Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2013 May 21;110(23):9237–9242. doi: 10.1073/pnas.1217546110

Heat transport in bubbling turbulent convection

Rajaram Lakkaraju a,1, Richard J A M Stevens a,b, Paolo Oresta c,d, Roberto Verzicco a,e, Detlef Lohse a, Andrea Prosperetti a,b
PMCID: PMC3677508  PMID: 23696657

Abstract

Boiling is an extremely effective way to promote heat transfer from a hot surface to a liquid due to numerous mechanisms, many of which are not understood in quantitative detail. An important component of the overall process is that the buoyancy of the bubble compounds with that of the liquid to give rise to a much-enhanced natural convection. In this article, we focus specifically on this enhancement and present a numerical study of the resulting two-phase Rayleigh–Bénard convection process in a cylindrical cell with a diameter equal to its height. We make no attempt to model other aspects of the boiling process such as bubble nucleation and detachment. The cell base and top are held at temperatures above and below the boiling point of the liquid, respectively. By keeping this difference constant, we study the effect of the liquid superheat in a Rayleigh number range that, in the absence of boiling, would be between 2 × 106 and 5 × 109. We find a considerable enhancement of the heat transfer and study its dependence on the number of bubbles, the degree of superheat of the hot cell bottom, and the Rayleigh number. The increased buoyancy provided by the bubbles leads to more energetic hot plumes detaching from the cell bottom, and the strength of the circulation in the cell is significantly increased. Our results are in general agreement with recent experiments on boiling Rayleigh–Bénard convection.

Keywords: two-phase convection, latent heat, boundary layers, point bubble model, simulations


The greatly enhanced heat transfer brought about by the boiling process is believed to be due to several interacting components (13). With their growth the bubbles cause a microconvective motion on the heating surface, and as they detach by buoyancy, the volume they vacate tends to be replaced by cooler liquid. Especially in subcooled conditions, the liquid in the relatively stagnant microlayer under the bubbles can evaporate and condense on the cooler bubble top. This process provides for the direct transport of latent heat, which is thus able to bypass the low-velocity liquid region adjacent to the heated surface due to the no-slip condition. The bubble growth process itself requires latent heat and, therefore, also removes heat from the heated surface and the neighboring hot liquid. Finally, with their buoyancy, the bubbles enhance the convective motion in the liquid beyond the level caused by the well-known single-phase Rayleigh–Bénard (RB) convection mechanisms (4, 5). This last process is the aspect on which we focus in the present article.

In classical single-phase RB convection, the dimensionless heat transport, Inline graphic, the Nusselt number, is defined as the ratio of the total heat transported through the cell to the heat that would be transported by pure conduction with a quiescent fluid. This ratio increases well above 1 as the Rayleigh number Inline graphic is increased due to the onset of convective motion in the cell. Here g is the acceleration of gravity, β the isobaric thermal expansion coefficient, Inline graphic the difference between the temperature Inline graphic of the hot bottom plate and the temperature Inline graphic of the cold top plate, L the height of the cell, ν the kinematic viscosity, and κ the thermal diffusivity. Further, Inline graphic depends on the shape of the cell, its aspect ratio (defined for a cylindrical cell of diameter D as Inline graphic), and the Prandtl number Inline graphic of the liquid. For Inline graphic in the range Inline graphic and Inline graphic in the range Inline graphic, the heat transport satisfies an approximate scaling relation Inline graphic (4, 5).

How is this scaling modified if the hot plate temperature Inline graphic is above the fluid saturation temperature Inline graphic, so that phase change can occur? The present article addresses this question by focusing on the enhanced convection caused by the bubble buoyancy, rather than attempting a comprehensive modeling of the actual boiling process in all its complexity. We carry out numerical simulations in the range Inline graphic for a cylindrical cell with aspect ratio Inline graphic for Inline graphic, which is appropriate for water at 100 °C under normal conditions.

This work differs in two major respects from our earlier studies of the problem. In the first place, we are now able to reach a much higher Rayleigh number, Inline graphic as opposed to Inline graphic as in ref 6, and to include three times as many bubbles. Secondly, we now study the effect of the liquid superheat, which was held fixed before.

The extensive literature on boiling leads to the expectation that the appearance of bubbles would cause a substantial increase in Inline graphic with respect to single-phase convection (1). For RB convection, the effect of phase change has recently been studied in ref. 7 for the case of ethane near the critical point, and indeed a major increase of the heat transport has been found.

Model

The present article is based on the same mathematical model and numerical method that we have used in ref. 6. and several other recent papers (8, 9). Briefly, under the Boussinesq approximation, conservation of mass, momentum, and thermal energy equations for the liquid are:

graphic file with name pnas.1217546110eq1.jpg
graphic file with name pnas.1217546110eq2.jpg

and

graphic file with name pnas.1217546110eq3.jpg

Here u, p, and T are the liquid velocity, pressure, and temperature, and ρ and Inline graphic are the liquid density and specific heat, while Inline graphic is the total number of bubbles. The bubbles are modeled as point sources of momentum and heat for the liquid. The i-th bubble offers a mechanical forcing Inline graphic and a thermal forcing Inline graphic. Here Inline graphic is the radius of the i-th bubble, Inline graphic the bubble heat transfer coefficient, and the liquid temperature Inline graphic and acceleration Inline graphic are evaluated at the location Inline graphic of the bubble; Inline graphic is the bubble surface temperature assumed to be at saturation with respect to the static pressure.

The motion of each bubble, envisaged as a sphere, is followed in a Lagrangian way by means of an equation that, in addition to buoyancy, includes drag, added mass, and lift:

graphic file with name pnas.1217546110eq4.jpg

where Inline graphic, Inline graphic are the added mass and lift coefficients, and Inline graphic is the bubble velocity. CD is the drag coefficient (see ref. 6).

In its mechanical aspects, therefore, the model is similar to existing ones, which have been extensively used in the literature to simulate dilute disperse flows with bubbles and particles (10, 11). The novelty of our model lies in the addition of the thermal component. The heat exchange between the bubble and the liquid in its vicinity is modeled by means of a heat transfer coefficient dependent on the Péclet number of the bubble–liquid relative motion and on the Prandtl number of the liquid. The radial motion of the bubbles is slow enough that the vapor pressure remains essentially equal to the ambient pressure, which implies that the bubble surface temperature can be assumed to remain at the saturation value. The bubble volume, on which the enhanced buoyancy effect depends, is calculated by assuming that the entire heat absorbed by a bubble is used to generate vapor at the saturation density and pressure (for complete details, see ref. 6).

The calculation is carried out on a finite-difference grid based on cylindrical coordinates. The standard staggered-grid arrangement is used for the flow variables and the projection method for the calculation of the pressure and time stepping (12). No-slip conditions are applied on the bottom and top of the cell, and also on the lateral boundary. The Lagrangian treatment of the bubbles proceeds by means of a third-order Runge–Kutta method. The energy and force imparted by each bubble to the liquid are interpolated to the grid points of the cell containing the bubble in such a way as to preserve the total energy and the resultant and moment of the force.

Simulations are carried out on computational grids with the angular, radial, and axial directions discretized by means of Inline graphic, Inline graphic, Inline graphic, and Inline graphic nodes for Inline graphic, Inline graphic, Inline graphic, and Inline graphic. The simulations are therefore well resolved according to the requirements specified in refs. 13 and 14. We have also checked the global balances of appendix B in ref. 6, finding that they were satisfied to within Inline graphic.

When a bubble reaches the top cold plate, it is removed from the calculation to model condensation and a new bubble is introduced at a random position on the bottom hot plate so that the total number of bubbles in the calculation remains constant. We do not attempt to model the nucleation process, which, with the present state of knowledge, cannot be done on the basis of first principles and which would require addressing extremely complex multiscale issues. For our limited purpose of studying the bubble-induced increased buoyancy, it is sufficient to simply generate a new bubble at the hot plate. We do not model the process by which the bubble detaches from the plate but assume that it is free to rise immediately as it is introduced. The initial bubble radius is arbitrarily set at 38 μm. As shown in ref. 9, the initial bubble size is immaterial provided it is in the range of a few tens of microns. In view of their smallness, the latent heat necessary for their generation is very small and is neglected. We show results for three values of the total number of bubbles Inline graphic, namely Inline graphic, 50,000, and 150,000. Another parameter we vary is the degree of superheat, Inline graphic, which we express in the dimensionless form Inline graphic.

The Nusselt number shown in the following is defined as Inline graphic, where k is the liquid thermal conductivity and Inline graphic is the heat flux into the bottom plate. This quantity differs from Inline graphic, the heat flux at the upper plate, due to the heat stored in the bubbles. [The Nusselt number shown in our previous papers (6, 8, 9) are based on the average between Inline graphic and Inline graphic.] An important parameter introduced by the bubbles is the Jakob number Inline graphic, where ρ and Inline graphic are the densities of liquid and vapor and Inline graphic is the latent heat for vaporization. Physically, Inline graphic expresses the balance between the available thermal energy and the energy required for vaporization. With Inline graphic oC, Inline graphic varies between 0 and 1.68 as ξ varies between 0 and 1/2. For Inline graphic, the bubbles introduced at the hot plate can only encounter liquid at saturation temperature or colder, and therefore they cannot grow but will mostly collapse. On the other hand, for Inline graphic, they have significant potential for growth.

To give an impression of the physical situation corresponding to our parameter choices, we may mention that 100 °C water in a 15-cm-high cylinder with an imposed temperature difference Inline graphic °C would correspond to Inline graphic. The Kolmogorov length scale based on the volume- and time-averaged kinetic energy dissipation in single-phase RB convection is 3 mm for Inline graphic and 0.5 mm for Inline graphic (5) and is therefore always much larger than the initial size of the bubbles (e.g., 13 times larger for the highest Rayleigh number). In our simulations bubbles grow at most to a diameter of 130 dinj (see Fig. 5). The bubble volume fractions are less than 0.01% and hence use of the point bubble model is justified.

Fig. 5.

Fig. 5.

Bubble size (given in bubble diameters Inline graphic) distributions at various heights in the cylinder: (A–E) bubble density versus bubble diameter, and (F–J) PDF versus bubble diameter for Inline graphic and Inline graphic. From top to bottom, Inline graphic (A+F), 0.2 (B+G), 0.3 (C+H), 0.4 (D+I), and 0.5 (E+J). The bubble injection diameter at the hot plate is Inline graphic microns. Data symbols for different vertical heights Inline graphic (red squares), 0.25 (blue circles), 0.5 (green diamonds), 0.75 (brown triangles), and 0.95 (black stars). Note that all of the vertical axes are logarithmic except D and E.

Observations on Heat Transport and Flow Organization

In Fig. 1, the dependence of Inline graphic on the Rayleigh number Inline graphic and the dimensionless superheat ξ is shown for Inline graphic bubbles. Here Inline graphic is normalized by Inline graphic, the single-phase Nusselt number corresponding to the same value of Inline graphic. Each symbol shows the result of a separate simulation carried out for the corresponding values of Inline graphic and ξ. A colored surface is interpolated through the computed results with the color red corresponding to Inline graphic and the color blue to Inline graphic.

Fig. 1.

Fig. 1.

Inline graphic for boiling convection normalized by the corresponding single-phase value Inline graphic for Inline graphic bubbles. Here ξ is the normalized superheat, Inline graphic. The symbols correspond to Inline graphic (square), Inline graphic (triangles), Inline graphic (circles), and Inline graphic (stars).

The same data are shown on a 2D plot of Inline graphic versus ξ in Fig. 2A for four different Rayleigh numbers in descending order; here the dashed lines are drawn as guides to the eye. It is evident that the relative enhancement of the heat transport is a decreasing function Inline graphic. This statement, however, does not apply to the absolute heat transport shown in Fig. 2B, where Inline graphic is not normalized by the single-phase value. Here Inline graphic increases in ascending order, which shows that the bubbles always have a beneficial effect on the heat transport. For very small superheat, the heat transport approaches the single-phase value as shown in the inset of Fig. 2B.

Fig. 2.

Fig. 2.

Inline graphic (A) and Inline graphic (B) as functions of the normalized superheat ξ for 50,000 bubbles. The symbols correspond to Inline graphic (squares), Inline graphic (triangles), Inline graphic (circles), and Inline graphic (stars). The inset shows a detail for small superheat ξ for Inline graphic (squares) and Inline graphic (triangles) with quadratic fits to the data.

Figs. 1 and 2 show results calculated keeping the bubble number fixed. This procedure, therefore, does not faithfully reflect physical reality, as it is well known that the number of bubbles is an increasing function of superheat. The dependence is actually quite strong, with the number of bubbles proportional to Inline graphic raised to a power between 3 and 4 (1). However, varying independently Inline graphic and ξ permits us to investigate separately the effect of these quantities.

The effect of changing the bubble number from 50,000 to 150,000 at the same ξ is shown in Fig. 3 A and B for Inline graphic and Inline graphic, respectively. In the latter case, we also include results for Inline graphic. For small ξ, the heat transfer enhancement is small as the bubbles will mostly encounter colder liquid, condense, and add very little to the system buoyancy. As the superheat ξ increases, however, the effect of the bubbles becomes stronger and stronger, and larger the larger their number.

Fig. 3.

Fig. 3.

Inline graphic versus ξ for three different bubble numbers, Inline graphic (squares, B), 50,000 (triangles), and 150,000 (circles); A is for Inline graphic and B for Inline graphic. The curved dash line is a fit to the experimental data of Zhong et al. (7) shown by the filled symbols. The inset is a blow-up for the range Inline graphic. Error bars are shown inside the hollow symbols.

In Fig. 3B, the solid symbols are the data of ref. 7 taken at a higher Rayleigh number, Inline graphic. (This article reports data for both increasing and decreasing superheat. We show here only the latter data because, for increasing superheat, there is a threshold for fully developed boiling conditions that pushes the onset of bubble appearance beyond Inline graphic. For decreasing ξ, on the other hand, fully developed boiling conditions prevail all the way to small values of ξ.) The inset in the figure shows our computed results and the experimental data for Inline graphic. A major difference between our simulations and the experiment is that, in the latter, the number of bubbles increases with the superheat, while it remains constant with ξ in the simulations. We can nevertheless attempt a comparison as follows. Quadratic interpolation using our results for the three values of Inline graphic suggests that, to match the experimental values, we would need Inline graphic for Inline graphic and Inline graphic for Inline graphic. If, as suggested by experiment, the actual physical process results in a relation of the form Inline graphic, we find Inline graphic, which falls in the experimental range Inline graphic mentioned before. With this value of m, we can estimate the number of bubbles necessary to account for the measured Inline graphic at Inline graphic. Using Inline graphic, we find Inline graphic for Inline graphic and Inline graphic for Inline graphic. These values are in agreement and consistent with the fact that our computed result at Inline graphic is somewhat higher than the measured value for Inline graphic. The picture that emerges from these considerations is therefore in reasonable agreement with the experiment. A similar exercise cannot be carried out for larger values of ξ as in the experiment bubbles then become so large that they coalesce and form slugs with nonnegligible dimensions. Our model, in which the vapor volume fraction is assumed to be so small as to be negligible, clearly cannot be applied to this situation.

The heat transport in single-phase RB convection can be approximated by an effective scaling law Inline graphic. In the present Inline graphic range, the experimental data are well represented with the choices Inline graphic and Inline graphic. How does the effective scaling law change for boiling convection? Fig. 4A shows the Nusselt number versus Rayleigh number for different values of ξ for Inline graphic bubbles. The two solid lines have slopes Inline graphic and Inline graphic, while the dashed line shows the single-phase values. If we fit Inline graphic for the boiling case again with an effective scaling law Inline graphic, we obtain the effective exponents Inline graphic shown in the inset of the figure (as squares). Of course, Inline graphic and, as ξ increases, Inline graphic decreases to a value close to 0.20. In the range Inline graphic, the numerical results for Inline graphic are well represented by Inline graphic, which monotonically increases from 1 to 34.15 for Inline graphic. How strongly does the prefactor Inline graphic and the effective scaling exponent Inline graphic depend on Inline graphic? In the inset of the same figure, we show Inline graphic for Inline graphic bubbles (see circles), to compare with the Inline graphic case. The functional dependence Inline graphic is very close for the two cases. Further, we find Inline graphic for 150,000 bubbles—that is, a stronger ξ dependence compared with the Inline graphic case, reflecting the enhanced number of bubbles.

Fig. 4.

Fig. 4.

(A) Inline graphic versus Inline graphic and (B) Inline graphic versus ξ for 50,000 bubbles. In A, the numerical results are shown as crosses Inline graphic, squares Inline graphic, triangles Inline graphic, circles Inline graphic, diamonds Inline graphic, and stars Inline graphic. Simulations without bubbles are also shown for comparison as a dashed line joining small dots and data from the LB simulations of ref. 15 as filled circles, light gray for no-boiling, and dark gray for boiling. In the inset, the effective scaling exponent Inline graphic obtained from power-law fits of the form Inline graphic is shown as a function of ξ for 50,000 (squares) and 150,000 (circles) bubbles. In B, the effective buoyancy has been computed from Eq. 5. The symbols are the same as in Fig. 2A. Error bars are shown inside the hollow symbols.

It is tempting to regard the increased heat transport as due to the additional buoyancy provided by the bubbles. In this view, the Rayleigh number should be based on an effective buoyancy Inline graphic in place of the pure liquid buoyancy β. An expression for Inline graphic can then be found by equating Inline graphic to Inline graphic with the result:

graphic file with name pnas.1217546110eq5.jpg

The quantity Inline graphic as given by this relation is shown in Fig. 4B as function of ξ and Inline graphic for Inline graphic. For the same ξ, Inline graphic decreases as Inline graphic increases as expected on the basis of Figs. 1 and 2. For fixed Inline graphic, Inline graphic increases with ξ, also as expected. It is quite striking that Inline graphic can exceed β by nearly three orders of magnitude for Inline graphic and small Rayleigh number. Note that one cannot directly compare the numerical values for Inline graphic shown in Fig. 4B with an experiment in which ξ is increased in a given cell, as in our plot Inline graphic is fixed, whereas in the experiment Inline graphic with Inline graphic as discussed above.

A recent Lattice–Boltzmann (LB) simulation of finite-size bubbles also found heat transport enhancement (15). The results of this study for Inline graphic are shown by filled circles in Fig. 4A. The heat transfer enhancements achieved are much smaller than ours, most likely due to the significantly smaller number of bubbles (only a few hundreds), as well as other differences (the values of Inline graphic, Inline graphic, etc.) of lesser importance.

Where are the bubbles in the flow, and how are they distributed in size, depending on their location? In Fig. 5, the statistics on the bubble diameter computed at different vertical heights in the cylinder are shown. To get an immediate impression on how large the bubbles have grown, we have normalized the bubble diameter Inline graphic with the initial injection diameter Inline graphic microns. We calculate the time-averaged bubble density in thin horizontal slices positioned at five different vertical heights in the cylinder for various diameter ranges (Fig. 5 A–E). For small superheat Inline graphic, the bubble nuclei do not grow much: most of them only up to a diameter 12 times the injection size and only very few toward 25 times the injection diameter (Fig. 5A). Moreover, they do not make it up to one-quarter of the cell height, as they encounter cold liquid and condense. As we increase ξ, the bubbles grow to larger sizes and can even reach the top plate (Fig. 5 B–E). Although the number density at a given cross-section decreases, a wide range of bubble size emerges, leading to poly-dispersity. The bubbles can grow up to a size of even 100 times the initial injection diameter. Note that for large Inline graphic and even more at Inline graphic, at any plane away from the boundary layers, the number density shows a similar trend for bubble size distribution, reflecting the homogeneously boiling situation. In the right column of Fig. 5 F–J, we show the corresponding probability density functions (PDFs) versus the bubble diameter, now all in log-linear form. Again we see that both the bubble maximum and the most probable diameter increase as we increase ξ.

We now come to the local flow organization. As well known, the boundary layers formed on the bottom (and top) plate are marginally stable and occasional intermittent eruptions of hot (or cold) liquid occur at their edges. Vapor bubbles subject these boundary layers to intense fluctuations, which enhance the convective effects. As an example, Fig. 6 shows sample time records of the dimensionless vertical velocity Inline graphic (Fig. 6A), and temperature Inline graphic (Fig. 6B) versus normalized time Inline graphic near the axis at Inline graphic—that is, just outside the hot thermal boundary layer. The velocity scale Inline graphic is defined by Inline graphic and Inline graphic. The dashed lines are results for the single-phase case. The immediate observation is that the small-scale fluctuations are much stronger in the two-phase case. As expected, the positive and negative velocity fluctuations are correlated with warm and cold temperature fluctuations, respectively.

Fig. 6.

Fig. 6.

The solid lines show the dimensionless vertical velocity Inline graphic (A) and temperature θ (B) as functions of dimensionless time in the hot liquid at a height Inline graphic near the axis. The dashed lines show similar results for simulations without bubbles; here Inline graphic, Inline graphic, and Inline graphic.

To give an impression of the difference brought about by the presence of bubbles on the convective motions in the cell, we show in Fig. 7 snapshots of the dimensionless temperature in a vertical plane through the axis of the cell for Inline graphic in the single-phase (Fig. 7A) and two-phase (Fig. 7B) cases, the latter for Inline graphic and Inline graphic. We notice that bubbles considerably thicken the layer of hot fluid near the base and make it more energetic compared with the single-phase situation. Chunks of hot liquid can be seen all of the way up near the cold plate, presumably caused by the latent heat deposited by condensing bubbles in the bulk liquid. The up–down symmetry of the single-phase case that can be seen in Fig. 7A is markedly absent in the two-phase case because of the tendentially upward motion of the bubbles, which condense on encountering liquid colder than Inline graphic. This mechanism is evidently quite different from the symmetry-breaking process observed in non-Boussinesq systems, which is due to the temperature dependence of viscosity (16).

Fig. 7.

Fig. 7.

Instantaneous dimensionless temperature field in a vertical plane through the cell axis for convection without bubbles (A) and with bubbles (B). The color varies from red for Inline graphic to blue for Inline graphic; here Inline graphic, Inline graphic, and Inline graphic.

It is found that, for Inline graphic, the time- and area-averaged mean temperature in the cell is very close to 0.5—that is, Inline graphic, except in the two boundary layers near the plates. A detailed view of the temperature distribution in these layers is provided in Fig. 8. The figure makes evident that the temperature distribution in the upper and lower layers is not symmetric. Furthermore, the layers become thinner as Inline graphic increases, a clear manifestation of the enhanced convective circulation promoted by the bubbles.

Fig. 8.

Fig. 8.

(A) Inline graphic versus Inline graphic near cold plate, and (B) Inline graphic versus Inline graphic near hot plate. Symbols are circles (RB), triangles Inline graphic, and squares Inline graphic. Boundary layer thickness Inline graphic based on the wall gradient is also indicated in B for single-phase convection. Here Inline graphic and Inline graphic.

For the hot plate, one can define the thermal boundary layer thickness as Inline graphic, where Inline graphic is the mean temperature gradient at the hot plate. Replotting the data of the bottom panel in Fig. 8 as functions of Inline graphic, we find that the three sets of data collapse on a single line in the range Inline graphic (Fig. 9). The small differences farther away from the wall reflect differences in the shape factor of the boundary layers.

Fig. 9.

Fig. 9.

Normalized mean temperature 1 − Inline graphic versus Inline graphic in the hot thermal boundary layer. Here ξ = 0.5, Ra = 2 × 108, and Nb = 0 (circles), 50,000 (triangles), and 150,000 (squares).

Summary and Conclusions

In summary, our investigation of a simple model of RB convection with boiling has demonstrated the effect of the degree of superheat and of the bubble number on heat transport. Comparison with existing data suggests a basic conformity of our results with some physical features of a real system. Vapor bubbles significantly enhance the heat transport primarily by increasing the strength of the circulatory motion in the cell. The velocity and thermal fluctuations of the boundary layers are increased and, by releasing their latent heat upon condensation in the bulk fluid, the bubbles also act as direct carriers of energy. We have shown that the heat transfer enhancement can be interpreted in terms of an enhanced buoyancy, which is shown in Eq. 1 and Fig. 4B. The relative effect of the bubbles diminishes as Inline graphic increases.

Acknowledgments

We thank G. Ahlers and F. Toschi for providing valuable data and numerical results and L. Biferale and C. Sun for helpful discussions. Computations have been performed on the Huygens cluster of SURFSara in Amsterdam. This research is part of the Foundation for Fundamental Research on Matter (FOM) and Industrial Partnership Program on Fundamentals of Heterogeneous Bubbly Flows. We acknowledge financial support from FOM and the National Computing Facilities sponsored by Netherlands Organization for Scientific Research. P.O. gratefully acknowledges support from Fund for Investments in Basic Research (FIRB) Grant RBFR08QIP5_001.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

References

  • 1.Dhir VK. Boiling heat transfer. Annu Rev Fluid Mech. 1998;30(1):365–401. [Google Scholar]
  • 2.Carey VP. Liquid-Vapor Phase-Change Phenomena. New York: Hemisphere; 1992. [Google Scholar]
  • 3. Lienhard JH IV, Lienhard JH V (2008) A Heat Transfer Textbook (Phlogistan Press, Cambridge, MA) 3rd Ed.
  • 4.Ahlers G, Grossmann S, Lohse D. Heat transfer and large scale dynamics in turbulent Rayleigh-Bénard convection. Rev Mod Phys. 2009;81(2):503. [Google Scholar]
  • 5.Lohse D, Xia KQ. Small-scale properties of turbulent Rayleigh-Bénard convection. Annu Rev Fluid Mech. 2010;42:335. [Google Scholar]
  • 6.Oresta P, Verzicco R, Lohse D, Prosperetti A. Heat transfer mechanisms in bubbly Rayleigh-Bénard convection. Phys Rev E Stat Nonlin Soft Matter Phys. 2009;80(2):503–537. doi: 10.1103/PhysRevE.80.026304. [DOI] [PubMed] [Google Scholar]
  • 7.Zhong JQ, Funfschilling D, Ahlers G. Enhanced heat transport by turbulent two-phase Rayleigh-Bénard convection. Phys Rev Lett. 2009;102(12):335–364. doi: 10.1103/PhysRevLett.102.124501. [DOI] [PubMed] [Google Scholar]
  • 8.Schmidt LE, et al. Modification of turbulence in Rayleigh-Bénard convection by phase change. New J Phys. 2011;13(2):025002. [Google Scholar]
  • 9.Lakkaraju R, et al. Effect of vapor bubbles on velocity fluctuations and dissipation rates in bubbly Rayleigh-Bénard convection. Phys Rev E Stat Nonlin Soft Matter Phys. 2011;84(3):036312. doi: 10.1103/PhysRevE.84.036312. [DOI] [PubMed] [Google Scholar]
  • 10.Balachandar S, Prosperetti A, editors. IUTAM Symposium on Computational Approaches to Multiphase Flow. The Netherlands: Springer, Dordrecht; 2004. [Google Scholar]
  • 11.Prosperetti A, Tryggvasonn G. Computational Methods for Multiphase Flow. Cambridge, UK: Cambridge Univ Press; 2007. [Google Scholar]
  • 12.Verzicco R, Camussi R. Numerical experiments on strongly turbulent thermal convection in a slender cylindrical cell. J Fluid Mech. 2003;477:19–49. [Google Scholar]
  • 13.Lakkaraju R, et al. Spatial dependence of fluctuations and flux in turbulent Rayleigh-Bénard convection. Phys Rev E Stat Nonlin Soft Matter Phys. 2012;86(5):056315. doi: 10.1103/PhysRevE.86.056315. [DOI] [PubMed] [Google Scholar]
  • 14.Stevens RJAM, Verzicco R, Lohse D. Radial boundary layer structure and Nusselt number in Rayleigh-Bénard convection. J Fluid Mech. 2010;643:495–507. [Google Scholar]
  • 15.Biferale L, Perlekar P, Sbragaglia M, Toschi F. Convection in multiphase fluid flows using lattice boltzmann methods. Phys Rev Lett. 2012;108(10):104502. doi: 10.1103/PhysRevLett.108.104502. [DOI] [PubMed] [Google Scholar]
  • 16.Ahlers G, et al. Non-Oberbeck-Boussinesq effects in strongly turbulent Rayleigh-Bénard convection. J Fluid Mech. 2006;569:409–445. [Google Scholar]

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES