acknowledgments Partial financial support by Androsch International Consulting GmbH, by the OeAD GmbH, projects nr. CZ 8/212 and CZ 6/214, and by the Ministry of Education, Youth and Sports of the Czech Republic, grants 7AMB12AT1 and 7AMB14AT11, is gratefully acknowledged. References [1] W. Schneider: Acta Mech. Vol. 47 (1983), p. 15. [2] G. Hohenbichler and W. Schneider: Z. angew. Math. Mech. Vol. 71 (1991), p. T 475; Corrigendum ibid, p. 364. [3] T. Loimer: J. Membr. Sci. Vol. 31 (27), p. 17. [4] T. Loimer, P. Uchytil, R. Petrickovic, and K. Setnickova: J. Membr. Sci. Vol. 383 (211), p. 14. [5] H. Rhim and S.-T. Hwang: J. Colloid Interf. Sci. Vol. 52 (1975), p. 174. [6] K.-H. Lee and S.-T. Hwang: J. Colloid Interf. Sci. Vol. 11 (1986), p. 544.

[7] J.-G. Choi, D. D. Do, and H. D. Do: Ind. Eng. Chem. Res. Vol. 4 (21), p. 45. [8] P. Uchytil: J. Membr. Sci. Vol. 97 (1994), p. 139. [9] P. Uchytil and Z. Brož: J. Membr. Sci. Vol. 97 (1994), p. 145. [1] P. Uchytil: J. Mater. Sci. Vol. 31 (1996), p. 6293. [11] S. Thomas, R. Schäfer, J. Caro, and A. Seidel-Morgenstern: Catalysis Today Vol. 67 (21), p. 25. [12] A. Hussain, A. Seidel-Morgenstern, and E. Tsotsas: J. Porous Media Vol. 12 (29), p. 749. [13] G. Hohenbichler, A. Köppl, and W. Schneider: Acta Mech. Vol. 17 (1994), p. 21. [14] L. R. Fisher and J. N. Israelachvili: J. Colloid Interf. Sci. Vol. 81 (1981), p. 528.