ArticlePDF Available

Abstract and Figures

While the heat transfer and the flow dynamics in a cylindrical Rayleigh-Bénard (RB) cell are rather independent of the aspect ratio Γ (diameter/height) for large Γ, a small-Γ cell considerably stabilizes the flow and thus affects the heat transfer. Here, we first theoretically and numerically show that the critical Rayleigh number for the onset of convection at given Γ follows Ra_{c,Γ}∼Ra_{c,∞}(1+CΓ^{-2})^{2}, with C≲1.49 for Oberbeck-Boussinesq (OB) conditions. We then show that, in a broad aspect ratio range (1/32)≤Γ≤32, the rescaling Ra→Ra_{ℓ}≡Ra[Γ^{2}/(C+Γ^{2})]^{3/2} collapses various OB numerical and almost-OB experimental heat transport data Nu(Ra,Γ). Our findings predict the Γ dependence of the onset of the ultimate regime Ra_{u,Γ}∼[Γ^{2}/(C+Γ^{2})]^{-3/2} in the OB case. This prediction is consistent with almost-OB experimental results (which only exist for Γ=1, 1/2, and 1/3) for the transition in OB RB convection and explains why, in small-Γ cells, much larger Ra (namely, by a factor Γ^{-3}) must be achieved to observe the ultimate regime.
Content may be subject to copyright.
Aspect Ratio Dependence of Heat Transfer in a Cylindrical Rayleigh-B´enard Cell
Guenter Ahlers,4,1 Eberhard Bodenschatz ,1,3,7,8 Robert Hartmann ,2Xiaozhou He ,6,1 Detlef Lohse ,2,3,1
Philipp Reiter ,1Richard J. A. M. Stevens ,2Roberto Verzicco,5,9,2 Marcel Wedi,1Stephan Weiss ,1,3
Xuan Zhang ,1Lukas Zwirner ,1and Olga Shishkina 1,*
1Max Planck Institute for Dynamics and Self-Organization, 37077 Göttingen, Germany
2Physics of Fluids Group, J. M. Burgers Center for Fluid Dynamics and MESA+ Institute,
University of Twente, 7500 AE Enschede, Netherlands
3Max PlanckUniversity of Twente Center for Complex Fluid Dynamics, 7500 AE Enschede, Netherlands
4Department of Physics, University of California, Santa Barbara, California 93106, USA
5Dipartimento di Ingegneria Industriale, University of Rome Tor Vergata,Via del Politecnico 1, Roma 00133, Italy
6School of Mechanical Engineering and Automation, Harbin Institute of Technology, Shenzhen, 518055 China
7Institute for the Dynamics of Complex Systems, Georg-August-University Göttingen, 37073 Göttingen, Germany
8Laboratory of Atomic and Solid-State Physics and Sibley School of Mechanical and Aerospace Engineering,
Cornell University, Ithaca, New York 14853, USA
9Gran Sasso Science InstituteViale F. Crispi, 767100 LAquila, Italy
(Received 20 April 2021; accepted 13 January 2022; published 24 February 2022)
While the heat transfer and the flow dynamics in a cylindrical Rayleigh-B´enard (RB) cell are rather
independent of the aspect ratio Γ(diameter/height) for large Γ, a small-Γcell considerably stabilizes the flow
and thus affects the heat transfer. Here, we first theoretically and numerically show that the critical Rayleigh
number for the onset of convection at given Γfollows Rac;ΓRac;ð1þCΓ2Þ2, with C1.49 for
Oberbeck-Boussinesq (OB) conditions. We then show that, in a broad aspect ratio range ð1=32ÞΓ32,the
rescaling Ra RalRa½Γ2=ðCþΓ2Þ3=2collapses various OB numerical and almost-OB experimental
heat transport data NuðRa;ΓÞ. Our findings predict the Γdependence of the onset of the ultimate regime
Rau;Γ½Γ2=ðCþΓ2Þ3=2in the OB case. This prediction is consistent with almost-OB experimental results
(which only exist for Γ¼1,1=2, and 1=3) for the transition in OB RB convection and explains why, in small-Γ
cells, much larger Ra (namely, by a factor Γ3) must be achieved to observe the ultimate regime.
DOI: 10.1103/PhysRevLett.128.084501
Physics is abstraction, often assuming systems of infinite
size. In the real world, this is not possible and finite-size
effects come into play and thus must be understood. Here
we will do so for the Rayleigh-B´enard convection (RBC),
which has always been the most paradigmatic system to
study buoyancy driven heat transfer in turbulent flow [13],
which is of great importance in geophysical flows and in
industry. The dimensionless control parameters are the
Rayleigh number, the Prandtl number, and the aspect ratio
Γof the cell, defined, respectively, as
Ra αgΔH3=ðκνÞ;Pr ν=κ;ΓD=H; ð1Þ
where Hand Dare the height and diameter of the
cylindrical cell, αis the isobaric thermal expansion
coefficient, νis the kinematic viscosity, κis the thermal
diffusivity, gis the gravitational acceleration, and ΔTb
Ttis the temperature difference between the hot bottom
plate and the cold top plate. The boundary conditions (BCs)
are no-slip at all walls and the sidewalls are adiabatic.
Within the Oberbeck-Boussinesq (OB) approximation, the
flow dynamics for the velocity u, the temperature T, and the
kinematic pressure pis given by the continuity equation
·u¼0and the Navier-Stokes and convection-diffusion
equations
tuþu·uþp¼ν2uþαgTez;ð2Þ
tTþu·T¼κ2T: ð3Þ
The key response parameter is the Nusselt number (the
dimensionless heat transfer)
Nu huzTizκzhTiz
κΔ=H ¼H
κΔhuzT1;ð4Þ
Published by the American Physical Society under the terms of
the Creative Commons Attribution 4.0 International license.
Further distribution of this work must maintain attribution to
the author(s) and the published articles title, journal citation,
and DOI. Open access publication funded by the Max Planck
Society.
PHYSICAL REVIEW LETTERS 128, 084501 (2022)
0031-9007=22=128(8)=084501(6) 084501-1 Published by the American Physical Society
where h·izdenotes the average in time and over a horizontal
cross section at height zfrom the bottom and h·iis the time
and volume average.
One key questionclearly, since Kraichnans 1962
prediction of an ultimate regime [1315] (i.e., the asymp-
totic law of heat transport at fixed Pr and extremely
large Ra)is, what is the Nu(Ra) dependence for very
large Ra? However, achieving very large Ra and thus this
predicted ultimate regime is challenging, both experimen-
tally, as large-scale setups are required, and computationally,
as the number of grid points that can be handled is limited,
too. Driven by the aim to nonetheless achieve very large Ra,
one is tempted to perform experiments or simulations at as
small Γas possible. For a profound judgement on this, a
FIG. 1. Critical Rac;Γfor the onset of convection: Linear growth rates (colored vertically elongated boxes) from the linearized DNS
approach (
GOLDFISH
) compared to the neutral stability curves (blue lines) from the eigenvalue LSA for (a) 2D box with isothermal
sidewalls, (b) 2D box with adiabatic sidewalls, and (c) cylinder with adiabatic sidewall. Black lines show Rac;Γ¼1708ð1þC=Γ2Þ2
with a best-fit Cfor the linearized DNS data (dashed lines) and with theoretical Cfor isothermal sidewall (solid line). Pluses in (c) show
Rac;Γfrom the nonlinearized DNS data (
AF
i
D
)[4]. Temperature contours near the onset of convection are shown for some Γ, as obtained
from the linearized DNS. See details in [58] and the Supplemental Material [9].
FIG. 2. (a) Compensated Nu vs Ra, as obtained in OB experiments and DNSs of RBC in a cylinder for Pr 4.4(water) and different Γ.
Most data are for Γ¼1and 1=2, which form the shape of this dependence. The data for extremely small Γshow no discernible
dependence. (b) Compensated Nu vs Ra based on the proper length scale l, for the same data as in (a). In the main plot, the theoretical
value of C¼1.49 is taken, while in the inset C¼0.77, which corresponds to the best fit of the critical Rac;for the onset of convection.
Now the data for extremely small Γfollow the general trend.
PHYSICAL REVIEW LETTERS 128, 084501 (2022)
084501-2
good understanding of the Γdependence of the flow and the
heat transfer for small Γis mandatory. The Göttingen group
[34,3941,50,53] has built large-scale cylindrical cells with
1Γ1=3and heights up to H¼2.24 m, filled with
pressurized SF6(with low viscosity and nearly constant Pr)
and has experimentally studied the onset Rau;Γof the
ultimate regime in almost-OB RBC. Note that building
even larger setups is not prohibitive, but simply extremely
costly. The Göttingen group found that the onset occurs at
Ra around 1014 (consistent with the theoretical estimate of
Grossmann and Lohse [15]) and revealed a Γdependence as
Rau;ΓΓ3.04 [54]; i.e., smaller Γrequire considerably
larger Ra to observe the onset. Also Roche et al. [55,56], for
1.14 Γ0.23, found a strong Γdependence of Rau;Γwith
the same trend. Based on an analysis of different experi-
mental data [3943,50,55,5759], they also proposed that
for small Γthe onset Ra for the ultimate regime goes
approximately as Rau;ΓΓ3.
In fact, due to the stabilizing effect of the sidewalls in
small-Γcells, it is not surprising at all that flow transitions
are shifted toward much larger Ra. This already holds at the
onset of convection: While without lateral confinement
(i.e., Γ) this onset occurs at a critical Rac;1708
[60], for small Γthe critical Rac;Γis much larger [6171].In
the limit Γ0, Catton and Edwards [63] numerically
solved the linearized perturbation equations with approxi-
mate wall conditions and proposed the scaling Rac;ΓΓ4
for the onset Rac;Γin this limit.
In this Letter, we will derive the scaling relation Rac;Γ
Γ4for Γ0and, in fact, generalize it to any Γ, be it large
or small. We will then show that our numerically performed
linear stability analysis (LSA) is consistent with the
suggested generalized functional dependence of Rac;Γon
Γ. Our result can be cast in the form that the relevant length
scale in RBC is
lD= ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Γ2þC
p¼H= ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1þC=Γ2
q;ð5Þ
with a constant Cthat depends on the shape of the cell. We
then apply this insight to the fully turbulent case and are
able to collapse various heat transfer data NuðRa;ΓÞfrom
OB experiments and direct numerical simulations (DNSs)
for various 1=32 Γ32 onto one universal curve.
FIG. 3. (a) Compensated Nu vs Ra for OB RBC in a cylinder for Pr near 0.8 and in a 3D cell with periodic BCs for Pr ¼1and different
Γ. Vertical lines indicate the onset of the transition at high Ra, observed in Göttingen experiments (the onset moves to higher Ra with
decreasing Γ). (b) Compensated Nu vs Ra based on the proper length scale l, for the same data as in (a). Now the transition happens at
the same location for all Γ[the vertical lines from (a) merge into one line]. The here presented experimental data from Chavanne et al.
[42,43] hold δρ=ρ<0.2for the density variation and δκ=κ<0.2for the thermal diffusivity variation, as well as αΔ<0.2and
0.68 Pr 1, i.e., similar almost-OB conditions as in [34,3941,50] (however, in [34,3941,50] the upper bounds for the fluid
parameter variations are even slightly stricter). Data for Pr ¼0.74 (gas N2) and Pr ¼0.84 (gas SF6) were taken using the same apparatus
as in [47] but were not published there. The inset shows an enlargement at the highest Ra in normal representation for both axes (see also
Supplemental Material [9], which includes [1012]).
PHYSICAL REVIEW LETTERS 128, 084501 (2022)
084501-3
Theoretical background.We first recall that the mean
kinetic energy dissipation rate ϵuand the thermal dissipa-
tion rate ϵθfulfil the exact relations [72,73]
ϵuνuÞ2αghuzTν3
H4ðNu 1ÞRa
Pr2;ð6Þ
ϵθκTÞ2i¼ðκΔ2=H2ÞNu:ð7Þ
Decomposing the temperature field as
TTlþθ;T
lðzÞTbðz=HÞΔ;ð8Þ
and taking into account huziz¼0for any z, one obtains
huzTiz¼huzθizand, hence,
huzTi¼huzθi:ð9Þ
From (4) and (7)(9), we get
huzθi¼ðκH=ΔÞhðθÞ210Þ
and then with (6) and (1) we obtain
uÞ2Ra½κ=ðΔH3ÞhuzTi:
From this, applying successively (9), the Cauchy-Schwarz
inequality, and relation (10), we derive
Ra ¼ΔH3
κuÞ2i
huzTi¼ΔH3
κuÞ2ihuzθi
huzθi2
ΔH3
κuÞ2ihuzθi
hu2
zihθ2iH4uÞ2ihðθÞ2i
hu2ihθ2i:ð11Þ
For a slightly supercritical Ra Rac;Γthe flow is sym-
metric so that hu0and hθ0holds. Therefore, we
can apply the Poincar´e-Friedrichs inequality to the right-
hand side of (11) to obtain
Rac;ΓH4uÞ2ihðθÞ2i
hu2ihθ2iΛ2;ð12Þ
where Λis the smallest relevant eigenvalue of the Laplacian
in a cylindrical domain with a unit height and aspect ratio Γ,
for certain integers m,n, and k,
Λ¼m2π2þ4α2
nkΓ21þCΓ2:ð13Þ
For Dirichlet or Neumann boundary conditions, αnk are the
first relevant roots of the Bessel function Jnor of its
derivative, respectively. Under the assumption that the
relevant eigenvalues admit positive as well as negative
values of θand uin both horizontal and vertical directions,
we obtain an estimate of the smallest relevant value of Λfor
m¼2,n¼k¼1, leading to C1.49.
For an infinite fluid layer (or for a cell with an infinite
diameter D, i.e., Γ)Ra
c;1708. Using this, rela-
tions (13) and (12), under assumption that Γand Rac;are
independent parameters, we obtain
Rac;ΓRac;ð1þCΓ2Þ2ð14Þ
as estimate for the critical Rac;Γfor the onset of convection
in a container with finite aspect ratio Γ.
Similarly, we estimate the growth of Nu near Rac;Γfrom
(11), the Poincar´e-Friedrichs inequality, and hθ2iΔ2,
Ra ΛH2θÞ2i=hθ2iΛH2Δ2θÞ2i:ð15Þ
From (8),(7), and (15) we finally obtain Ra ΛðNu 1Þ,
which, when combined with (13), implies that close to the
onset of convection, the Nusselt number behaves as
Nu 1ð1þCΓ2Þ1Ra:ð16Þ
From this and the fact that, in the classical turbulent regime
(for not too small Pr and not extremely high Ra), Nu
roughly grows as Ra1=3, one can expect a collapse of the
OB numerical and experimental data for various Γ, if these
are plotted as fðNu 1ÞRa1=3against
RalRað1þCΓ2Þ3=2ð17Þ
(for fixed Pr). Close to the onset of convection, this
dependence reduces to fRa2=3
l, while in the developed,
statistically steady convective flow fRa0
lconst. The
variable Ralis nothing else but a Rayleigh number not
based on the cell height H, but on the proper length scale l,
Eq. (5). In the limit Γ, the length scale lequals H,
while for Γ0,itisD.
Numerical LSA.We have verified the estimate (14) for
the Γdependence of the critical Rac;Γfor the onset of
convection with linearized DNSs for the 2D and 3D cases
and with the eigenspectrum LSA for the 2D case. The
growth rates obtained with both methods are in a very good
agreement, see Figs. 1(a) and 1(b). The numerically
obtained Rac;Γas function of Γ[Eq. (14)] for the isothermal
sidewalls are in excellent agreement with the analytical
estimates. Equation (14) captures the trend and reflects well
also the shape of the neutral curve for the case of adiabatic
sidewalls. The best-fit constants C(C0.52 for the 2D
domain and C0.77 for the cylinder) are, however,
smaller than the theoretical predictions for the isothermal
sidewalls, see Figs. 1(b) and 1(c). Isosurfaces of the
temperature of the flow fields near the onset of convection
are shown for some Γin Fig. 1as well. The azimuthal-
mode transition found for the cylinder between Γ¼1and 2
is consistent with the experiments [68].
Comparison with heat transfer data from OB experi-
ments and DNS.Our above theoretical analysis has
PHYSICAL REVIEW LETTERS 128, 084501 (2022)
084501-4
suggested the rescaling Ra Ralas a central step to
collapse the heat transfer data NuðRa;ΓÞfor given Γ, see
Eq. (17). This rescaling reflects that the relevant length
scale in RBC for general Γis lD= ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Γ2þC
p¼
H= ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1þC=Γ2
p, see Eq. (5), and not simply the height
H. For large Γone recovers l¼H, but for small Γone has
l¼D. We will now check whether this collapse holds and
plot the compensated Nusselt number fðNu 1Þ=Ra1=3
from OB experiments and well-resolved DNSs [74] for
various Γ, both vs Ra and vs Ral(with C¼1.49). We do so
for two different Pr, namely, for water (Pr 4.4, Fig. 2) and
for gas (Pr 0.8, Fig. 3) at room temperature. While in
Figs. 2(a) and 3(a) [fðRaÞ], the data for small Γshow no
trend and seem to scatter, in Figs. 2(b) and 3(b) [fðRalÞ],
they nicely collapse on one curve and on the theoretical
curve of the unifying theory for turbulent thermal con-
vection [2830]. A comparison with non-OB data for
cryogenic gaseous helium [42,43,57,75,76] is given in
the Supplemental Material [9]. As the derivation of
the scaling relations is for OB conditions, we do not
expect non-OB data to fulfil these relations, and indeed,
in general, they do not (see [34,77,78] and Supplemental
Material [9]).
Let us now estimate the Γdependence of the onset of the
ultimate regime of thermal convection, i.e., Rau;Γ. (The
other aspects of the ultimate regime are beyond the scope of
this Letter.) The Γdependence of Rau;Γhas been observed
in the Göttingen data [34,3941,50], with increasing Rau;Γ
for decreasing 1Γ1=3; see the vertical lines for large
Ra in Fig. 3(a). However, as suggested by our theory, in the
rescaled Fig. 3(b), these vertical lines collapse at the same
Ral;u 2.4×1013. This implies that the Γdependence of
Rau;Γin the OB case is
Rau;ΓRal;u½Γ2=ðCþΓ2Þ3=2;ð18Þ
which for Γ1simplifies to the estimate Rau;ΓΓ3,in
agreement with the experimental data [54]. Note that in
Fig. 3the agreement between the derived relation (18) and
measurements is demonstrated for all available almost-OB
experimental data, that is, for Γ¼1,1=2, and 1=3. Figure 3
and Eq. (18) also show that the presented DNS for small Γ
by far do not have large enough Ra to see the expected
onset of the ultimate regime.
In conclusion, we have developed a theory to account for
the Γdependence of the heat transfer in buoyancy driven
convection under OB conditions in cylindrical cells. In
particular, we find the Γdependence of the onset of
convection Rac;Γ[Eq. (14), consistent with the LSA] and
of the onset of the ultimate regime Rau;Γ[Eq. (18),
consistent with the Göttingen experiments]. Both equations
reflect that the relevant length scale in OB RBC is
l¼D= ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Γ2þC
p¼H= ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1þC=Γ2
p, which only in the
limiting cases Γor Γ0become the cell height
Hor the cell diameter D, respectively. Speaking more
generally, our results show how strongly finite-size effects
affect scaling relations and that small-ΓOB DNSs or
(almost) OB experiments require much large Ra to achieve
the ultimate regime.
The authors acknowledge the Deutsche Forschu-
ngsgemeinschaft (SPP1881 Turbulent Superstructures
and Grants No. Sh405/7, No. Sh405/8, and No. Sh405/
10), the Twente Max-Planck Center, the European
Research Council (ERC Starting Grant No. 804283
UltimateRB), the National Natural Science Foundation
of China (Grant No. 91952101), PRACE (Projects
No. 2020235589 and No. 2020225335), and the Gauss
Centre for Supercomputing e.V. for providing computing
time in the GCS Supercomputer SuperMUC at Leibniz
Supercomputing Centre.
*Olga.Shishkina@ds.mpg.de
[1] G. Ahlers, S. Grossmann, and D. Lohse, Rev. Mod. Phys.
81, 503 (2009).
[2] D. Lohse and K.-Q. Xia, Annu. Rev. Fluid Mech. 42, 335
(2010).
[3] F. Chill`a and J. Schumacher, Eur. Phys. J. E 35, 58 (2012).
[4] E. P. van der Poel, R. Ostilla Mónico, J. Donners, and R.
Verzicco, Computers & Fluids 116, 10 (2015).
[5] V. Theofilis, Annu. Rev. Fluid Mech. 43, 319 (2011).
[6] P. J. Schmid, Annu. Rev. Fluid Mech. 39, 129 (2007).
[7] P. Reiter, X. Zhang, R. Stepanov, and O. Shishkina, J. Fluid
Mech. 913, A13 (2021).
[8] P. Reiter, https://github.com/preiter93/pylsa (2021).
[9] See Supplemental Material at http://link.aps.org/
supplemental/10.1103/PhysRevLett.128.084501 for linear
stability analysis and comparison with non-OB data, which
includes Refs. [1012].
[10] G. Grötzbach, J. Comput. Phys. 49, 241 (1983).
[11] L. Zwirner, A. Tilgner, and O. Shishkina, Phys. Rev. Lett.
125, 054502 (2020).
[12] K. P. Iyer, J. D. Scheel, J. Schumacher, and K. R.
Sreenivasanan, Proc. Natl. Acad. Sci. U.S.A. 117, 7594
(2020).
[13] R. Kraichnan, Phys. Fluids 5, 1374 (1962).
[14] E. A. Spiegel, Annu. Rev. Astron. Astrophys. 9, 323 (1971).
[15] S. Grossmann and D. Lohse, Phys. Fluids 23, 045108
(2011).
[16] R. Hartmann, K. Chong, R. Stevens, R. Verzicco, and D.
Lohse, Europhys. Lett. 135, 24004 (2021).
[17] A. Nikolaenko, E. Brown, D. Funfschilling, and G. Ahlers,
J. Fluid Mech. 523, 251 (2005).
[18] X. Zhang, R. E. Ecke, and O. Shishkina, J. Fluid Mech. 915,
A62 (2021).
[19] S. Weiss and G. Ahlers, J. Fluid Mech. 676, 5 (2011).
[20] R. J. A. M. Stevens, J. Overkamp, D. Lohse, and H. J. H.
Clercx, Phys. Rev. E 84, 056313 (2011).
[21] R. J. A. M. Stevens, H. J. H. Clercx, and D. Lohse, Phys.
Rev. E 86, 056311 (2012).
[22] D. Funfschilling, E. Brown, A. Nikolaenko, and G. Ahlers,
J. Fluid Mech. 536, 145 (2005).
PHYSICAL REVIEW LETTERS 128, 084501 (2022)
084501-5
[23] S. Horn and O. Shishkina, Phys. Fluids 26, 055111 (2014).
[24] J.-Q. Zhong and G. Ahlers, J. Fluid Mech. 665, 300 (2010).
[25] R. Lakkaraju, R. J. A. M. Stevens, R. Verzicco, S.
Grossmann, A. Prosperetti, C. Sun, and D. Lohse, Phys.
Rev. E 86, 056315 (2012).
[26] O. Shishkina and A. Thess, J. Fluid Mech. 633, 449 (2009).
[27] O. Shishkina, M. S. Emran, S. Grossmann, and D. Lohse,
Phys. Rev. Fluids 2, 103502 (2017).
[28] S. Grossmann and D. Lohse, J. Fluid Mech. 407, 27 (2000).
[29] S. Grossmann and D. Lohse, Phys. Rev. Lett. 86, 3316
(2001).
[30] R. J. A. M. Stevens, E. P. van der Poel, S. Grossmann, and
D. Lohse, J. Fluid Mech. 730, 295 (2013).
[31] L. Zwirner and O. Shishkina, J. Fluid Mech. 850, 984
(2018).
[32] R. J. A. M. Stevens, D. Lohse, and R. Verzicco, in Direct
and Large Eddy Simulation XII, edited by M. García-
Villalba, H. Kuerten, and M. Salvetti (Springer, Cham,
2020), Vol. 27, p. 215, 10.1007/978-3-030-42822-8_28.
[33] R. J. A. M. Stevens, A. Blass, X. Zhu, R. Verzicco, and D.
Lohse, Phys. Rev. Fluids 3, 041501(R) (2018).
[34] X. He, E. Bodenschatz, and G. Ahlers, J. Fluid Mech. 931,
A7 (2022).
[35] J. Bailon-Cuba, M. S. Emran, and J. Schumacher, J. Fluid
Mech. 655, 152 (2010).
[36] R. J. A. M. Stevens, R. Verzicco, and D. Lohse, J. Fluid
Mech. 643, 495 (2010).
[37] X. Zhang, D. P. M. van Gils, S. Horn, M. Wedi, L. Zwirner,
G. Ahlers, R. E. Ecke, S. Weiss, E. Bodenschatz, and O.
Shishkina, Phys. Rev. Lett. 124, 084505 (2020).
[38] M. Wedi, D. P. van Gils, S. Weiss, and E. Bodenschatz,
J. Fluid Mech. 912, A30 (2021).
[39] G. Ahlers, X. He, D. Funfschilling, and E. Bodenschatz,
New J. Phys. 14, 103012 (2012).
[40] X. He, D. Funfschilling, H. Nobach, E. Bodenschatz, and G.
Ahlers, Phys. Rev. Lett. 108, 024502 (2012).
[41] X. He, D. Funfschilling, H. Nobach, E. Bodenschatz, and G.
Ahlers, Phys. Rev. Lett. 110, 199401 (2013).
[42] X. Chavanne, F. Chilla, B. Castaing, B. Hebral, B. Chabaud,
and J. Chaussy, Phys. Rev. Lett. 79, 3648 (1997).
[43] X. Chavanne, F. Chill`a, B. Chabaud, B. Castaing, and B.
ebral, Phys. Fluids 13, 1300 (2001).
[44] M. S. Emran and O. Shishkina, J. Fluid Mech. 882,A3
(2020).
[45] J. D. Scheel and J. Schumacher, Phys. Rev. Fluids 2, 123501
(2017).
[46] Data acquired using the small convection apparatus as
described in [47].
[47] S. Weiss, P. Wei, and G. Ahlers, Phys. Rev. E 93, 043102
(2016).
[48] G. L. Kooij, M. A. Botchev, E. M. A. Frederix, B. J. Geurts,
S. Horn, D. Lohse, E. P. van der Poel, O. Shishkina,
R. J. A. M. Stevens, and R. Verzicco, Comput. Fluids
166, 1 (2018).
[49] S. Wagner, O. Shishkina, and C. Wagner, J. Fluid Mech.
697, 336 (2012).
[50] X. He, D. Funfschilling, E. Bodenschatz, and G. Ahlers,
New J. Phys. 14, 063030 (2012).
[51] O. Shishkina and C. Wagner, Phys. Fluids 19, 085107
(2007).
[52] O. Shishkina and C. Wagner, J. Fluid Mech. 546, 51 (2006).
[53] X. He, E. Bodenschatz, and G. Ahlers, Proc. Natl. Acad.
Sci. U.S.A. 117, 30022 (2020).
[54] E. Bodenschatz, X. He, D. van Gils, and G. Ahlers, 15th
European Turbulence Conference (Delft University of
Technology, Netherlands, 2015), http://etc15.fyper.com/
proceedings/proceedings/documents/210.pdf.
[55] P.-E. Roche, F. Gauthier, R. Kaiser, and J. Salort, New J.
Phys. 12, 085014 (2010).
[56] P.-E. Roche, New J. Phys. 22, 073056 (2020).
[57] J. J. Niemela, L. Skrbek, K. R. Sreenivasan, and R. J.
Donnely, Nature (London) 404, 837 (2000).
[58] J. J. Niemela and K. R. Sreenivasan, J. Fluid Mech. 481, 355
(2003).
[59] J. J. Niemela and K. R. Sreenivasan, J. Low Temp. Phys.
143, 163 (2006).
[60] S. Chandrasekhar, Hydrodynamic and Hydromagnetic
Stability (Clarendon Press, Oxford, 1961).
[61] G. Charlson and R. Sani, Int. J. Heat Mass Transfer 13, 1479
(1970).
[62] I. Catton, J. Heat Transfer 92, 186 (1970).
[63] I. Catton and D. K. Edwards, AlChE J. 16, 594 (1970).
[64] J. C. Buell and I. Catton, J. Heat Transfer 105, 255 (1983).
[65] J. Mizushima, J. Phys. Soc. Jpn. 64, 2420 (1995).
[66] A. Y. Gelfgat, J. Comput. Phys. 156, 300 (1999).
[67] K. Boronska and L. S. Tuckerman, J. Fluid Mech. 559, 279
(2006).
[68] F. Hebert, R. Hufschmid, J. Scheel, and G. Ahlers, Phys.
Rev. E 81, 046318 (2010).
[69] B. F. Wang, D. J. Ma, C. Chen, and D. J. Sun, J. Fluid Mech.
711, 27 (2012).
[70] J. Yu, A. Goldfaden, M. Flagstad, and J. D. Scheel, Phys.
Fluids 29, 024107 (2017).
[71] O. Shishkina, Phys. Rev. Fluids 6, 090502 (2021).
[72] L. N. Howard, in Applied Mechanics, edited by H. Görtler
(Springer, Berlin, 1966), pp. 11091115.
[73] E. Siggia, Annu. Rev. Fluid Mech. 26, 137 (1994).
[74] O. Shishkina, R. J. A. M. Stevens, S. Grossmann, and D.
Lohse, New J. Phys. 12, 075022 (2010).
[75] P. Urban, P. Hanzelka, V. Musilova, T. Kralik, M. L. Mantia,
A. Srnka, and L. Skrbek, New J. Phys. 16, 053042
(2014).
[76] S. Weiss, X. He, G. Ahlers, E. Bodenschatz, and O.
Shishkina, J. Fluid Mech. 851, 374 (2018).
[77] D. D. Gray and A. Giorgini, Int. J. Heat Mass Transfer 19,
545 (1976).
[78] D. Funfschilling, E. Bodenschatz, and G. Ahlers, Phys. Rev.
Lett. 103, 014503 (2009).
PHYSICAL REVIEW LETTERS 128, 084501 (2022)
084501-6
... ± 0.04) × 10 −4 mol/m 3 Pa and P 0 = 1.0 bar, D = (1.85 ± 0.02) × 10 −9 m 2 /s the diffusion coefficient of CO 2 in water, H = 17.6 ± 0.35 mm the height of the liquid barrier, and ν = 9.5 × 10 −7 m 2 /s the kinematic viscosity of water [8,30,31]. We obtain Ra H ≈ (8.8 ± 0.5) × 10 6 , which is well above the critical Rayleigh number, Ra H,c = 1.29 × 10 6 , based on the minimal aspect ratio (Γ max = d/H = 0.17) of our experimental setup [32]. ...
... ± 0.6) × 10 3 , by taking the average critical value for the twelve experiments shown in figure 6. We compare this value to the critical Rayleigh number from Ahlers et al. for Rayleigh-Bénard convection in a cylinder with adiabatic sidewalls, which we believe to be the closest available approximation to our system [32]: ...
... intersect, is exactly δ = δ * = 1.13 mm, which is reasonably close to z 1 , with a corresponding Ra δ * = Ra c = 2.51 × 10 3 . This further emphasises the difficulty in defining the thickness for the self-similar diffusion boundary layer, as by selecting a lower intensity threshold, and thus higher concentration cut-off C δ , we could have reproduced the prediction from Ahlers et al. [32]. ...
Preprint
Full-text available
The dissolution and subsequent mass transfer of carbon dioxide gas into liquid barriers plays a vital role in many environmental and industrial applications. In this work, we study the downward dissolution and propagation dynamics of CO2 into a vertical water barrier confined to a narrow vertical glass cylinder, using both experiments and direct numerical simulations. Initially, the dissolution of CO2 results in the formation of a CO2-rich water layer, which is denser in comparison to pure water, at the top gas-liquid interface. Continued dissolution of CO2 into the water barrier results in the layer becoming gravitationally unstable, leading to the onset of buoyancy driven convection and, consequently, the shedding of a buoyant plume. By adding sodium fluorescein, a pH-sensitive fluorophore, we directly visualise the dissolution and propagation of the CO2 across the liquid barrier. Tracking the CO2 front propagation in time results in the discovery of two distinct transport regimes, a purely diffusive regime and an enhanced diffusive regime. Using direct numerical simulations, we are able to successfully explain the propagation dynamics of these two transport regimes in this laterally strongly confined geometry, namely by disentangling the contributions of diffusion and convection to the propagation of the CO2 front.
... In a recent meta study by Ahlers et al. (2022), data from a great number of experiments were compiled and a correction with respect to the aspect ratio of the convection cell was proposed, which improved the collapse of the data. The data were divided into two sets, where Pr ≈ 4.4 in the first and Pr ≈ 0.8 in the second. ...
... Only after the data have been corrected the curves show a common point of transition. Nevertheless, Ahlers et al. (2022) ...
... The thickness of the thermal boundary layers can be estimated from the prefactor in Nu = aRa 1/3 , as δ T ≈ 0.5a −3/4 η B . With a ≈ 0.05 (Iyer et al. 2020;Ahlers et al. 2022), we obtain δ T ≈ 5η B , which seems reasonable. Sun, Cheung & Xia (2008) report δ T = 0.58 mm and η = 0.4 mm from an experiment at Ra = 2.5 × 10 10 with water (Pr = 4.3). ...
Article
Full-text available
We consider the Nusselt–Rayleigh number problem of Rayleigh–Bénard convection and make the hypothesis that the velocity and thermal boundary layer widths, $\delta _u$ and $\delta _T$ , in the absence of a strong mean flow are controlled by the dissipation scales of the turbulence outside the boundary layers and, therefore, are of the order of the Kolmogorov and Batchelor scales, respectively. Under this assumption, we derive $Nu \sim Ra^{1/3}$ in the high $Ra$ limit, independent of the Prandtl number, $\delta _T/L \sim Ra^{-1/3}$ and $\delta _u/L \sim Ra^{-1/3} Pr^{1/2}$ , where $L$ is the height of the convection cell. The scaling relations are valid as long as the Prandtl number is not too far from unity. For $Pr \sim 1$ , we make a more general ansatz, $\delta _u \sim \nu ^{\alpha }$ , where $\nu$ is the kinematic viscosity and assume that the dissipation scales as $\sim u^3/L$ , where $u$ is a characteristic turbulent velocity. Under these assumptions we show that $Nu \sim Ra^{\alpha /(3-\alpha )}$ , implying that $Nu \sim Ra^{1/5}$ if $\delta _u$ were scaling as in a Blasius boundary layer and $Nu \sim Ra^{1/2}$ (with some logarithmic correction) if it were scaling as in a standard turbulent shear boundary layer. It is argued that the boundary layers will retain the intermediate scaling $\alpha = 3/4$ in the limit of high $Ra$ .
... The heat transport properties are then related to the scaling behavior between the dimensionless heat flux (characterized by the Nusselt number Nu) and the dimensionless temperature difference (characterized by the Rayleigh number Ra), i.e., Nu ∼ Ra β where β is the scaling exponent. Decades of studies on RB setup show the emergence of universal scaling exponent in the constitutive law [9][10][11][12][13][14][15][16][17][18] . Typically, one theoretically arguments β = 1/3 from the elegant theory of marginal stability 9,16,18 , or β ≈ 0.3 from experimental observations 13 in the classical regime, and β = 1/2 in the ultimate regime predicted by a mixing length model assuming that the heat flux is fully controlled by turbulence 11,12 . ...
... Decades of studies on RB setup show the emergence of universal scaling exponent in the constitutive law [9][10][11][12][13][14][15][16][17][18] . Typically, one theoretically arguments β = 1/3 from the elegant theory of marginal stability 9,16,18 , or β ≈ 0.3 from experimental observations 13 in the classical regime, and β = 1/2 in the ultimate regime predicted by a mixing length model assuming that the heat flux is fully controlled by turbulence 11,12 . Both heat transport scaling relations are extensively examined by various experimental and numerical investigations [13][14][15][16][17][18] . ...
... Typically, one theoretically arguments β = 1/3 from the elegant theory of marginal stability 9,16,18 , or β ≈ 0.3 from experimental observations 13 in the classical regime, and β = 1/2 in the ultimate regime predicted by a mixing length model assuming that the heat flux is fully controlled by turbulence 11,12 . Both heat transport scaling relations are extensively examined by various experimental and numerical investigations [13][14][15][16][17][18] . ...
Preprint
Full-text available
The emergence of unified constitutive law is a hallmark of convective turbulence, i.e., $Nu \sim Ra^\beta$ with $\beta \approx 0.3$ in the classical and $\beta=1/2$ in the ultimate regime, where the Nusselt number $Nu$ measures the global heat transport and the Rayleigh number $Ra$ quantifies the strength of thermal forcing. In recent years, vibroconvective flows have been attractive due to its ability to drive flow instability and generate ``artificial gravity'', which have potential to effective heat and mass transport in microgravity. However, the existence of constitutive laws in vibroconvective turbulence remains unclear. To address this issue, we carry out direct numerical simulations in a wide range of frequencies and amplitudes, and report that the heat transport exhibits a universal scaling law $Nu \sim a^{-1} Re_\mathrm{os}^\beta$ where $a$ is the vibration amplitude, $Re_\mathrm{os}$ is the oscillational Reynolds number, and $\beta$ is the universal exponent. We find that the dynamics of boundary layers plays an essential role in vibroconvective heat transport, and the $Nu$-scaling exponent $\beta$ is determined by the competition between the thermal boundary layer (TBL) and vibration-induced oscillating boundary layer (OBL). Then a physical model is proposed to explain the change of scaling exponent from $\beta=2$ in the OBL-dominant regime to $\beta = 4/3$ in the TBL-dominant regime. We conclude that vibroconvective turbulence in microgravity defines a distinct universality class of convective turbulence. This work elucidates the emergence of universal constitutive laws in vibroconvective turbulence, and opens up a new avenue for generating a controllable effective heat transport under microgravity or even microfluidic environment in which gravity is nearly absent.
... Thus the onset Rayleigh number for convection Ra c cannot be determined directly. Following Wei (2021), we determine Ra c from the normalised time-averaged flow strength δ / T. When convection sets in, the first unstable mode is the azimuthal m = 1 mode in a cylindrical cell with Γ = 1 (Hébert et al. 2010;Ahlers et al. 2022). Thus Ra c can be determined once δ / T is larger than the experimentally detectable temperature differences. ...
... It is seen that E 1 / E t is always larger than 0.85 in liquid metal convection, suggesting the LSF is in the form of a single-roll structure. When Ra < Ra t , the LSF is the cell structure observed just beyond the onset of convection, i.e. the m = 1 azimuthal mode (Hébert et al. 2010;Ahlers et al. 2022 liquid metal convection is a residual of the cell structure near the onset of convection. ...
Article
Full-text available
We present an experimental study of Rayleigh–Bénard convection using liquid metal alloy gallium-indium-tin as the working fluid with a Prandtl number of $Pr=0.029$ . The flow state and the heat transport were measured in a Rayleigh number range of $1.2\times 10^{4} \le Ra \le 1.3\times 10^{7}$ . The temperature fluctuation at the cell centre is used as a proxy for the flow state. It is found that, as $Ra$ increases from the lower end of the parameter range, the flow evolves from a convection state to an oscillation state, a chaotic state and finally a turbulent state for $Ra>10^5$ . The study suggests that the large-scale circulation in the turbulent state is a residual of the cell structure near the onset of convection, which is in contrast with the case of $Pr\sim 1$ , where the cell structure is transiently replaced by high order flow modes before the emergence of the large-scale circulation in the turbulent state. The evolution of the flow state is also reflected by the heat transport characterised by the Nusselt number $Nu$ and the probability density function (p.d.f.) of the temperature fluctuation at the cell centre. It is found that the effective local heat transport scaling exponent $\gamma$ , i.e. $Nu\sim Ra^{\gamma }$ , changes continuously from $\gamma =0.49$ at $Ra\sim 10^4$ to $\gamma =0.25$ for $Ra>10^6$ . Meanwhile, the p.d.f. at the cell centre gradually evolves from a Gaussian-like shape before the transition to turbulence to an exponential-like shape in the turbulent state. For $Ra>10^6$ , the flow shows self-similar behaviour, which is revealed by the universal shape of the p.d.f. of the temperature fluctuation at the cell centre and a $Nu=0.19Ra^{0.25}$ scaling for the heat transport.
... The resulting dynamical system is governed by the Rayleigh number Ra = gα∆T H 3 /(νκ) and the Prandtl number Pr = ν/κ which are defined by acceleration due to gravity g, the thermal expansion coefficient α, the temperature difference between heating and cooling plate ∆T , the domain height H, the kinematic viscosity ν and the thermal diffusivity κ of the fluid. Additionally, the aspect ratio Γ = W/H as the ratio of domain width W and height H and the container's shape affects the flow [39,40]. In the present large aspect ratio experiment, so-called turbulent superstructures emerge [41][42][43][44]. ...
Preprint
Full-text available
The measurement of the transport of scalar quantities within flows is oftentimes laborious, difficult or even unfeasible. On the other hand, velocity measurement techniques are very advanced and give high-resolution, high-fidelity experimental data. Hence, we explore the capabilities of a deep learning model to predict the scalar quantity, in our case temperature, from measured velocity data. Our method is purely data-driven and based on the u-net architecture and, therefore, well suited for planar experimental data. We demonstrate the applicability of the u-net on experimental temperature and velocity data, measured in large aspect ratio Rayleigh-Bénard convection at Pr = 7.1 and Ra = 2 × 10 5 , 4 × 10 5 , 7 × 10 5 . We conduct a hyper-parameter optimization and ablation study to ensure appropriate training convergence and test different architectural variations for the u-net. We test two application scenarios that are of interest to experimentalists. One, in which the u-net is trained with data of the same experimental run and one in which the u-net is trained on data of different Ra . Our analysis shows that the u-net can predict temperature fields similar to the measurement data and preserves typical spatial structure sizes. Moreover, the analysis of the heat transfer associated with the temperature showed good agreement when the u-net is trained with data of the same experimental run. The relative difference between measured and reconstructed local heat transfer of the system characterized by the Nusselt number Nu is between 0.3% and 14.1% depending on Ra . We conclude that deep learning has the potential to supplement measurements and can partially alleviate the expense of additional measurement of the scalar quantity.
... For any given Pr and Ra-range, the theory provides accurate predictions of the value of γ, for containers of aspect ratio Γ ≳ 1. For Γ ≪ 1, the data can be rescaled according to the method suggested in [22,1], which we do not discuss here, as in the present study Γ = 1. ...
Preprint
Full-text available
In magnetoconvection, the flow of electromagnetically conductive fluid is driven by a combination of buoyancy forces, which create the fluid motion due to thermal expansion and contraction, and Lorentz forces, which distort the convective flow structure in the presence of a magnetic field. The differences in the global flow structures in the buoyancy-dominated and Lorentz-force-dominated regimes lead to different heat transport properties in these regimes, reflected in distinct dimensionless scaling relations of the global heat flux (Nusselt number $\textrm{Nu}$) versus the strength of buoyancy (Rayleigh number $\textrm{Ra}$) and electromagnetic forces (Hartmann number $\textrm{Ha}$). Here, we propose a theoretical model for the transition between these two regimes for the case of a quasistatic vertical magnetic field applied to a convective fluid layer confined between two isothermal, a lower warmer and an upper colder, horizontal surfaces. The model suggests that the scaling exponents $\gamma$ in the buoyancy-dominated regime, $\textrm{Nu}\sim\textrm{Ra}^\gamma$, and $\xi$ in the Lorentz-force-dominated regime, $\textrm{Nu}\sim(\textrm{Ha}^{-2}\textrm{Ra})^\xi$, are related as $\xi=\gamma/(1-2\gamma)$, and the onset of the transition scales with $\textrm{Ha}^{-1/\gamma}\textrm{Ra}$. These theoretical results are supported by our Direct Numerical Simulations for $10\leq \textrm{Ha}\leq2000$, Prandtl number $\textrm{Pr}=0.025$ and $\textrm{Ra}$ up to $10^9$ and data from the literature.
... The phase diagram in Figure 6 is obtained in rectangular convection cells by narrowing the width only. If the geometrical confinement is applied to both lateral directions, a recent study [42] showed that the critical Ra number for the onset of convection follows a power law of Ra c ∼ 1708(1 + C/Γ 2 ) 2 under Oberbeck-Boussinesq conditions, where C is a constant that depends on the shape of the convection cells. It is not clear at this stage how the boundaries of the plume-controlled regime (dotted and dashed lines in Figure 6) will be reshaped in convection systems with other geometries. ...
Article
Full-text available
Tuning transport properties through the manipulation of elementary structures has received a great success in many areas, such as condensed matter physics. However, the ability to manipulate coherent structures in turbulent flows is much less explored. This article reviews a recently discovered mechanism of tuning turbulent heat transport via coherent structure manipulation. We first show how this mechanism can be realized by applying simple geometrical confinement to a classical thermally-driven turbulence, which leads to the condensation of elementary coherent structures and significant heat-transport enhancement, despite the resultant slower flow. Some potential applications of this new paradigm in passive heat management are also discussed. We then explain how the heat transport behaviors in seemingly different turbulence systems can be understood by this unified framework of coherent structure manipulation. Several future directions in this research area are also outlined.
Article
The dissolution and subsequent mass transfer of carbon dioxide gas into liquid barriers plays a vital role in many environmental and industrial applications. In this work, we study the downward dissolution and propagation dynamics of CO2 into a vertical water barrier confined to a narrow vertical glass cylinder, using both experiments and direct numerical simulations. Initially, the dissolution of CO2 results in the formation of a CO2-rich water layer, which is denser in comparison to pure water, at the top gas-liquid interface. Continued dissolution of CO2 into the water barrier results in the layer becoming gravitationally unstable, leading to the onset of buoyancy-driven convection and, consequently, the shedding of a buoyant plume. By adding sodium fluorescein, a pH-sensitive fluorophore, we directly visualize the dissolution and propagation of the CO2 across the liquid barrier. Tracking the CO2 front propagation in time results in the discovery of two distinct transport regimes, a purely diffusive regime and an enhanced diffusive regime. Using direct numerical simulations, we are able to successfully explain the propagation dynamics of these two transport regimes in this laterally strongly confined geometry, namely by disentangling the contributions of diffusion and convection to the propagation of the CO2 front.
Article
Full-text available
Moderate spatial confinement enhances the heat transfer in turbulent Rayleigh-B\'enard (RB) convection [Chong et al., PRL 115, 264503 (2015)]. Here, by performing direct numerical simulations, we answer the question how the shape of the RB cell affects this enhancement. We compare three different geometries: a box with rectangular base (i.e., stronger confined in one horizontal direction), a box with square base (i.e., equally confined in both horizontal directions), and a cylinder (i.e., symmetrically confined in the radial direction). In all cases the confinement can be described by the same confinement parameter Γ^{−1}, given as height-over-width aspect ratio. The explored parameter range is 1≤Γ^{−1}≤64$, 10^7≤\Ra≤10^{10} for the Rayleigh number, and a Prandtl number of Pr=4.38. We find that both the optimal confinement parameter Γ^{−1}_opt for maximal heat transfer and the actual heat transfer enhancement strongly depend on the cell geometry. The differences can be explained by the formation of different vertically-coherent flow structures within the specific geometries. The enhancement is largest in the cylindrical cell, owing to the formation of a domain-spanning flow structure at the optimal confinement parameter Γ^{−1}_opt.
Article
Full-text available
To study turbulent thermal convection, one often chooses a Rayleigh-Bénard flow configuration, where a fluid is confined between a heated bottom plate, a cooled top plate of the same shape, and insulated vertical sidewalls. When designing a Rayleigh-Bénard setup, for specified fluid properties under Oberbeck-Boussinesq conditions, the maximal size of the plates (diameter or area), and maximal temperature difference between the plates, Δmax, one ponders: Which shape of the plates and aspect ratio Γ of the container (ratio between its horizontal and vertical extensions) would be optimal? In this article, we aim to answer this question, where under the optimal container shape, we understand such a shape, which maximizes the range between the maximal accessible Rayleigh number and the critical Rayleigh number for the onset of convection in the considered setup, Rac,Γ. First we prove that Rac,Γ∝(1+cuΓ−2)(1+cθΓ−2), for some cu>0 and cθ>0. This holds for all containers with no-slip boundaries, which have a shape of a right cylinder, whose bounding plates are convex domains, not necessarily circular. Furthermore, we derive accurate estimates of Rac,Γ, under the assumption that in the expansions (in terms of the Laplace eigenfunctions) of the velocity and reduced temperature at the onset of convection, the contributions of the constant-sign eigenfunctions vanish, both in the vertical and at least in one horizontal direction. With that we derive Rac,Γ≈(2π)4(1+cuΓ−2)(1+cθΓ−2), where cu and cθ are determined by the container shape and boundary conditions for the velocity and temperature, respectively. In particular, for circular cylindrical containers with no-slip and insulated sidewalls, we have cu=j112/π2≈1.49 and cθ=(j̃11)2/π2≈0.34, where j11 and j̃11 are the first positive roots of the Bessel function J1 of the first kind or its derivative, respectively. For parallelepiped containers with the ratios Γx and Γy, Γy≤Γx≡Γ, of the side lengths of the rectangular plates to the cell height, for no-slip and insulated sidewalls we obtain Rac,Γ≈(2π)4(1+Γx−2)(1+Γx−2/4+Γy−2/4). Our approach is essentially different to the linear stability analysis, however, both methods lead to similar results. For Γ≲4.4, the derived Rac,Γ is larger than Jeffreys' result Rac,∞J≈1708 for an unbounded layer, which was obtained with linear stability analysis of the normal modes restricted to the consideration of a single perturbation wave in the horizontal direction. In the limit Γ→∞, the difference between Rac,Γ→∞=(2π)4 for laterally confined containers and Jeffreys' Rac,∞J for an unbounded layer is about 8.8%. We further show that in Rayleigh-Bénard experiments, the optimal rectangular plates are squares, while among all convex plane domains, circles seem to match the optimal shape of the plates. The optimal Γ is independent of Δmax and of the fluid properties. For the adiabatic sidewalls, the optimal Γ is slightly smaller than 1/2 (for cylinder, about 0.46), which means that the intuitive choice of Γ=1/2 in most Rayleigh-Bénard experiments is right and justified. For the given plate diameter D and maximal temperature difference Δmax, the maximal attainable Rayleigh number range is about 3.5 orders of magnitudes smaller than the order of the Rayleigh number based on D and Δmax. Deviations from the optimal Γ lead to a reduction of the attainable range, namely, as log10(Γ) for Γ→0 and as log10(Γ−3) for Γ→∞. Our theory shows that the relevant length scale in Rayleigh-Bénard convection in containers with no-slip boundaries is ℓ∼D/Γ2+cu=H/1+cu/Γ2. This means that in the limit Γ→∞, ℓ equals the cell height H, while for Γ→0, it is rather the plate diameter D.
Article
Full-text available
This work addresses the effect of travelling thermal waves applied at the fluid layer surface, on the formation of global flow structures in two-dimensional (2-D) and 3-D convective systems. For a broad range of Rayleigh numbers (10^3≤Ra≤10^7) and thermal wave frequencies (10^−4≤Ω≤10^0), we investigate flows with and without imposed mean temperature gradients. Our results confirm that the travelling thermal waves can cause zonal flows, i.e. strong mean horizontal flows. We show that the zonal flows in diffusion dominated regimes are driven purely by the Reynolds stresses and end up always travelling retrograde. In convection dominated regimes, however, mean flow advection, caused by tilted convection cells, becomes dominant. This generally leads to prograde directed mean zonal flows. By means of direct numerical simulations we validate theoretical predictions made for the diffusion dominated regime. Furthermore, we make use of the linear stability analysis and explain the existence of the tilted convection cell mode. Our extensive 3-D simulations support the results for 2-D flows and thus provide further evidence for the relevance of the findings for geophysical and astrophysical systems.
Article
Full-text available
The long-standing puzzle of diverging heat transport measurements at very high Rayleigh numbers (Ra) is addressed by a simple model based on well-known properties of classical boundary layers. The transition to the ‘ultimate state’ of convection in Rayleigh–Bénard cells is modeled as sub-critical transition controlled by the instability of large-scale boundary-layer eddies. These eddies are restricted in size either by the lateral wall or by the horizontal plates depending on the cell aspect ratio (in cylindrical cells, the cross-over occurs for a diameter-to-height ratio around 2 or 3). The large-scale wind known to settle across convection cells is assumed to have antagonist effects on the transition depending on its strength, leading to wind-immune, wind-hindered or wind-assisted routes to the ultimate regime. In particular winds of intermediate strength are assumed to hinder the transition by disrupting heat transfer, contrary to what is assumed in standard models. This phenomenological model is able to reconcile observations from more than a dozen of convection cells from Grenoble, Eugene, Trieste, Göttingen and Brno. In particular, it accounts for unexplained observations at high Ra, such as Prandtl number and aspect ratio dependences, great receptivity to details of the sidewall and differences in heat transfer efficiency between experiments.
Article
Full-text available
The large-scale circulation (LSC) of fluid is one of the main concepts in turbulent thermal convection as it is known to be important in global heat and mass transport in the system. In turbulent Rayleigh-Bénard convection (RBC) in slender containers, the LSC is formed of several dynamically changing convective rolls that are stacked on top of each other. The present study reveals the following two important facts: (i) the mechanism which causes the twisting and breaking of a single-roll LSC into multiple rolls is the elliptical instability and (ii) the heat and momentum transport in RBC, represented by the Nusselt (Nu) and Reynolds (Re) numbers, is always stronger (weaker) for smaller (larger) number n of the rolls in the LSC structure. Direct numerical simulations support the findings for n=1,…,4 and the diameter-to-height aspect ratio of the cylindrical container Γ=1/5, the Prandtl number Pr=0.1 and Rayleigh number Ra=5×105. Thus, Nu and Re are, respectively, 2.5 and 1.5 times larger for a single-roll LSC (n=1) than for a LSC with n=4 rolls.
Chapter
Full-text available
Heat transfer mediated by a fluid is omnipresent in nature as well as in technical applications and it is always among the fundamental mechanisms of the phenomena. The performance of modern computer processors has reached a plateau owing to the inadequacy of the fluid-based cooling systems to get rid of the heat flux which increases with the operating frequency [1]. On much larger spatial scales, circulations in the atmosphere and oceans are driven by temperature differences whose strength is key for the evolution of the weather and the stability of regional and global climate [2].
Article
We report measurements of the temperature frequency spectra $P(\,f, z, r)$ , the variance $\sigma ^2(z,r)$ and the Nusselt number $Nu$ in turbulent Rayleigh–Bénard convection (RBC) over the Rayleigh number range $4\times 10^{11} \underset{\smash{\scriptscriptstyle\thicksim}} { and for a Prandtl number $Pr \simeq ~0.8$ ( $z$ is the vertical distance from the bottom plate and $r$ is the radial position). Three RBC samples with diameter $D = 1.12$ m yet different aspect ratios $\varGamma \equiv D/L = 1.00$ , $0.50$ and $0.33$ ( $L$ is the sample height) were used. In each sample, the results for $\sigma ^2/\varDelta ^2$ ( $\varDelta$ is the applied temperature difference) in the classical state over the range $0.018 \underset{\smash{\scriptscriptstyle\thicksim}} { can be collapsed onto a single curve, independent of $Ra$ , by normalizing the distance $z$ by the thermal boundary layer thickness $\lambda = L/(2 Nu)$ . One can derive the equation $\sigma ^2/\varDelta ^2 = c_1\times \ln (z/\lambda )+c_2+c_3(z/\lambda )^{-0.5}$ from the observed $f^{-1}$ scaling of the temperature frequency spectrum. It fits the collapsed $\sigma ^2(z/\lambda )$ data in the classical state over the large range $20 \underset{\smash{\scriptscriptstyle\thicksim}} { . In the ultimate state ( $Ra \underset{\smash{\scriptscriptstyle\thicksim}} { > } Ra^*_2$ ) the data can be collapsed only when an adjustable parameter $\tilde \lambda = L/(2 \widetilde {Nu})$ is used to replace $\lambda$ . The values of $\widetilde {Nu}$ are larger by about 10 % than the experimentally measured $Nu$ but follow the predicted $Ra$ dependence of $Nu$ for the ultimate RBC regime. The data for both the global heat transport and the local temperature fluctuations reveal the ultimate-state transitions at $Ra^*_2(\varGamma )$ . They yield $Ra^*_2 \propto \varGamma ^{-3.0}$ in the studied $\varGamma$ range.