Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Transport Phenomena in Heat and Mass Transfer
Transport Phenomena in Heat and Mass Transfer
Transport Phenomena in Heat and Mass Transfer
Ebook1,762 pages

Transport Phenomena in Heat and Mass Transfer

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Theoretical, numerical and experimental studies of transport phenomena in heat and mass transfer are reported in depth in this volume. Papers are presented which review and discuss the most recent developments in areas such as: Mass transfer; Cooling of electronic components; Phase change processes; Instrumentation techniques; Numerical methods; Heat transfer in rotating machinery; Hypersonic flows; and Industrial applications. Bringing together the experience of specialists in these fields, the volume will be of interest to researchers and practising engineers who wish to enhance their knowledge in these rapidly developing areas.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780444599797
Transport Phenomena in Heat and Mass Transfer

Related to Transport Phenomena in Heat and Mass Transfer

Mechanical Engineering For You

View More

Related categories

Reviews for Transport Phenomena in Heat and Mass Transfer

Rating: 5 out of 5 stars
5/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Transport Phenomena in Heat and Mass Transfer - J.A. Reizes

    (Singapore)

    NATURAL AND FREE CONVECTION

    APPLICATION OF THE ELECTROCHEMICAL MASS TRANSFER TECHNIQUE TO THE STUDY OF BUOYANCY-DRIVEN FLOWS

    H.D. Chiang and R.J. Goldstein,     Energy and Resources Laboratories, Industrial Technology Research Institute, Taiwan, ROC; Department of Mechanical Engineering, University of Minnesota, Minneapolis, MN 55455 USA

    ABSTRACT

    The applicability of an electrochemical mass transfer technique in studying buoyancy-driven convection is examined. Emphasis is placed on the copper deposition system. A detailed description of the system, the physical properties of the solutions used, and the related methodology are summarized. Justification of the analogy between the electrochemical technique and comparable heat transfer studies is presented.

    1 INTRODUCTION

    The mass transfer process involved in an electrochemical system was initially studied almost exclusively by electrochemists and chemical engineers. Their interests lay in understanding the physics and chemistry involved and applying mass transfer to electrochemical processing. The analogy between electrochemical systems and the corresponding heat transfer systems expanded the scope of the technique to a wide range of applications.

    The electrochemical system employs a diffusion-controlled electrolytic reaction (or pair of reactions) to study the desired transport phenomena. With an externally applied potential difference across two electrodes in the electrolytic solution, a current will flow from the anode to the cathode, generating a mass transfer process within the solution. This mass transfer could be by migration, diffusion, and, perhaps after exceeding a critical threshold, by convection. The study of the mass transfer coefficient at the cathode surface can be used to increase understanding of the mass transfer process in its own right or to infer corresponding heat transfer phenomena through analogy.

    Advantages of an electrochemical system over a conventional heat transfer system in the study of buoyancy-driven convection include: (1) high precision and local measurements are more easily made, (2) large Rayleigh numbers (Ra) can be achieved in a moderate-sized apparatus, (3) boundary conditions can be controlled better, and (4) sidewall conduction and radiation effects are eliminated. The electrochemical systems normally used operate with high Schmidt number (Sc) fluids limiting the analogy to high-Prandtl-number (Pr) fluids. Even so, for researchers in the heat and mass transfer communities the electrochemical technique provides a promising alternative as a measurement tool.

    A wealth of information on electrochemical systems is available. For readers interested in the fundamentals, books by Levich (1962) and Newman (1973) provide extensive coverage of the theoretical aspects of electrochemical mass transfer. On the application side, reviews by Mizushina (1971) and Wragg (1977) give in-depth surveys of the literature. These, however, do not provide the novice with sufficient information on the subtle know-how required to apply the technique properly. A detailed review paper by Selman and Tobias (1978) covers the various operating conditions and their constraints in a systematic fashion. Due to its broad scope, however, the paper left open some questions concerning the use of the technique, especially for those in the heat transfer field who are interested in the analogy between electrochemical systems and their heat transfer counterparts. In the following sections, a simplified description of the electrochemical systems used for the study of natural convection is given, along with property correlations and procedures for concentration measurement.

    2 THE ELECTROCHEMICAL METHOD

    A typical electrochemical system consists of an electrolytic solution as the working fluid and two electrodes, anode and cathode (there could be more than one cathode). Mass transfer is induced in the solution by applying an external electric potential difference across the anode and cathode(s). Positive ions (cations) of the electrolyte move toward the cathode while negative ions (anions) move toward the anode. The movement of the ions is controlled by: (1) electric migration due to the electric field, (2) diffusion because of ion-density gradient, and (3) convection, if the fluid is in motion. Migration is the movement of ions under the influence of an electric field. Fluid motion can be driven by pressure drop in forced flows or by density gradients in buoyancy-driven flow.

    With heat transfer, convection and diffusion processes are present, but there is no equivalent to migration. In order to use ionic transport as an analog to the heat transfer process, the ionic migration has to be made negligible. This is achieved by introducing a second electrolyte–the so-called supporting electrolyte. This is normally an acid or base with a concentration many times that of the active electrolyte, and which is selected such that its ions do not react at the electrodes over the range of potential difference used in the experiment. The excess of supporting electrolyte will sharply reduce the electric field in the bulk of the solution, and the migration effect on the charge carrier will only be a minor correction.

    Among the more commonly used electrolytic solutions are:

    A Cupric Sulphate-Sulphuric Acid solution (CuSO4-H2SO4-H2O)

    B Potassium Ferrocyanide-Potassium Ferricyanide-Sodium Hydroxide solution (K3]Fe(CN)6[-K4]Fe(CN)6[-NaOH-H2O)

    With a cupric-sulphate solution, copper is dissolved from the anode and deposited on the cathode (metal-deposition reaction). For the other system (also known as redox-couple system), only charge transfer occurs at the electrodes. The respective reactions at the cathode surface are:

    (1a)

    and

    (1b)

    The copper deposition system is usually employed for natural convection studies. Hence, emphasis will be placed on the copper system here, but the general principles apply to both systems.

    The mass transfer coefficient for species i is,

    (2)

    is the transfer flux of species i due to diffusion and convection, and ΔCi is the concentration difference of the species across the region of interest.

    The total flux at the cathode surface can be determined from Faraday’s Law

    (3)

    where I is the current density at the cathode surface, ni is the valence of the transferred ion, and F is Faraday’s constant.

    As mentioned above, the migration effect on the active electrolyte is negligible after the addition of the supporting electrolyte (H2SO4 when using CuSO4). This is, strictly speaking, valid only for the limiting case,

    Otherwise, some migration effect will exist. The migration flux can be related to the current density by the introduction of the transference number ti. The transference number, a function of the solution concentrations (see Section 6), is proportional to the migration flux:

    (4)

    Combining Eq. 4 with Eqs. 2 and 3 gives

    (5)

    Usually, the concentration difference is determined from the bulk and surface concentrations. The bulk concentration is usually assumed to be constant and can be measured by chemical analysis; however, the surface concentration is an unknown. In a heat transfer study, the temperature is usually continuous across a solid-fluid interface. Thus, the interface temperature can be determined from measurement on the solid side. Measurement of surface concentration is not that direct. This is resolved by using the limiting current condition. As the externally applied potential across the electrodes is increased, the current increases monotonically until a plateau - on a graph of current vs. voltage – occurs (cf. Fig. 1). For the cupric sulphate system, the concentration of copper ions at the cathode surface will be negligible at the limiting current. For the redox-couple system (ferri-ferro-cyanide), the concentration of [Fe(CN)6]³- at the cathode will be negligible at the limiting current.

    FIGURE 1 The limiting current potential

    The limiting current density is the maximum current density attainable at the cathode by the transport of the reacting ions. To understand the cause of this limit and its resulting plateau in the current-potential plot, the various components of the cell potential should be introduced.

    In an electrolytic solution, the driving force for the passage of current between the anode and cathode is the electric potential difference across them. This potential difference can be separated into three distinct parts: (1) ohmic potential, ΔΦohm, (2) concentration overpotential, η, and (3) surface overpotential, ζ. The overall cell potential difference can be written as

    (6)

    where subscripts a and c refer to anode and cathode conditions, respectively.

    The ohmic potential has its usual meaning here. It is due to the electrical resistance of the bulk solution. The addition of supporting electrolyte drastically reduces this resistance, and, thus, the ohmic potential difference is small. The surface overpotential is the potential difference across the solid-liquid interface, which is required to drive the surface reaction given in Eq. 1. This potential difference normally does not effect the rate of ion transfer and will be ignored here. Finally, the concentration overpotential is a potential difference caused by a concentration gradient within the solution. In an electrolytic solution, the concentration gradient usually occurs near the electrodes (i.e. across the concentration boundary layer). The concentration overpotential is given by

    (7)

    where subscripts s and b refer to surface and bulk conditions, respectively.

    For a metal-deposition reaction in a dilute aqueous solution, the anode reaction is almost unrestrained (unless the bulk concentration of cupric sulphate is greater than half its solubility), while the cathode reaction rate has an upper bound. From Eq. 7

    It takes an infinitely large potential difference to reduce the surface concentration of the reacting ions to zero; this limits the maximum ionic transfer rate from the solution to the cathode. In reality, of course, the concentration overpotential does not increase to infinity; a secondary reaction will take over at some higher potential, increasing the current density. In an aqueous solution, this secondary reaction is usually the hydrogen evolution reaction. This leads to the familiar S-shaped limiting-current plateau on the current-voltage plot as shown in Fig. 1.

    Measurement of the limiting-current density and knowledge that the cathode surface concentration of the reacting ions is close to zero at the limiting current permit calculation of the mass transfer coefficient from Eq. 5. From this, one can infer that measurement of localized limiting-current densities under constant surface potential will yield information on local mass transfer coefficients.

    3 REVIEW OF LITERATURE

    The nature of the electrolytical reaction implies the importance of convection on the current transfer at the electrodes. Early theoretical studies on natural convection at a vertical electrode (Agar, 1947; Wagner, 1949; Keulegan, 1951; Tobias et al., 1952) and convection at a rotating disc (Levich, 1947) led to systematic treatments of the subject by Levich (1962) and Newman (1973). The development of the limiting diffusion current technique (LDCT) for mass transfer measurements generated numerous applications to natural- and forced-convection flows. Extensive reviews are available (Mizushina, 1971; Wragg, 1977; Selman and Tobias, 1978). We shall restrict our discussion to natural-convection flows.

    Natural convection at vertical electrodes under laminar flow was studied by Wilke, Eisenberg, and Tobias (1953). They presented a detailed account of the various parameters studied and their data reduction procedure. They used a CuSO4-H2SO4-H2O system and correlated the overall Sherwood number

    (8)

    with K = 0.673. In a companion paper, Wilke, Tobias, and Eisenberg (1953) reported the above results together with results for three other systems. A summary of the correlations is given in Table 1, where K is the coefficient in Eq. 8. The authors presented an overall correlation for all the data with a constant K = 0.66.

    TABLE 1

    Natural Convection Mass Transfer Results from Wilke, Tobias, and Eisenberg (1953)

    Natural convection at upward-facing, horizontal electrodes has been studied by several investigators who considered overall correlations, the effect of electrode size and shape, flow patterns, and transient and periodic fluctuations (Fenech and Tobias, 1960; Tobias and Boeffard, 1966; Wragg, 1968; Wragg and Loombs, 1970; Lloyd and Moran, 1974; Patrick and Wragg, 1975; Antonini et al., 1978). The more general case of a single inclined electrode has also been studied (Fouad and Ahmed, 1969; Lloyd et al., 1972; Moran and Lloyd, 1975; Patrick et al., 1977). Many different flow regimes are observed on an inclined electrode. They range from laminar to turbulent plumes for horizontal or near-horizontal, upward-facing electrodes, to the formation of longitudinal vortices, and, finally, to boundary layer flow for near-vertical and downward-facing electrodes. Patrick, Wragg, and Pargeter (1977) present a conjectural map of the flow regimes on a Ra-θ, where θ is the angle of the plate from the horizontal. However, no confirmation of the transition boundaries is available. They also propose a g·sinθ scaling for overall mass transfer from single inclined surfaces over θ* < θ < 180° where θ* is the transition angle between longitudinal rolls and unicellular flow. Their correlation equation is

    (9)

    Turbulent-flow conditions have also been studied. Fouad and Ibl (1960) extended the vertical-electrode results to turbulent flows up to Ra=10¹⁵. Their results for 4×10¹³ < Ra < 10¹⁵ indicate that

    (10)

    This deviation from an expected 1/3-power law seems to be due to the averaging over laminar and turbulent regions. Local measurements of turbulent flows on vertical and inclined electrodes performed by Lloyd, Sparrow, and Eckert (1972) yielded the 1/3-power law

    (11)

    However, the coefficient Kt depends on θ in the domain studied: 45° < θ < 90°. Patrick, Wragg, and Pargeter (1977) measured overall turbulent mass transfer for 0° < θ < 50°. The resulting constant K also has a θ dependence.

    Mass transfer measurements for other geometries have also been performed. The studies include cylindrical and spherical electrodes (Schutz, 1963; Sedahmed and Nirdosh, 1990), multiple cylinders (Smith and Wragg, 1974; Wragg et al., 1975), rod-shaped vertical electrodes (Selman and Tavakoli-Attur, 1980), horizontal (Sedahmed and Shemilt, 1981) and vertical annuli (Sedahmed and Shemilt, 1982), inclined cylinders (Sedahmed and Shemilt, 1982), cuboids (Worthington et al., 1957) and cavities (Butterworth, 1983; Sommerscales and Kessemi, 1985). Double-diffusion problems involving simultaneous heat and mass transfer have also been studied (DeLeeuw Den Bouter et al., 1968; Wragg and Nasiruddin, 1974; Wragg and Patrick, 1974; Hamotani et al., 1985; Wang et al., 1989).

    Most of the above studies of natural convection involved overall measurements. Local measurements in natural convection have only been performed by a few investigators. Local mass transfer coefficients on a cylinder and a sphere were measured by Schutz (1963). Lloyd, Sparrow, and Eckert (1972) and Moran and Lloyd (1975) measured local mass transfer on vertical and inclined surfaces under turbulent conditions. Lateral distributions of mass transfer coefficient over an inclined surface under a longitudinal-roll flow regime were measured by Lloyd, Sparrow, and Eckert (1972).

    Another interesting feature is that almost all studies involve an isolated cathode, thus simulating a single heated surface. In these studies, the anode is generally much larger than the cathode; so the current density and concentration difference near the anode is relatively small. The electrochemical technique has also been applied to the study of enclosure problems (See, 1976; Sayer, 1977; Hogerton, 1980) and double diffusion in a vertical enclosure (Hamotani et al., 1985).

    4 MEASUREMENT OF THE LIMITING CURRENTS

    The limiting-current density (or densities) can be determined after the establishment of the limiting-current plateau. Before a discussion of the generation of the plateau, a few comments about the electrode potential are needed.

    4.1 The Electrode Potential and the Reference Electrode

    On first sight, this seems to be a trivial issue. Since total external potential across the electrodes is known. However, most researchers to date have been using a reference electrode in their measurements of potential. It is thus beneficial to discuss the concept briefly.

    Recall in Section 2 the components of the cell potential as given in Eq. 6. Now, the boundary conditions for the potential gradient term are not simply given by the external potential. The surface overpotentials have to be known before any analytical and numerical solutions can be obtained. In their experiments, electrochemists and chemical engineers usually try to measure only the cathode concentration overpotential, ηc, using a reference electrode. The measured overpotential is then used to generate the limiting-current plateau. Most early applications of the electrochemical technique to natural convection involve the use of the cathode alone, i.e. most early investigators tried to isolate the hydrodynamic effect of the anode such that the cathode behaves like a single isolated surface. This was well-justified in those early attempts to understand the physics of the electrochemical systems. A reference electrode was needed to measure the cathode concentration overpotential, which is then used for the analytical prediction of the limiting currents.

    In enclosure studies, the overall potential could be used, because accurate measurement of the electrode overpotential, though important in accurately defining the boundary conditions, is not necessary for the measurement of the limiting current. Using the overall potential might shift the limiting-current plateau along the potential axis, i.e. change the limiting potential value, but it would not alter the actual value of the limiting-current density. Hence, as long as a well-defined limiting-current plateau can be established, the use of a reference electrode is really not needed. Details about the construction and deployment of a reference electrode is available in Selman and Tobias (1978) and will not be covered here.

    4.2 DETERMINATION OF THE LIMITING-CURRENT PLATEAU

    The external potential difference and the electrode current density are the two variables for the formation of the limiting-current plateau. They, however, are not independent of each other. Two approaches are available (Selman and Tobias, 1978): (1) A galvanodynamic approach, where the current is increased by using a constant-current source. The limiting-current plateau is determined by the point at which a steep increase in the potential occurs for a very small increase in the current. (2) A potentiodynamic approach, where the potential is increased by using a constant-voltage source and the current is monitored. The limiting-current plateau is determined by the region where increases in potential results in a very small increase in the current.

    In both approaches, the controlled current or voltage may be increased in steps or continuously, i.e. a step change approach or a ramp change approach. Finally, the limiting-current condition could be imposed instantaneously by applying a single potential step. This last approach is used for local measurements and can only be applied, if the external potential drop needed for the establishment of the limiting current condition is known a priori.

    The choice of a limiting-current measurement approach out of the five different combinations mentioned above will usually depend upon the particular circumstances encountered. For some studies, it might only be a matter of preference; for others, it might be dictated by the experimental requirements.

    One other important point is that an electrochemical natural-convection experiment will usually start with the electrolytic solution initially at rest. The measurement of the steady-state mass transfer will thus only be available after sufficient time has elapsed and the transients have subsided. Hence, the transient time constant of the flow and mass transfer system must be known beforehand or determined in preliminary tests.

    A true horizontal plateau or constant current readings over the flat part of the S-shaped curve does not occur in practice. Instead, a small positive slope in the plateau region has to be tolerated, depending on the uncertainty tolerance of the overall measurement. For the copper deposition system used, the flat portion of the plateau is typically approximately 200 mV wide. The center of the plateau is between 400–500 mV at a copper concentration of 0.015 Molar and 700–800 mV at a copper concentration of 0.10 Molar.

    For the copper deposition system, hydrogen evolution is the secondary reaction that occurs at higher potential differences. Hydrogen evolution should be kept to a minimum, since it will increase the copper concentration of the bulk solution. The rate of hydrogen evolution increases rapidly toward the right-hand end of the plateau region and is responsible for the rapid increase in the current density at higher potential.

    Wilke, Eisenberg, and Tobias (1953) employed the galvanodynamic approach and used a step change of current settings. Various time intervals between current setting and potential reading (10, 30 & 50s) and between successive current settings (60 and 180s) were used. They reported a maximum deviation of 3.2% from their average values in the limiting current.

    Fenech and Tobias (1960) used a sudden potential step followed by a constant ramp change in voltage to minimize the time required to reach the limiting current. They encountered difficulty at a copper concentration of 0.01 Molar, reporting a limiting-current plot in the shape of a camel’s back. A camel back shape was reported by Somerscales and Kassemi (1985) for copper concentrations less than 0.05 Molar. Rates of potential increase of 1 mV/s and 0.5 mV/s were used for 0.1 and 0.05 Molar cupric sulfate, respectively. They mentioned the elimination of the camel back shape at a lower rate of increase, but they were too concerned about the increase in copper plating to lower the copper sulfate concentration and the rate of increase. For both studies, the total elapsed time to the generation of the limiting-current plateau was about 10–20 minutes.

    The use of a constant rate of change (ramp-change) in potential or current will only yield accurate results if the time constant (time to reach a steady state) of the flow system is less than the total elapsed time of the experiment. That is, the flow and concentration fields have to reach a steady state before the potential attains the limiting value. This, however, may not be true in inclined enclosures and other complex geometries, where the transient time needed for the establishment of core stratification could be of the order of hours or even days.

    The study of local mass transfer profiles and the turbulent behavior of the system cannot be done with a ramp change approach. A single potential step (corresponding to the limiting potential, which was determined in advance) would be needed. An appropriate potential is applied externally at time zero and maintained constant throughout the run during which local mass transfer readings are gathered. Typical experimental runs for turbulent measurements in a natural-convection study would last 2 to 4 hours.

    In using the copper deposition system, care should be taken to avoid prolonged runs, as these would lead to surface roughness and, thus, alter the boundary conditions. Run time of longer than about 10 hours is not recommended at a copper sulfate concentration of 0.02 Molar and should be reduced appropriately at higher concentrations. If the system under investigation has a longer transient time, the redox-couple system (Selman and Tobias, 1978) might be more appropriate. To be prudent, preliminary investigations should always be performed to verify the operating characteristics, especially when complex geometries are involved.

    5 PHYSICAL PROPERTIES

    From the previous sections, we established the measurement needs as (1) surface flux(es), (2) bulk and surface concentrations, and (3) physical properties. In natural-convection mass transfer, the dimensionless quantities of interest are the Sherwood number, the Rayleigh number, and the Schmidt number. Hence, the properties that are needed are the density, viscosity, diffusion coefficients and transference numbers. In general, a physical property, p (eg. ρ,μ,ti), is a function of solution concentrations and of temperature,

    (12)

    For the diffusion coefficient, the Stoke-Einstein equation gives the mobility product (Selman and Tobias, 1978)

    (13)

    A brief survey of the available property data for the supported copper deposition system will be given here. After their 1953 paper (Wilke et al., 1953) on the correlation of limiting currents under free convection at vertical plates, Eisenberg, Tobias and Wilke (1956) published a tabulation of their measured ρ & μ values at a sulfuric-acid concentration of 1.5 Molar. A few years later, Fenech and Tobias (1960) presented correlation equations for all properties at 22°C, but they did not provide the raw data used. Lloyd, Sparrow, and Eckert (1972) proposed correlations of the general form of Eq. 12. Their correlations for ρ & μ are based on the previous two references. The research group led by Professor Tobias performed extensive property measurements, and the results for ρ & μ at 25°C were published in an internal report (Arapkoske and Selman, 1971). Unfortunately, a copy of this report is now virtually impossible to obtain. Improved correlations (also at 25°C) are provided by Selman and Tavakoli-Attar (1980). Temperature-dependent correlations for ρ & μ are available from Professor Selman.

    In natural convection the driving force for flow is due to Δρ. To evaluate Δρ, knowledge of the surface concentrations of Cu++ and H+ is required. Even at the limiting current, the surface concentration of H+ is still an unknown. This is a secondary effect and can be approximated. Wilke, Eisenberg, and Tobias (1953) proposed an approximate procedure by equating the convective-diffusive flux of H+ to its migration flux at the surface. Selman and Newman (1971) analytically determined the limiting-current density and the sulfuric-acid concentration at the cathode surface for laminar natural convection at a vertical electrode. The applicability of their laminar-flow results to turbulent flows is uncertain.

    We next turn to the property that has created perhaps the most difficulty – the diffusion coefficient. The electrochemical system used is a multi-component solution. In a multi-component system, there is a diffusion coefficient for every ionic species. Cole and Gordon (1936) measured the molecular diffusivities of cupric sulfate in an aqueous solution of sulfuric acid. Diaphragm cells were used and all measurements were performed at 18°C. These data were used by Wilke, Eisenberg, and Tobias (1953), who tabulated the average diffusivities for their experimental runs. The range of temperatures of the runs was 18.0°C to 27.2°C. Fenech and Tobias (1960) proposed a correlation for copper diffusivity in 1960 using the Stoke-Einstein Equation (Eq. 13). All their property correlations were given at 22°C, and have since been used by many investigators.

    The mobility product correlation given by Fenech & Tobias for copper sulfate is

    (14)

    The validity of Eq. 14 was questioned by Arvia, Bazan and Carrozza (1966). They measured the effective diffusivity of copper sulfate in aqueous and aqueous-glycerol (glycerol was added to the solution to increase the viscosity) solutions with sulfuric acid using a rotating disk electrode. Since the limiting current density for a rotating disk electrode is known analytically, the only unknown in its expression, the diffusivity, can be determined from the experimentally measured limiting current. In forced convection, the analogy of the electrochemical mass transfer and heat transfer is almost exact (except for the small normal component of velocity at the electrode surface due to mass transfer). Hence, the resulting diffusivity data can be used for mass transfer measurements in other complex geometries.

    The range of concentrations in Arvia, Bazan, and Carrozza (1966) was: 0.010-0.077 Molar for copper sulfate; 0.477-1.752 Molar for sulfuric acid and 0-9.51 Molar for glycerol. They proposed

    The 16% maximum deviation is usually unacceptable. Moreover, closer inspection of their data revealed a dependence on the glycerol concentrations. For solutions with no glycerol, their data could be better represented by Wragg and Ross (1968)

    (15)

    with nine data points. Wragg and Ross (1968) used eq. 15 to re-evaluate their results for laminar flow in an annular duct; they found excellent agreement with theoretical prediction, thus suggesting the validity of using the rotating disk electrode method in obtaining the average diffusivity.

    Subsequently, the Berkeley group measured effective diffusivities of copper ions. Extending the concentrations of copper sulfate to cover the whole solubility range (up to 0.7 Molar). From these data, and also because of the need for higher copper sulfate concentrations in practical electrochemical processes, Selman and Tobias (1978) propoosed a more general correlation for the mobility product which was applicable for all copper sulfate concentrations:

    (16)

    Eq. 16, based on 56 data points, has a standard error of 8%. The effect of the addition of glycerol was also correlated for CCuSO4 < 0.1 Morlar, CH2SO4∼ 1.5 Molar, and Cglycerol < 6 Molar, with a standard error of 0.05:

    (17)

    A comparison of Eq. 14 with either Eq. 15 or 16 shows disagreement of about 20%. To explain this disagreement, the difference between molecular and ionic diffusivities was proposed (Newman, 1973; Selman and Tobias, 1978). This line of reasoning, however, does not resolve the issue, since ionic diffusivities cannot be measured directly in practice and theoretical values are only available for solutions of infinite dilution. Thus, no easy comparison could be made. A quick look at the literature reveals authors quoting the term ionic diffusivity while, at the same time, using Eq. 14 of Fenech and Tobias, which was supposed to correlate molecular diffusivities! After a careful review of the literature and a closer look at the raw data, our conclusion is that Eq. 14 seems to overestimate the copper diffusivity.

    The diaphragm cell technique might really measure the molecular diffusivity and is, therefore, not applicable to multi-component systems. Possibly, the data were an overestimation all along. Since that was the only diaphragm cell measurement to date, we feel that, without more measurements, focusing on the important point - the agreement with theoretical prediction - would be more productive.

    Eq. 15 apparently gives a fairly good estimate of the effective diffusivity for C < 0.1 Molar, since it works for forced-convection flow (Wragg and Ross, 1968). Eq. 16, though having a wider range of applicability, also has a larger uncertainty. Thus, unless the accuracy in the effective diffusivity correlation can be improved, it will be a major source of uncertainty in the evaluation of property values at present.

    The availability of a correlation for the effective diffusivity in 1966 (Arvia et al., 1966) and the extension in 1978 (Selman and Tobias, 1978) did not attract widespread notice. For the many natural-convection applications since then, only the group at the University of Exeter, U.K. has adopted the new correlations. The group used the new diffusivities to re-evaluate their horizontal-surface results (Wragg and Loombs, 1970), and several other studies of horizontal and inclined surfaces (Patrick and Wragg, 1975; Patrick et al., 1977), multiple cylinders (Smith and Wragg, 1974; Wragg et al., 1975) and double-diffusion problems (Wragg and Nasiruddin, 1974; Wragg and Patrick, 1974).

    Besides the uncertainty in diffusivity, the evaluation of Δρ could contribute significantly to potential error. Thus, care should be taken while interpreting and applying the results from natural-convection studies.

    Finally, the evaluation of the physical properties of the solution requires the knowledge of the bulk concentrations of the electrolytes. The solution preparation and one technique of concentration determination for the copper deposition system are summarized in Appendix A.

    6 DATA REDUCTION CORRELATIONS

    As mentioned in Section 5, various property correlations and interpolation procedures are employed to calculate the Sherwood number and Rayleigh number. After some comparative studies, the following set of correlations and the following procedure were chosen.

    For the solution density and viscosity, the temperature-dependent correlations provided by Professor Selman were selected. They are

    (18)

    (19)

    where ΔT=(T-25.0), T in °C, CCuSO4 and CH2SO4 in Molar.

    For the diffusion coefficient, the following two correlations were chosen (Selman and Tobias, 1978). For CCuSO4< 0.09 Molar.

    (20a)

    and for CCuSO4>0.09 Molar

    (20b)

    The determination of Δρ requires knowledge of the surface concentrations. At the limiting-current condition, the cathode surface concentration of cupric sulfate is close to zero, and the concentration of the sulfuric acid at the cathode surface is calculated by the following procedure.

    Since sulfuric acid is an inactive species, we can equate its migration flux (Eq. 4) to its convection-diffusion flux (Eq.2) at the surface to give

    (21)

    Evaluating Ilim from Eq. 5 gives

    (22)

    The ratio of the mass transfer coefficient of cupric sulfate to that of the sulfuric acid is flow-dependent. Consider a flow with the following relationship:

    (23)

    Then

    (24)

    Substituting Eq.24 into Eq.22 gives

    (25)

    For the evaluation of Eq. 25, the following three correlations are also needed (Fenech and Tobias, 1960):

    (26)

    (27)

    (28)

    Finally, if the test cell and solution can be maintained at a constant 25°C throughout the experimental run, the following more accurate and simpler correlations for the density and viscosity can be used (Selman and Tavakoli-Attar, 1980)

    (29)

    (30)

    with standard errors of 0.0018 g/cm³ and 0.0076 poise, respectively.

    7 VERIFICATION OF THE ANALOGY ASSUMPTION

    One argument for the validity of the property relation would be agreement of measurement with known results. For natural-convection flow, theoretical results for the laminar boundary layer flow along a vertical, flat plate are available; the local Sherwood number is given by

    (31)

    At a Schmidt number of 2000, Kx= 0.500. If the electrochemical technique is to be of practical value, it should be able to reproduce this result (though the reverse may not necessarily follow). Since this paper is intended as a summary of the technique for applications to natural-convection flows, only the comparison with this theoretical result will be presented here.

    Fig. 2 presents the limiting-current plateau for a single, vertical, flat plate, and Fig. 3 shows local Sherwood number measurements and their comparison with Eq. 31. The best 1/4-power law fit of the 28 data points gives a constant Kx= 0.507 with a standard deviation of 0.013. Such excellent agreement suggests the validity of the analogy and the reasonableness of the property correlations, at least for the range studied. The experimental apparatus for the study will be presented in a companion paper (Chiang et al., in progress).

    FIGURE 2 Limiting current plateau for a single vertical flat plate

    FIGURE 3 Local Sherwood number vs. Rayleigh number for a single vertical flat plate

    NOMENCLATURE

    SUBSCRIPTS

    APPENDIX A

    TEST SOLUTION PREPARATION AND CONCENTRATION DETERMINATION

    For someone without proper exposure to the techniques of analytical chemistry, handling the acidic test solution and the process of titration and standardization can be intimidating. In fact, easier ways for the determination of the cupric-sulfate concentration exist, e.g. using atomic absorption (Hogerton, 1980). However, the equipment involved is rather expensive and usually could not be justified for modest investigations.

    The method presented here for the determination of the bulk concentration of the cupric sulfate and sulfuric acid of the test solution is a standard titration technique. The sulfuric-acid concentration is determined through titration with a sodium hydroxide (NaOH) solution of known concentration, and the cupric sulfate concentration is determined by iodometric titration using sodium thiosulfate (Na2S2O3). Since sodium hydroxide and sodium thiosulfate in solid forms will unavoidably absorb water vapor, their concentrations cannot be readily predetermined through weighing. The concentrations have to be separately standardized using primary standard chemicals.

    In preparation of the CuSO4-H2SO4-H2O test solution, weighing is used to get the approximate concentration. (CuSO4 crystal comes in the form of CuSO4.5H2O). A typical H2SO4 concentration used is about 1.5 Molar. This is achieved by diluting concentrated 18 Molar (36 Normal) sulfuric acid into about 11 parts of distilled water. It should be emphasized that one should always add acid to water, never the other way around. Adding water to acid will splash acid out of the container, because acid dilution releases an enormous amount of thermal energy.

    The amount of cupric sulfate to be added to the acid solution depends on the required concentration. The water solubility of cupric sulfate is about 1.4 Molar; hence, the bulk cupric sulfate concentration should be kept below 0.7 Molar, if crystallization at the anode surface is to be avoided. In electrochemistry, high cupric sulfate concentration is usually desirable. In mass transfer applications, especially when the heat transfer analogy is also desired, the cupric sulfate concentration should be kept as low as possible so as not to significantly alter the surface geometry. A compromise between this and the increasing uncertainty in Δρ when the cupric-sulfate concentration is lowered leads to a usable range of approximately 0.015 to 0.1 Molar. The upper bound may be somewhat extended, but care should be exercised in the interpretation of the results when using a higher concentration.

    After choosing a cupric-sulfate concentration, the appropriate amount of cupric-sulfate crystal can be weighed by a simple calculation involving the final solution volume and the molecular weight of cupric sulfate. The properly mixed solution is then allowed to settle. Constant stirring is needed to enhance faster mixing and dissipation of excess heat generated from acid dilution. The experience from this study is to add the acid in several batches and allow some time for the solution to cool down. Several samples from different depths of the solution bath are collected and titrated to determine concentration. The stirring process is continued, until the concentration variation within the solution is less then the uncertainty of the titration. This process can take up to a few days depending on the volume of solution used.

    Before the sulfuric-acid concentration can be determined, the titrating sodium hydroxide concentration needs to be standardized. This is achieved by weighing a known fixed amount of analytical-grade potassium biphthalate, dissolving it in distilled water through heating, and then titrating it using the prepared sodium hydroxide solution. The preparation and standardization procedure for sodium hydroxide is shown in Exhibit A1.

    Similarly, sodium thiosulfate (Na2S2O3) is used to determine the concentration of cupric sulfate, and its own concentration is standardized using the primary standard potassium dichromate. Exhibit A2 shows the preparation and standardization of sodium thiosulfate.

    With the concentrations of the two titrating chemicals known, the concentrations of the test solution can be determined. Exhibit A3 presents the required procedure. The remaining paragraphs provide a sample calculation of the solution preparation. The test solution concentrations chosen are about 0.015 Molar of cupric sulfate and 1.5 Molar of sulfuric acid. The final volume of the test solution is about 25 liters. Reagent A.C.S chemicals and distilled water (sometimes boiled to remove the dissolved CO2) are used.

    Preparation of Solution

    a) H2SO4Use 18 Molar concentrated sulfuric acid

    Volume of H2SO4 needed = (25 × 1.5)18 = 2.1 liters

    b) CuSO4Use cupric sulfate crystal (CuSO4.5H2O) Molecular weight = 249.68 Required weight = 0.015 × 25 × 249.68 = 93.6 gm

    Preparation of NaOH

    To improve titration accuracy, i.e. to use the largest volume of NaOH during titration with a 50ml burette, leads to a NaOH concentration of about 2.0 Molar. A 2-liter supply of NaOH is prepared each time from dry, solid NaOH pallets.

    Note: NaOH pallets will readily absorb water and release a vast amount of thermal energy; thus care should be exercised while dissolving the pallets. Also, to obtain carbonate-free solution, boiled and cooled distilled water is used (to remove dissolved carbon dioxide).

    Phenolphthalein

    Used as indicator in the NaOH titration. Prepared by adding 0.1 gm of ph-ph to 100 ml of methanol.

    Preparation of Na2S2O3

    Chosen concentration is about 0.1 Molar. A one-liter supply is prepared.

    Molecular Weight = 248.19

    Required Weight = 0.1 × 1.0 × 248.19 = 24.8 gm

    Also add about 0.1 gm of Na2CO3 to the solution.

    Potassium Iodide (KI)

    Prepared by adding about 120 gm of KI to 1 liter of distilled water.

    Starch Solution

    Take 2 gm of starch soluble, add 25 ml of cold distilled water to dissolve. Separately heat 1 liter of distilled water to boiling point, add the starch solution, let boil for 2 minutes, and finally add 0.1 gm of boric acid crystal.

    Starch is quite unstable; even with the boric acid addition, it might not be usable after several days.

    Other Chemicals

    The two primary standards, potassium biphthalate and potassium dichromate, are accurately weighed using a precision balance to determine the amount used during standardization. Three other chemicals are required in concentrated formml: HCI, NH4OH and acidic acid glacial (CH3COOH). See exhibits A2 and A3 for their use.

    Appendix B1 PREPARATION AND STANDARDIZATION OF NaOH

    Preparation

    Mol. wt. of NaOH = 39.998, need about 80 g/1 for 2 Molar

    Weigh NaOH pellets (in gm) in a 150 ml beaker

    Initial container wt. = __________

    Weigh NaOH pellets (in gm) in a 150 ml beaker

    Initial container wt. = __________ __________

    Total wt. = __________ ____________

    Container wt. = __________ __________

    Chemical wt. = __________ __________

    Mix the pellets with required volume of boiled and cooled distilled (B-C-D) water, allow to cool.

    Standardization

    Standardize with Potassium Biphthalate (mol. wt. = 204.22)

    Weigh 8-9 gms in each 250 ml flask

    Flask # _____ _____ _____

    Total wt. = __________ __________ __________

    Flask wt. = __________ __________ __________

    Chemical wt. = __________ __________ __________

    Mix with 100 ml of B-C-D water, heat to dissolve, allow to cool, add 3 drops of ph-ph as indicator.

    Titrate with 50 ml burette, color changes from clear → pink

    Expected Vol. (ml)

    Appendix B2 PREPARATION AND STANDARDIZATION OF SODIUM THIOSULPHATE

    Preparation

    Mol. wt. of Na2S2O3·5H2O = 248.17

    To titrate 0.1 Molar Cu2SO4 solution, need about 25 g/1 of Na2S2O3 and 0.1 g/1 of Na2CO3 (Sodium Carbonate). Weigh Na2S2O3 in 125 ml flask and Na2CO3 in small glass container.

    Mix both chemicals with B-C-D water

    Standardization

    Standardize with Potassium Dichromate (K2Cr2O7, Mol. wt. = 294.19)

    Weigh about 0.1 gm in each 250 ml flask

    Add to each flask about 25 ml of KI solution (120 g/1 of KI) and about 8 ml. of conc. HCL. Dilute to about 100 ml (B-C-D)Titrate with 20 ml burette until color changes: dark brown → yellow green, then add about 5 ml of starch solution, color changes: yellow green → dark blue, titrate carefully until color changes: dark blue → clear green.

    APPENDIX B3 DETERMINATION OF H2SO4 AND CuSO4 MOLARITIES

    Molarity of H2SO4 (Mol. wt. = 98.07)

    Use 3 250 ml flasks, put in each:

    exactly 25 ml of test solution,

    about 25 ml of boiled distilled water,

    and 1 drop of Phenolphthalein (ph-ph)

    Titrate with 50 ml burette. Rinse and fill with NaOH. Titrate until:

    blue→ cloudy blue → pink-green blue

    (reading #1) (reading #2)

    Molarity of NaOH (A) = ____ M; Approx. H2SO4 Molarity (B) = ____M

    Est. vol. of NaOH needed = 50 × (B) / (A) = ____ ml

    Molarity of CuSO4

        Added to each flask used above

    5 drops of cone. NH4OH,

    5 ml of conC. glacial Acetic Acid, and

    25 ml KI solution (3g KI/25 ml)

    Titrate with 5 ml burette. Rinse and fill burette with Na2S2O3

    M; Cu++Molarity (B)≈ ____M Est. Vol. of Na2S2O3 needed =25 × (B) / (A) =____ml

    REFERENCES

    Agar, J.N. Diffusion and Convection at Electrodes. Discussion of Faraday Soc. 1947; 1:27–37.

    Antonini, G., Guiffant, G., Geiger, D. Periodic Fluctuation of Free Convection Mass Transfer from Horizontal Surface. Letter in Heat Mass Transfer. 1978; 5:187–195.

    Arapkoske, S.K., Selman, J.R.Transport Properties of CuSo - H SO -HO Solutions at 25°C, UCRL-20510. UC-Berkeley: Lawrence Radiation Laboratory, 1971.

    Arvia, A.J., Bazan, J.C., Carrozza, J.S.W. Electrochemical Study of the Diffusion of Cupric Ion in Aqueous and Aqueous-Glycerol Solutions Containing Sulphuric Acid. Electrochemica Acta. 1966; 11:881–889.

    Butterworth, M.G., 1983, Natural Convection Electrochemical Mass Transfer from Cubic Cavities and Vertical Cylinders, M.S. Thesis, Univ. of Minnesota, Minneapolis.

    Chiang, H.D., Goldstein, R.J., and Khan, V., (in progress).

    Cole, A.F.W., Gordon, A.R. The Diffusion of Copper Sulfate in Aqueous Solutions of Sulfuric Acid. J. Phy. Chem. 1936; 40:733–737.

    DeLeeuw Den Bouter, J.A., DeMunnik, B., Heerties, P.M. Simualtaneous Heat and Mass Transfer in Laminar Free Convection from a Vertical Plate. Chem. Eng. Sci. 1968; 23:1185–1190.

    Eisenberg, M., Tobias, C.W., Wilke, C.R. Selected Physical Properties of Ternary Electrolytes Employed in Ionic Mass Transfer Studies. J. Electrochem. Soc. 1956; 103:413–416.

    Fenech, E.J., Tobias, C.W. Mass Transfer by Free Convection at Horizontal Electrodes. Electrochemica Acta. 1960; 2:311–325.

    Fouad, M.G., Ahmed, A.M. Mass Transfer by Free Convection at Inclined Electrodes. Electrochemica Acta. 1969; 14:651–666.

    Fouad, M.G., Ibl, N. Natural Convection Mass Transfer at Vertical Electrodes under Turbulent Flow Conditions. Electrochemica Acta. 1960; 3:233–243.

    Hamotani, Y., Wang, L.W., Ostrach, S., Jiang, H.D. Experimental Study of Natural Convection in Shallow Enclosures with Horizontal Temperature and Concentration Gradients. Int. J. Heat Mass Transfer. 1985; 28:165–173.

    Hogerton, P., 1980, An Experimental Study of Mass Transfer by Natural Convection at Moderate Rayleigh Number in a Horizontally Oriented Electrochemical Cell, M.S. Thesis, Univ. of Minnesota, Minneapolis.

    Keulegan, G.H. Hydrodynamics of Cathode Films. J. of Research (NBS). 1951; 47:156–169.

    Levich, B. The Theory of Concentration Polarisation. Discussion of Faraday Soc. 1947; 1:37–43.

    Levich, V.G., 1962, Physicochemical Hydrodynamics, Prentice-Hall, N.J.

    Lloyd, J.R. and Moran, W.R., 1974, Natural Convection Adjacent to Horizontal Surface at Various Planform, ASME Paper 74-WA/HT-66.

    Lloyd, J.R., Sparrow, E.M., Eckert, E.R.G. Laminar, Transition and Turbulent Natural Convection Adjacent to Inclined and Vertical Surfaces. Int. J. Heat Mass Transfer. 1972; 15:457–473.

    Lloyd, J.R., Sparrow, E.M., Eckert, E.R.G. Local Natural Convection Mass Transfer Measurements. J. Electrochem. Soc. 1972; 119:702–707.

    Mizushina, T. The Electrochemical Method in Transport Phenomena. Adv. in Heat Transfer. 1971; 7:87–161.

    Moran, W.R., Lloyd, J.R. Natural Convection Mass Transfer Adjacent to Vertical and Downward-Facing Inclined Surfaces. J. Heat Transfer. 1975; 97:472–474.

    Newman, J. S., 1973, Electrochemical Systems, Prentice-Hall, N.J.

    Patrick, M.A., Wragg, A.A. Optical and Electrochemical Studies of Transient Free Convection Mass Transfer at Horizontal Surface. Int. J. Heat Mass Transfer. 1975; 18:1397–1407.

    Patrick, M.A., Wragg, A.A., Pargeter, D.M. Mass Transfer by Free Convection During Electrolysis at Inclined Electrodes. The Canadian J. Chem. Eng. 1977; 55:432–438.

    Sayer, E., 1977, Turbulent Natural Convection Mass Transfer Across Inclined Fluid Layers, M.S. Thesis, Univ. of Minnesota, Minneapolis.

    Schutz, G. Natural Convection Mass-Transfer Measurements on Spheres and Horizontal Cylinders by an Electrochemical Method. Int. J. Heat Mass Transfer. 1963; 6:873–879.

    Sedahmed, G.H., Nirdosh, I. Free Convection Mass Transfer at Horizontal Cylinders with Active Ends. Int. Comm. Heat Mass Transfer. 1990; 17:355–366.

    Sedahmed, G.H., Shemilt, L.W. Natural Convection Mass Transfer in Horizontal Annuli. Letter in Heat Mass Transfer. 1981; 8:515–523.

    Sedahmed, G.H., Shemilt, L.W. Free Convection Mass Transfer in Vertical Annuli. Chem. Eng. Commun. 1982; 14:307–316.

    Sedahmed, G.H., Shemilt, L.W. Natural Convection Mass Transfer at Cylinders in Different Positions. Chem. Eng. Sci. 1982; 37:159–166.

    See, D.L., 1976, Natural Convection Mass Transfer in a Horizontal Fluid Layer, M.S. Thesis, Univ. of Minnesota, Minneapolis.

    Selman, J.R., personal communication.

    Selman, J.R., Newman, J. Free-Convection Mass Transfer with a Supporting Electrolyte. J. Electrochem. Soc. 1971; 118:1070–1078.

    Selman, J.R., Tavakoli-Attar, J. Free Convective Mass Transfer to a Rod-Shaped Vertical Electrode. J. Electrochem. Soc. 1980; 127:1049–1055.

    Selman, J.R., Tobias, C.W. Mass Transfer Measurements by the Limiting Current Technique. Adv. Chem. Eng. 1978; 10:211–318.

    Smith, A.F.J., Wragg, A.A. An Electrochemical Study of Mass Transfer in Free Convection at Vertical Arrays of Horizontal Cylinders. J. Applied Electrochem. 1974; 4:219–228.

    Somerscales, E.F.C., Kassemi, M. Electrochemical Mass Transfer Studies in Open Cavities. J. Applied Electrochem. 1985; 15:405–413.

    Tobias, C.W., and Boeffard, A.L., 1966, Ionic Mass Transfer by Free Convection at Horizontal Electrodes, 17th General Meeting of CITCE, Sept. 5–9, 1966, Tokyo, pp. 23–24.

    Tobias, C.W., Eisenberg, M.A., Wilke, C.R. Diffusion and Convection in Electrolysis–A Theoretical Review. J. Electrochemical Soc. 1952; 99:359C–365C.

    Wagner, C. The Role of Natural Convection in Electrolytic Processes. Trans. Electrochemical Soc. 1949; 95:161–173.

    Wang, L.W., Chai, A.T., and Sun, D.J., 1989, Convective Flows in Enclosures with Vertical Temperature or Concentration Gradients, 27th Aerospace Sciences Meeting, AIAA, Reno, Jan. 1989, NASA Technical Memorandum 101373, AIAA-89–0069.

    Wilke, C.R., Eisenberg, M., Tobias, C.W. Correlation of Limiting Currents under Free Convection Condition. J. Electrochemical Soc. 1953; 100:513–523.

    Wilke, C.R., Tobias, C.W., Eisenberg, M. Free-Convection Mass Transfer at Vertical Plates. Chem. Eng. Prog. 1953; 49:663–674.

    Worthington, D.H., Patrick, M.A., Wragg, A.A. Effect of Shape on Natural Convection Heat and Mass Transfer at Horizontally Oriented Cuboids. Chem. Eng. Research and Design. 1987; 65:131–138.

    Wragg, A.A. Free Convection Mass Transfer at Horizontal Electrodes. Electrochemica Acta. 1968; 13:2159–2165.

    Wragg, A.A. Applications of the Limiting Diffusion Current Technique in Chemical Engineering. The Chem. Engineer (London). 1977; 39–44. [Jan. 1977 (316)].

    Wragg, A.A., Loombs, R.P. Free Convection Flow Patterns at Horizontal Surface with Ionic Mass Transfer. Int. J. Heat Mass Transfer. 1970; 13:439–442.

    Wragg, A.A., Nasiruddin, A.K. Ionic Mass Transfer by Free Convection with Simultaneous Heat Transfer. Electrochemica Acta. 1974; 18:929–931.

    Wragg, A.A., Patrick, M.A. On the Structure of Free Convective Flow in Simultaneous Heat & Mass Transfer above Horizontal Electrodes. Electrochemica Acta. 1974; 19:929–931.

    Wragg, A.A., Patrick, M.A., Mustoe, L.H. Transient and Optical Studies of Free Convective Mass Transfer During Electrodeposition at Single and Multi-Cylinder Cathodes. J. Applied Electrochem. 1975; 5:359–361.

    Wragg, A.A., Ross, T.K. Diffusivity and Ionic Mass Transfer in The Cupric Sulphate System. Electrochemica Acta. 1968; 13:2192–2194.

    UNSTEADY HEAT TRANSFER IN DIFFERENTIALLY HEATED CAVITIES

    John C Patterson and Steven W Armfield,     Centre for Water Research, University of Western Australia, Nedlands, Western Australia, 6009

    ABSTRACT

    The thermal boundary layers on the vertical walls of a differentially heated cavity at early time are known to exhibit a complex travelling wave during growth to steady state. The signal is observed in both the thermal boundary layers and the horizontal intrusions emanating from the downstream ends of the layers, with the result that measures of the total heat transfer across the the cavity also display the signal. In addition, the heat transfer is in some cases also subject to a lower frequency oscillation, previously linked to the first mode internal wave triggered by the interaction of the intrusions with the far wall boundary layer. Although there had been a number of numerical demonstrations of both of these effects, until recently no experimental verification was available, and that was for a single case in the parameter space. This paper briefly reports the results of a number of further experiments with increasing Prandtl number which show the dependence on Pr of both effects. Further, the high frequency signals are shown to be linked to the stability of the thermal boundary layer, and the low frequency signal is shown to follow the scaling previously introduced.

    1 INTRODUCTION

    Unsteady natural convection in rectangular cavities with differentially heated sidewalls is a heat transfer process which is relevant to many geophysical and industrial applications; horizontal transport in reservoirs and lakes, atmospheric circulation, reactor cooling systems, crystal growth procedures, and aircraft fuel containers are typical examples. In many of these cases, the application of the horizontal temperature gradient is unsteady in some sense, and it is the transient response of the flow and heat transfer characteristics which is of interest. In particular, if the forcing is periodic, as in many geophysical examples, the interaction of the response time scales with the forcing time scales may profoundly influence the heat transfer characteristics of the system. Consequently, in the last decade there has been substantial interest in the development of natural convection flows following a change to the imposed thermal boundary conditions.

    In particular, the case of sudden heating and cooling of the opposing end walls of a rectangular cavity containing an isothermal fluid initially at rest has been examined in some detail, beginning with the scaling and numerical analysis by Patterson and Imberger (1980). This paper classified the flow development into several regimes, depending on the relative values of various combinations of the Rayleigh number (Ra), the Prandtl number (Pr) of the fluid, and the aspect ratio (A) of the cavity. Of special interest was the numerical prediction of a decaying oscillatory approach to steady state for Ra > Pr⁴ A−4.

    This oscillatory signal was strongly evident in the value of the Nusselt Number (Nu) integrated over the vertical extent of the cavity at the centre line. The presence of the oscillations meant that, transiently, heat was being stored in or depleted from either end of the cavity alternatively, which may be significant in some of the applications mentioned above.

    Although there were many numerical observations of the presence of the oscillatory behaviour (eg. Staehle and Hahne, 1982; Gresho et. al., 1980; Schladow et. al., 1989; Schladow, 1990), it was not until the recent paper of Patterson and Armfield (1990) that the oscillations were confirmed experimentally. Based on the experimental and numerical results presented, Patterson and Armfield were able to identify these waves as first mode internal waves resulting from disturbance of the stratification in the cavity.

    The experimental results and some of the numerical results above also showed an indication of a much higher frequency oscillatory component superimposed on the lower frequency approach to steady state. This component appeared in all cases in two short-lived discrete bursts early in the flow development. The first burst of high frequency signal was associated with the start up of the thermal boundary layer on each of the vertical walls, and the second burst with the impact of the horizontal intrusions emanating from the vertical walls with the vertical boundary layers on the opposite walls (Patterson and Armfield, 1990). The high frequency signal appeared to be of similar character to the waves observed numerically by LeQuere and Alziary de Roquefort (1985) in the case of a sudden increase in Rayleigh number from an initial steady flow. In that case at particular Ra, Pr, A values, a quasi steady periodic flow was obtained, with high frequency waves travelling around the cavity. Paolucci and Chenoweth (1989) also obtained solutions to this problem, and were able to simulate the onset of chaotic flow as these high frequency waves became unstable. Paolucci and Chenoweth identified these waves with the concept of the downstream wave train shed by an internal hydraulic jump as the horizontal intrusions exited from the emergent corners, an idea first suggested by Ivey (1984). However, Schladow (1990) and Patterson and Armfield (1990) showed that there were inconsistencies with this

    Enjoying the preview?
    Page 1 of 1