You are on page 1of 18

GRAVITY EFFECTS ON FLOW BOILING

GAUTAM WUDDI
ABSTRACT

This review of Gravity effects on Flow Boiling attempts to summarize the most promising results of the research performed in this area. While a reasonable amount of data has been accumulated over the years for microgravity, there is very little for hypergravity. This may be because the major focus of most of the investigations has been to clarify ow mechanisms and behaviour in reduced gravity environments. Another reason is the sheer expense and difculty of collecting quality data in parabolic ights. Future experiments on the International Space Station might resolve some of the difculties involved, besides giving very high quality data.
INTRODUCTION

The quantum jump in research being conducted in space environments and the advanced microprocessor technology being used in endurance military aircraft has exposed the need for better heat transfer methods than the single phase cooling thus far adopted. The focus has now fallen upon heat transfer using two phase ow cooling methods wherein both the sensible as well as latent energy of the uid or refrigerant can be leveraged. Two phase ows are also very conducive to transporting heat at nearly isothermal conditions over long distances. Both of these characteristics give them very high heat transfer coefcients. The fact that the latent heat is being utilized allows us to reduce the volumetric ow rate and thus the pumping requirements, which is critical in the applications mentioned above. The major parameters that have been considered in this study are the different ow regimes that occur in a microgravity environment, the criteria for transition between different ow regimes, the heat transfer coefcients and correlations for the frictional pressure drop and the critical heat ux.
Basic Terminology:

For a channel with cross-sectional area A, the mass ux or mass velocity G is given by G = (m/t)/A. The volume fraction is the ratio of the volume of a phase over the the total volume of ow. If the volume consists of a cross sectional area times the length, the volume fraction may be simplied to an area fraction. The vapor phase volume fraction is referred to as the void fraction, denoted by or . The slip is the ratio of the phase velocities - if both phases are moving with the same velocity, i.e., slip is equal to 1, then the ow is referred to as homogenous.

The equilibrium quality ((h - hf)/hfg) may be thought of as a measure of the degree of superheating or sub-cooling of a uid. In conditions where thermodynamic equilibrium exists between the phases, the quality corresponds to the ow fraction of the vapor; it can also be referred to by the term dryness fraction. The volumetric ux of a phase is dened as the volumetric ow rate of the phase over the total ow area. Physically, it has the units of velocity and may be thought of as the velocity that a phase has if it is owing through the channel alone. It is also more commonly referred to as the supercial velocity. Several ow regime maps have been developed for terrestrial gravity conditions, which help with keeping the ow in a recommended regime during design of equipment. Usually bubbly and annular ow are preferred and slug ow is avoided since it causes vibrations in the channels which may result in structural damage.
Dimensionless numbers:

Jakob number: It is the ratio between the sensible and the latent heat absorbed during a phase change process. Ja = Cp(Ts - Tsat)/hfg

Bond number: It is the ratio between the gravitational and the surface tension forces. B = d2/

Weber number: It is the ratio between the forces of inertia and surface tension. We = U2d/

Reynolds number: It is the ratio between the inertial and viscous forces. Re = Ud/

Suratman number: It is the ratio of surface tension to the momentum dissipation in a uid. It is also called the Laplace number and is used as a characteristic length in boiling. Su = La = L/2. Considering the supercial phase velocities instead of the mean ow velocities in the above dimensionless quantities give rise to the supercial dimensionless numbers, denoted with a subscript S.

THEORETICAL INVESTIGATIONS

The ow is said to be in the bubbly ow regime when the size of the vapor bubbles is less than or equal to that of the diameter of the tube. The ow is designated as slug ow when the size of the bubbles are greater in length than the diameter of the tube. Annular ow is the condition where the liquid never bridges the tube [1]. Although there has been one more regime - frothy slug annular ow or transition regime reported in microgravity ows, the three mentioned above are the widely accepted. The transition of ow from bubbly to slug occurs when the bubble concentration and size is such that adjacent bubbles come into contact. The conventional treatment of bubble coalescence requires that there be a liquid layer trapped between the two coalescing bubbles, and when the liquid lm drains out, over a certain time, an instability mechanism will result in lm rupture and bubble coalescence [2]. In a microgravity environment, the only forces that inuence motion and coalescence of bubbles is the motion induced by turbulence. Kamp et al.[3] conducted a study in which a numerical model for bubble coalescence on the basis of the collision frequency and drainage time are veried using experimental results conducted in a microgravity environment. A conclusion of their investigation is that bubble growth due to coalescence between inlet and outlet diminishes with increasing liquid ow rate. Slug ow is characterized by a plug of liquid, which is followed by a long Taylor bubble. The liquid plug may have smaller bubbles dispersed in it. The diameter of the Taylor bubble is close to that of the tube diameter and liquid is forced to ow around the bubble. If a co-ordinate system which moves along with the Taylor bubble is considered, the liquid lm around the bubble ows backwards - whether for terrestrial gravity or for microgravity conditions - and a conventional force analysis may be performed, such as that for vertical upward ow under gravity, with the surface tension taken into consideration[4]. A numerical solution using the Runge-Kutta method reveals that the surface tension may be neglected in the range where the lm thickness is less than about one-third of the pipe radius. The surface tension cannot, however, be neglected at the nose of the Taylor bubble. From the knowledge of the lm thickness prole, estimation of the pressure drop can be performed by considering a control volume around a slug. Pu = uglu + (sD/A)ls fSf/A dz

where the average slug density is u = u G + (1 - u) L. Dukler also attempted to theoretically estimate the values of all the parameters for annular ow, using a control volume approach. He then extended the ow regime map for 1-g to microgravity.


Data from a microgravity experiment in a 6 mm tube plotted on a Dukler map Figure from Celata et al. [30]

A theoretical study by Eastman et al. [5] reported that ow pattern transitions in different gravity elds could be reasonably well predicted using the criteria of Weisman et al. [6]: G = B (1/x) (x/1-x)0.2232 g0.4107 for the transition from intermittent to annular G = C/(1-x) g0.25 for the transition from annular to dispersed ow.

Karri and Mathur [7] extrapolated the 1-g models of Taitel et al. and Weismen et al. for horizontal and vertical ows to microgravity. The modied models predicted that the transition boundaries would move towards lower supercial velocities of both the phases as the gravity level was reduced. At a gravity level of 10-5g, the modied ow regime maps showed that there existed only three ow regimes: bubbly, intermittent and annular. An extensive theoretical and experimental study of two phase frictional pressure drop in tubes has been conducted and a number of correlations have been proposed for normal gravity. The methods developed so far can be classied into two categories: homogenous ow and separated ow. The former treats two phase ow as a pseudo single phase ow with suitably averaged properties of the liquid and vapor phases. The latter considers the two phase ow to be essentially ow in two different pipes, with the velocity of the phase constant in its pipe. Two phase ow multipliers are a characteristic of this approach. Correlations for terrestrial pressure drop data have been extended to microgravity ow, by factoring in the reduction in the gravitational force, and considering the usually neglected surface tension. Under conditions of microgravity, the phase distribution is symmetrical and the slip ratio is very small, though there is agreement about the latter only for bubbly ow. Although this is a good starting point, experimental
4

studies are needed to conrm the validity of extending terrestrial designed correlations to microgravity. For a two phase ow system, under reduced gravity conditions, the reduction of buoyancy has two effects: i. decreases the liquid fraction due to a decrease in the slip ratio (leading to a higher liquid velocity, and consequently a higher pressure loss); ii. reduce the turbulence amplication induced by bubble movement. The balance between the two effects will determine the frictional pressure gradient [8]. When the ow is dominated, to a large extent, by inertia, no signicant changes in pressure gradient due to gravity or the lack of it were observed.
EXPERIMENTAL INVESTIGATIONS

FLOW REGIME

An early observation and classication of the various ow regimes in a microgravity environment was done by Zhao and Rezkallah [9]. They grouped the ow regimes into three major regions: a surface tension dominated region, an inertia dominated region and a transitional region in between. The bubbly and slug ows would be the surface tension dominated regimes, the transitional region would comprise of the frothy slug annular ow and the inertia dominated region was the annular ow. They distinguished the bubbly and slug ows using the void fraction - based on the supercial velocities; when the void fraction was smaller than 0.18, bubbly ow was said to exist, when larger, the ow pattern would be a slug ow. Since, in this ow condition, the surface tension force is dominant, the "Taylor" slugs observed had a spherical nose and so did the smaller bubbles between the slugs. Close visual observation reveals that the distance between the smaller bubbles and the slugs remains constant throughout the ow regime. This indicates that there is negligible slip, if at all existent, in the bubbly/ slug ow regimes. As the gas ow rate is gradually increased, the increasing density of the slugs breaks them up into smaller bubbles, which are entrained in the ow. At this point, there is a balance between the surface tension and the inertia forces and it gives rise to the frothy slug annular ow regime. In this regime, the liquid ows at the wall of the tube, while the gas ows in the center with frequent appearances of frothy slugs in it. As the gas ow rate is further increased, inertia dominates and annular ow - liquid at the wall and uninterrupted gas ow in the center - ensues. An investigation by Macgillivray et al. [10] on the variation of annular ow lm thickness with gravity - with a water-air mixture, for vertical upward ow - concluded that the effect of gravity on the average lm thickness values is minimal.

Flow regimes in microgravity - Figure from Valota et al. [17]

The Weber number is dened as the ratio of inertial forces to the surface tension. Given the absence of gravitational effects i.e. buoyancy, the Weber number is one parameter which can describe the criteria for microgravity ow transitions. Rezkallah [11] plotted the Weber numbers in terms of the liquid and vapor velocities for the waterair data of Elkow & Rezkallah and the water-air and glycerine/water-air data sets of Bousman. On the basis of this plot, Rezkallah specied the gas phase Weber number for the transition from bubbly/slug to transitional ow as being around 2, and that for the transition to fully developed annular ow to occur at a gas phase Weber number of 20. Both the liquid viscosity and the diameters of the tube were found to have a minimal inuence on the transitions.


Flow pattern map on the basis of actual liquid and vapor Weber numbers Figure from Rezkallah [11]

However, Parang and Chao observed that the Weber number could not predict the transition data accurately at higher liquid phase velocities [12]. They also contended that, among all the studies that had been performed, the range of Weber numbers had been varied only by varying the velocity of the ow but not the uid properties. It is also highly unlikely that the same force balance between inertial and surface tension forces would hold good in ow patterns as varied as bubbly, slug and annular ows. Based on the data taken from Bousman [13], they suggested that, for the transition from slug to annular ow, the correct scaling parameter would be Weg/Rel and that the transition would occur at 10-3. For the transition from bubbly to slug ow, the parameter was Weg/ Rel1.8 and the transition occurred at 5x10-8. They concluded that their scaling parameters would capture the balance between turbulent shear and gas inertia. Jayawardena et al., using dimensional analysis, theorized that the transitions should depend on the liquid and gas phase supercial velocity Reynolds numbers and the liquid Suratman number [14]. Experimental data, along with some empirical parameters, were used to derive the following boundaries for transition from slug to annular ow. ReSG/ReSL = K2Su- for Su < 106 ReSG = K3Su2 for Su > 106 where the liquid Suratman number SuL = Re2SL/WeSL K2(= 4641.6) and K3(= 2 x 10-9) are empirical parameters.


Microgravity ow pattern maps for Su<106 and Su>106 Figure from Sen [16]

Zhao and Hu extended the drift ux model of Reinarts and proposed to use the supercial Weber numbers of the phases, in dimensionless form, along with experimentally determined values for the empirical parameters C0 and K [15]. This model predicts the boundaries quite well in the bubbly ow regime but is not substantiated with data in the slug ow regime.

A theoretical investigation by Sen on slug to annular ow transition reinforces Jayawardane's results and attempts to provide a physical basis for it [16]. For the transition from slug ow to annular ow, a higher gas ow rate, resulting in a higher ReSG, is required. It is accepted that as the gas phase Reynolds number increases, the wave height decreases and vice versa. Expressing the same in non-dimensional terms, the wave height increases as the ratio of ReSG/ReSL decreases. As the Suratman number increases, for a constant tube diameter, the surface tension must increase, which should result in a decrease in the wave height or a smoother gas liquid interface. The model proposed by Jayawardena et al. was found to predict the transition boundaries quite accurately. Valota et al. conducted experiments using a capacitance sensor for void fraction measurements to determine if the statistical parameters obtained from the uctuation of the void fraction measurements could be used to identify the ow regimes [17]. They reported that the variance and the signal to noise ratio (smaller for annular ow, larger values for slug ow) suited that function well. The semi-empirical void fraction models proposed by Dukler, Reinarts, Bousman and Zhao suffer from their dependence on the empirical values of interfacial friction factor (C0) and drift ux distribution coefcient (K) for slug ow.
BUBBLE DYNAMICS

Cooper et al., after investigation of bubble ebullition on a vertical at plate in a laminar up-ow of supersaturated hexane, concluded that bubble growth was dependent primarily on the thermal diffusivity and the Jacob number, and that the inuence of surface tension and viscosity were negligible [18]. Wang et al. reported that bubbles would slide and roll on the heater surface when the heat ux supplied was beyond a threshold value [19]. The average bubble velocities after departure from their nucleation sites ranged between 80-90% of the free-stream velocity. Ma and Chung experimentally investigated bubble dynamics in FC-72 in a drop tower [20]. All system parameters and results were normalized. They reported that, in microgravity, the bubble's shape developed from oblong to nearly perfect spherical, with the size being larger than that observed in earth gravity. They concluded that gravity effects could be neglected when forced convection - inertia effects - were dominant. As the Reynolds number of the ow increased, they observed that the size of the bubble decreased, the angle of its inclination with the heater surface increased and so did the rate of bubble departure from their nucleation sites. They also reported that, at very high Reynolds numbers, suppression of boiling occurred. They observed that the diameter of a bubble during its growth was proportional to the one-third power of the time of its growth and proposed a correlation for bubble growth in microgravity: D/La = cb Re-1/3t*1/3 where cb is a function of liquid-heater properties and bulk temperature.

At low to moderate Reynolds numbers, the frequency of bubble generation was observed to increase with the ow rate, but was still lower than that in conditions of terrestrial gravity. At higher Reynolds numbers, the inuence of gravity on bubble generation is negligible.


Bubble growth with increase of Reynolds number V* and t* are the normalized volume and time, at departure.

The results of Ma and Chung were conrmed in an experimental investigation by Serret et al. for convective boiling between two plates separated by a distance of 1 mm. They also reported that there was a slight decrease in the average nucleation site temperature, which was consistent with the frequent rewetting of the surface due to higher frequency of bubble departure [21]. Clarke and Rezkallah attempted to numerically simulate the behaviour of bubbles in a two phase ow at different Reynolds numbers of ow and validate them against the experimental data [22]. They concluded that for increasing values of the bubble size and the liquid Reynolds number and for lower values of the surface tension, the drift velocity of the bubble increased. This explains the tendency of the bubbles at the center to be spherical in shape than those at the walls.
HEAT TRANSFER

Misawa and Anghaie performed drop tower experiments on Freon 113 and observed that the slip ratio in conditions of microgravity is less than one [23]. The increased void fraction and thereby acceleration resulted in a larger pressure drop than that predicted by the homogenous model. Investigations conducted aboard a KC-135 aircraft by Kawaji et al. showed that there is a reduction of heat transfer rate in microgravity due to the formation of a thicker vapor layer on the surface of the tube [24]. Flow pattern observation of subcooled Freon 113 showed that there were marked
9

differences in the shape of the bubbles in the dispersed ow regions. Reinarts et al. working with R12, reported that the condensation heat transfer coefcients at microgravity were 26% lower than those at 1-g conditions [25]. Saito et al. conducted investigations in parabolic aircraft with a rod shaped heater concentric with a square channel [26]. It was observed that, under conditions of microgravity, the uid temperature above the heater rod decreases, while the uid temperature below the heater rod remained relatively unchanged. The wall temperature of the heater rod increased slightly, both on the top as well as on the bottom surfaces. Unlike in conditions of gravity, the sub-coolings of the upper stream and lower stream remained fairly constant. They also reported that the local heat transfer coefcients increased slightly on the top of the heater rod, while those at the bottom of the heater rod decreased slightly. The differences in the local heat transfer coefcients were still observed to be small, despite the large variation in the ow regimes under earth gravity and microgravity. All of these can be attributed to the diminished inuence of natural convection at low inlet velocities. They also reported that under conditions of low inlet velocity, the bubbles did not leave the surface of the heater but started growing larger due either to continued heating and vaporization at the surface of the heater or due to coalescence. As the inlet uid velocities were increased, thus increasing the uid inertia, the bubbles were dragged along the surface of the heater. Ohta [27] conducted an extensive series of experiments, in parabolic aircraft, using three different inlet conditions and heat uxes. Under conditions of (a/g)=1 and sub-cooled liquid, the upstream ow was bubbly, but, with increase in quality, it alternated between froth and annular ow. For (a/g)>1, the diameter of the bubbles was observed to decrease, while at the same time, their velocity relative to the velocity of the bulk liquid increased. The instantaneous void fraction decreased. In microgravity, as may be expected, the trend was reversed, with a decrease in bubble velocity, leading to an increase in the void fraction, resulting ultimately in a transition to annular ow at a lower dryness fraction. For conditions where the inlet quality itself was high, annular ow - independent of gravity - resulted across the length of the tube. Under conditions of moderate quality, leading to annular ow, (i) at low heat ux, nucleate boiling is completely suppressed and the heat transfer coefcient increases with gravity and vice versa; (ii) at high heat ux, the heat transfer is dominated by nucleate boiling and is completely independent of gravity. Ohta also attempts to clarify the heat transfer coefcient for the annular ow regime using a method analogous to single phase ow at a heated wall, with certain idealizations. According to him, for an annular ow regime, the heat transfer coefcient would be dened as = q0/(t0 - tsat)

where q0 is the heat ux at the tube wall, tsat is the saturation temperature and t0 is the wall temperature.

10

Moderate quality, annular ow regime, low heat ux Figure from Professor H. Ohta

Rite and Rezkallah, experimenting with a vertically oriented water-air test section, observed that values of the Nusselt number were lower for microgravity than for earth gravity conditions, when the quality was low. As the gas ow rate was increased, increasing the inertia and transitioning to the annular ow regime, the earth gravity and microgravity data points almost coincide, implying that the differences between heat transfer for different gravity levels may be ow regime dependent [28]. They also reported that, for microgravity conditions, the local heat transfer coefcients at the inlet were higher than those at the outlet, unlike the behaviour in earth gravity. In microgravity, the lack of buoyancy impedes the turbulence generating ability of the vapor bubbles, which then ow mainly in the center of the tube and do not affect the laminar sub-layer. They conclude that, for two phase Reynolds numbers greater than 10000, and for moderate quality (0.002), the microgravity heat transfer was greater than earth gravity heat transfer. This might be explained by the greater frequency of bubble detachment from the nucleation sites, leading to more turbulence in the ow, at high Reynolds numbers. Heat transfer experiments, in tubular test sections, using subcooled R113, were carried out by Lui et al [29]. They reported heat transfer coefcients that were upto 20% higher in microgravity than in earth gravity. Investigations by Celata et al. conrmed the above results [30]. They also found that the Dukler ow regime map could best predict the microgravity data.
11

Separate investigations by Celata et al. revealed that gravity did not have any inuence on the heat transfer, for annular ow. They concluded that, for a lower mass ux of uid, gravity would have a larger effect on the heat transfer coefcients [31].
CRITICAL HEAT FLUX

Ma and Chung conducted forced convection boiling experiments with FC-72 and concluded that the CHF in microgravity is lesser than that in normal gravity at low ow rates [32]. But, as the ow rates were increased, the curves for both microgravity and gravity began to coincide, implying that gravity effects diminish with higher ow rates. Critical heat ux is a very transient phenomena, which under the most benign conditions can lead to burnout of equipment. The power budget available in a space environment is very limited, which places severe constraints on the pumping power and thereby the ow rates. While sub-cooling has been found to helpful in earth-gravity conditions, the dimensional constraint limiting the size of condensers in a space environment makes it unfeasible. The combination of these two factors carries the risk of producing small values of CHF in reduced gravity [33]. Extensive experiments - leading to a publication rate of atleast one every two years - have been carried out by Mudawar and Hassan on the inuence of various parameters on the CHF in various orientations and thereby various gravity conditions. They have proposed a theory which provides an intuitive understanding of the CHF mechanism, with a predictive error of about 30%. There are four major models that attempt to explain the phenomenon of ow boiling CHF: boundary layer separation, bubble crowding, micro-layer dryout and interfacial lift-off. The interfacial lift-off model proposed by Galloway and Mudawar [34] is based on a study Hino and Ueda whose study of bubble residence time near the wall revealed that discrete bubbles were replaced by a continuous wavy vapor layer at heat uxes even less than the CHF. This model postulates that the liquid layer makes contact with the walls at wetting fronts which correspond to the troughs and heat transfer due to boiling occurs at these locations. When the pressure force resulting from the interfacial curvature is exceeded by the vapor momentum due to the vigorous boiling at the localized wetting fronts, the wavy layer lifts off leading to an insulating vapor blanket. When this phenomenon occurs at one wetting front, the heat that was being transferred there will have to be transferred from some other neighboring wetting front, leading to more vigorous boiling and hence quicker evaporation there. This chain reaction results in an accelerated formation of the insulating vapor layer, resulting in burnout. Mudawar and Hasan derived an expression for the lift-off heat ux and found it to be proportional to the 1/2 which may be considered a wave curvature parameter, being the amplitude and the wavelength of the idealized sinusoidal wave. This model also predicts that there exists an orientation range, over which, for low velocities, the interface is always stable, thereby producing a continuous vapor lm [35]. Overall, this model was found to be valid for all orientations at near saturated ow at high velocities.
12

CHF trigger mechanism according to the Interfacial LiftOff model Figures from Zhang et al. [36]

An experimental investigation of the effects of the three major forces - inertia, body force and the surface tension force - on the ow boiling CHF revealed that for conditions of saturated ow and velocities below 0.2 m/s, the CHF is very sensitive to the orientation of ow. The CHF was found to be much less sensitive to orientation at velocities higher than 0.5 m/s, in concurrence with other studies. Mudawar and Hassan also compared their data with those predicted by existing correlations and found that the Mudawar and Maddox correlation predicted the data well for subcooled ow, whereas for saturated ow, the Sturgis and Mudawar model was to be preferred. An investigation of the trigger mechanism of ow boiling at CHF performed by Zhang et al. shows that the Interfacial LiftOff model accurately describes the CHF mechanism in both lunar gravity as well as microgravity [36]. They performed their experiment controlling the heat ux as a percentage of the CHF, which allowed them to better visualize the vapor layer formation. In addition to conrming the results of earlier studies, they also concluded that, at ow velocities of about 1.5 m/s, inertia effects exerted dominance and the effect of gravity or the lack of it could be entirely neglected, which means that, if the ow rate could be maintained above 1.5 m/s, equipment designed for use in earth gravity could safely be used in microgravity as well. The extension of any model to sub-cooled ow is complicated by the variation of the bulk liquid enthalpy along the length of the heated surface and the partitioning of the wall energy between the latent and the sensible heats. Zhang et al. modied the Interfacial Liftoff model for sub-cooled ow, introducing a heat utility ratio - dened as
13

the fraction of the surface heat ux that converts liquid at the vicinity of the wall and at the bulk liquid temperature, to vapor. A constant value of 0.5 was recommended for the interfacial friction coefcient. The resulting expression for the value of the CHF for subcooled ow boiling was found to agree with the experimental results within an error range of +-25% [37].
PRESSURE DROP

A pioneering investigation on pressure drop of gas liquid ow in microgravity was conducted by Heppner et al. For a water-air uid ow in a horizontal pipe, they reported that the pressure drop in microgravity was higher than that at terrestrial gravity. Chen et al. reported - for the same dryness fraction, of saturated R114 - that the pressure drop was greater - by about 40% - in microgravity than in earth gravity conditions [38]. An experimental investigation by Colin et al. revealed that for conditions of microgravity, pressure drop were strongly affected by the tube diameter [39]. They reported that, for bubbly turbulent ow, in tubes of diameters 19 and 40 mm, the semiempirical friction factor fp = 0.079 Re-1/4 correctly predicted the friction factor. However, for slug ow, the above equation under-predicts the friction factor for the 19 mm tube, while correctly predicting fp for the 40 mm tube. They also observed that, for tubes of smaller diameters, the friction factor was nearly proportional to the inverse of the Reynolds number, although 8 to 10 times greater. They also reported that the difference between the experimental friction factor and that predicted using the Poiseuille relationship increased when the Reynolds number was decreased. Using an estimate of the void fraction to determine the frictional pressure drop, Zhao and Rezkallah, testing an air-water mixture, concluded that the pressure drop in microgravity conditions was slightly more than that at earth-gravity conditions, the differences between the two ranging from 1 to 14%. Plots of the two phase multiplier values versus the quality for both microgravity and terrestrial gravity conditions were found to coincide, for the same gas quality, supporting their observations. A possible reason for the great difference between the results of this investigation and that of Chen et al. and Colin et al. might be that the ow velocities were much higher than that in the earlier ones. The average Reynolds number of the mixture was reported to be around 104, putting it rmly in the turbulent ow regime. Another major reason for the difference is that the earlier investigations were conducted for a horizontal pipe, where stratication of ow would occur in terrestrial gravity conditions, whilst the latter was for vertical upward ow. Investigations by Ohta revealed that the pressure drop increases with the supercial velocities in microgravity. According to Ohta, for an annular ow with a vapor core, the interfacial friction factor may be dened as a function of the gravity, mass ow and the quality as follows: fi/fi0 = 1 + 0.08(1-x/x)0.9 Fr-1 : 0< a/g 2.

14

In the above equation, for fully vapor ow, the interfacial friction factors are equal (fi1g = fi2g = fi0), as might be expected. Using the Chisholm and Laird correlation for earth gravity conditions, Ohta extrapolates to nd the interfacial friction factor for conditions of micro and hyper gravity. The friction factor was observed to be proportional to gravity and the effect of gravity on the friction factor was reduced with an increase in the quality and thereby the mass velocity of vapor ow. Zhao et al. conducted experiments aboard the Mir space station in which they compared the pressure drops of two phase ow with some common correlations [40]. They concluded that the Friedel model provided relatively better agreement with the experimental results. A comparison by Fang et al. of the predictive capability of the various correlations to microgravity data revealed that the McAdams et al., Chisholm and Muller-Steinhagen and Heck performed best [41]. They also reported that the Zhao and Rekallah vertical ow data was reasonably well predicted by most of the correlations, which were dened for horizontal ow, implying that the effect of orientation on the two phase frictional pressure drop was minimal. They concluded that the practice of extending terrestrial gravity correlations to predict microgravity behaviour would be reasonable up to the rst approximation, but, the fact that even the best performing correlations end up underpredicting the experimental data implies that there needs a specic correlation for microgravity data. Fang et al., in a separate study [42], reported the development of a new correlation specically for microgravity two phase ow which could predict the experimental data better than all existing correlations: lo2 = Y2x0.87 + (1-x0.626)0.54[1 + 2x1.823(Y2 - 1) + 47.74x1.4] where lo2 is dened according to convention.
SUMMARY

In comparison with the single phase ow applications being presently utilized, two phase ows have many advantages which make them desirable for use in both microgravity as well as hyper-gravity environments. However, the current level of knowledge is not sufcient to reliably design a two phase loop system. There is consensus that there are three major ow regimes in microgravity: bubbly, slug and annular. However, there is considerable confusion regarding the boundaries between the different ow regimes. Various attempts - extrapolating 1-g ow regime maps to microgravity, classication using semi-empirical void fraction relations, theoretical force balance analyses - at dening the ow regime boundaries have yielded different transition criteria. Among those available, Duklers ow regime map is preferable. The dimensionless analysis using the liquid Suratman number - while giving
15

no reasons for the dramatic change in the behaviour of the two phase mixture at Su = 106, ts most of the data with reasonable accuracy. There has been some understanding gained about bubble dynamics in a microgravity environment. The shape of the bubble is nearly spherical, since the distortion effect due to buoyancy is eliminated. With increasing inlet velocities, the average bubble size decreases. This may be attributed to greater heat loss due to convection to the ambient ow. If the Reynolds number of the ow is very high, the bubble size and shape resemble that observed in normal gravity conditions, implying that gravity effects are no longer inuential when inertia dominates. The bubble does not depart from its nucleation site, unless it is forced by the inertial effects of the ow. The lack of buoyancy - which on earth would provide a lift force to the bubble - makes the bubble roll on the surface of the heater, while it grows in size due to heating from the heater or coalescence with other bubbles, until the velocity and shear gradients in the ow distort its spherical shape and lift it into the center of the ow. Once the bubble reaches the center of the ow, i.e., the shear force on the bubble surface is uniform, surface tension forces the bubble to the shape of a sphere again. The heat ux at the heater surface and the inlet ow velocity inuence bubble growth and frequency of bubble nucleation and departure. A greater understanding of the underlying mechanism would be helpful in clearing up the confusion regarding microgravity heat transfer. The study of microgravity heat transfer has been constrained by the relatively small amount of experimental data available and the lack of standardized test apparatus. On the basis of available data, the heat transfer has been found to be strongly dependent on the ow regime, with the heat transfer for annular ows being practically independent of gravity. This may be explained by the dependence of frequency of bubble departure and nucleation site density on the mass ow rates and heat uxes applied. The importance of having a thorough knowledge of the critical heat ux cannot be over-emphasized, since it is the surest way of destroying equipment. The interfacial liftoff model proposed by Galloway and Mudawar and extended by Zhang et al. gives a reasonably accurate prediction of the experimental data. The model has been found to work equally well for sub-cooled ows, with minor modications. The pressure drop, in a microgravity environment, has been found to be strongly dependent on the tube diameter and the void fraction. Although a number of empirical and semi-empirical correlations have been proposed to determine the pressure drop, the difculty of obtaining void fraction data has resulted in an incomplete understanding of this topic. A recent comparison of the existing correlations has revealed that the McAdams and the Muller-Steinhagen and Heck correlations performed best. The pressure drop was also found to be independent of the orientation. Further studies, with standardized test methods, is required to clear up the many hazy points in our present understanding of microgravity heat transfer, before any practical applications can be realized. One way of getting around this scientic gestation
16

period is to ramp up the allocation of energy to pumping appliances, since the effects of gravity, be it micro or hyper, greatly diminish when inertia is dominant.
ACKNOWLEDGEMENT

I would like to gratefully acknowledge the assistance rendered by Prof. Haruhiko Ohta in sending me several gures from his research.
REFERENCES

1. A.E.Dukler,J.A.Fabre,J.B.Mcquillen,R.Vernon,1988,Gasliquid?lowatmicrogravity
breakupofbubblesanddrops.SINTEFReportSTF24A02531

2. MarcDhainaut,2002,Literaturestudyonobservationsandexperimentsoncoalescenceand 3. A.M.Kamp,A.K.Chesters,C.Colin,J.Fabre,2001,Bubblecoalescenceinturbulent?lows:A
mechanisticmodelforturbulenceinducedcoalescenceappliedtomicrogravitybubbly?low. Int.J.MultiphaseFlowVol.27,13631396. 4. YehudaTaitel,LarryWhite,1996,Theroleofsurfacetensioninmicrogravityslug?low. Chem.Engg.Sci.Vol.51,695700. 5. EastmanR.E.,FeldmanisC.J.,HaskinW.L.,WeaverK.L.,1984,Twophasethermaltransport forspacecraft.TechnicalReportNo.AFWALTR843028. 6. J.Weisman,S.Y.Kang,1981,Flowpatterntransitionsinverticalandupwardlyinclinedlines. Int.J.MultiphaseFlowVol.7,271291. 7. S.B.R.Karri,V.K.Mathur,1988,Twophase?lowpatternmappredictionsundermicrogravity. AIChEVol.34,137139. 8. L.Zhao,K.S.Rezkallah,1995,Pressuredropingasliquid?lowundermicrogravityconditons. Int.J.MultiphaseFlow.Vol.21,837849. 9. L.Zhao,K.S.Rezkallah,1993,Gasliquid?lowpatternsatmicrogravityconditions.Int.J. MultiphaseFlow.Vol.19,751763. 10. R.M.MacGillivray,K.S.Gabriel,2003,Astudyofannular?low?ilmcharacteristicsin microgravityandhypergravityconditions.ActaAstronautica.Vol.53,289297. 11. K.S.Rezkallah,1996,Webernumberbased?lowpatternmapsforliquidgas?lowsat microgravity.Int.J.MultiphaseFlow.Vol.22,12651270. 12. M.Parang,D.Chao,1998,Twophase?lowmappingandtransitionundermicrogravity conditions,36thAIAAAerospaceSciencesMeetingandExhibit. 13. W.S.Bousman,1995,Studiesoftwophasegasliquid?lowinmicrogravity.NASAReportNo. 195434(N9524906). 14. S.S.Jayawardena,V.Balakotaiah,L.C.Witte,1997,Flowpatterntransitionmapsfor microgravitytwophase?low.AIChEVol.43,16371640. 15. J.F.Zhao,W.R.Hu,2000,Slugtoannular?lowtransitionofmicrogravitytwophase?low.Int. J.MultiphaseFlow.Vol.26,12951304. 16. NilavaSen,2010,Twophaseslugtoannular?lowpatterntransitioninmicrogravity.Acta Astronautica.Vol.66,13731377. 17. L.Valota,C.Kurwitz,A.Shephard,F.Best,2007,Microgravity?lowregimeandanalysis.Int.J. MultiphaseFlow.Vol.33,11721185. 18. M.G.Cooper,K.Mori,C.R.Stone,1983,Behaviourofvaporbubblesgrowingatawallwith forced?low.Int.J.HeatandMassTransfer.Vol.26,14891503. 19. T.C.Wang,T.J.Snyder,J.N.Chung,1996,Experimentalexaminationofforcedconvection subcoolednucleateboilinganditsapplicationinmicrogravity.J.ofHeatTransfer.Vol.118, 237241.

conditions:?lowpatternsandtheirtransitions.Int.J.MultiphaseFlowVol.14,389400.

17

20. YueMa,J.N.Chung,2001,Astudyofbubbledynamicsinreducedgravityforcedconvection 21. D.Serret,D.Brutin,O.Rahli,L.Tadrist,2010,Convectiveboilingbetween2Dplates:


boiling.Int.J.ofHeatandMassTransfer.Vol.44,399415. Microgravityin?luenceonbubblegrowthanddetachment.MicrogravitySci.Technol.Vol.22, 377385. 22. N.N.Clarke,K.S.Rezkallah,2001,Astudyofdriftvelocityinbubblytwophase?lowunder microgravityconditions.Int.J.ofMultiphaseFlow.Vol.27,15331554. 23. M.Misawa,S.Anghaie,1991,Zerogravityvoidfractionandpressuredropinaboiling channel.AIChESymp.Ser.87,226235. 24. M.Kawaji,C.J.Westbye,B.N.Antar,1991,Microgravityexperimentsontwophase?lowand heattransferduringquenchingofatubeand?illingofavessel.AIChESymp.Ser.87, 236243. 25. T.R.Reinarts,F.R.Best,W.S.Hill,1992,De?initionofcondensationtwophase?lowbehaviours forspacecraftdesign.AIPConferenceProc.No.246,Vol.1,12161225. 26. M.Saito,N.Yamaoka,K.Miyazaki,M.Kinoshita,Y.Abe,1994,Boilingtwophase?lowunder microgravity.NuclearEnggandDesign,Vol.146,451461. 27. H.Ohta,2003,Microgravityheattransferin?lowboiling,AdvancesinHeatTransfer,Vol.37, 176. 28. R.W.Rite,K.S.Rezkallah,1997,Localandmeanheattransfercoef?icientsinbubblyandslug ?lowsundermicrogravityconditions.Int.J.ofMultiphaseFlow.Vol.23,3754. 29. R.K.Lui,M.Kawaji,T.Ogushi,1994,Anexperimentalinvestigationofsubcooled?lowboiling heattransferundermicrogravityconditions.10thInternationalHeatTransferConference, BrightonVol.7,497502. 30. G.P.Celata,G.Zummo,2008,Flowboilinginmicrogravity.EBECEM20081Encontro BrasileirosobreEbulio,CondensaoeEscoamentoMultifsicoLquidoGs. 31. C.Baltis,G.P.Celata,M.Cumo,L.Saraceno,G.Zummo,2012,Gravityin?luenceonheat transferratein?lowboiling.MicrogravitySci.Technol.Vol.24,203213. 32. YueMa,J.N.Chung,2001,Anexperimentalstudyofcriticalheat?luxinmicrogravityforced convectionboiling.Int.J.ofMultiphaseFlow.Vol.27,17531767. 33. H.Zhang,I.Mudawar,M.M.Hasan,2002,Experimentalassessmentoftheeffectsofbody force,surfacetensionforceandinertiaon?lowboilingCHF.Int.J.ofHeatandMassTransfer. Vol.45,40794095. 34. J.E.Galloway,I.Mudawar,1993,CHFmechanismin?lowboilingfromashortheatedwall part2.TheoreticalCHFmodel.Int.J.ofHeatandMassTransfer.Vol.36,25272540. 35. H.Zhang,I.Mudawar,M.M.Hasan,2002,Experimentalandtheoreticalstudyoforientation effectson?lowboilingCHF.Int.J.ofHeatandMassTransfer.Vol.45,44634477. 36. H.Zhang,I.Mudawar,M.M.Hasan,2005,FlowboilingCHFinmicrogravity.Int.J.ofHeatand MassTransfer.Vol.48,31073118. 37. H.Zhang,I.Mudawar,M.M.Hasan,2007,CHFmodelforsubcooled?lowboilinginEarth gravityandmicrogravity.Int.J.ofHeatandMassTransfer.Vol.50,40394051. 38. I.Chen,R.Downing,E.Keshock,M.Alsharif,1991,Measurementsandcorrelationoftwo phasepressuredropundermicrogravityconditions.J.Thermophys.Vol.5,514523. 39. C.Colin,J.Fabre,1995,Gasliquidpipe?lowundermicrogravityconditions:in?luenceoftube diameteron?lowpatternsandpressuredrops.Adv.SpaceRes.Vol.16,137142. 40. J.F.Zhao,J.C.Xie,H.Lin,W.R.Hu,A.V.Ivanov,A.Yu.Belyaev,2001,Experimentalstudieson twophase?lowpatternsaboardtheMirspacestation.Int.J.ofMultiphaseFlow.Vol.27, 19311944. 41. XiandeFang,HonggangZhang,YuXu,XianghuiSu,2012,Evaluationofusingtwophase frictionaldropcorrelationsfornormalgravitytoreducedgravityandmicrogravity.Adv. SpaceRes.Vol.49,351364. 42. XiandeFang,YuXu,2013,Correlationsfortwophasefrictionpressuredropunder microgravity.Int.J.ofHeatandMassTransfer.Vol.56,594605.

18

You might also like