ArticlePDF Available

Abstract and Figures

Using complementary experiments and direct numerical simulations, we study turbulent thermal convection of a liquid metal (Prandtl number Pr0.03\textit {Pr}\approx 0.03 ) in a box-shaped container, where two opposite square sidewalls are heated/cooled. The global response characteristics like the Nusselt number Nu{\textit {Nu}} and the Reynolds number Re\textit {Re} collapse if the side height L is used as the length scale rather than the distance H between heated and cooled vertical plates. These results are obtained for various Rayleigh numbers 5×103RaH1085\times 10^3\leq {\textit {Ra}}_H\leq 10^8 (based on H ) and the aspect ratios L/H=1,2,3L/H=1, 2, 3 and 5 . Furthermore, we present a novel method to extract the wind-based Reynolds number, which works particularly well with the experimental Doppler-velocimetry measurements along vertical lines, regardless of their horizontal positions. The extraction method is based on the two-dimensional autocorrelation of the time–space data of the vertical velocity.
Content may be subject to copyright.
J. Fluid Mech. (2022), vol.932, A9, doi:10.1017/jfm.2021.977
Dynamics and length scales in vertical
convection of liquid metals
Lukas Zwirner1,, Mohammad S. Emran1, Felix Schindler2,SanjaySingh
2,
Sven Eckert2,TobiasVogt
2and Olga Shishkina1
1Max Planck Institute for Dynamics and Self-Organisation, Am Fassberg 17, 37077 Göttingen
2Helmholtz-Zentrum Dresden-Rossendorf, Bautzner Landstr. 400, 01328 Dresden
(Received 16 April 2021; revised 7 October 2021; accepted 1 November 2021)
Using complementary experiments and direct numerical simulations, we study turbulent
thermal convection of a liquid metal (Prandtl number Pr 0.03) in a box-shaped
container, where two opposite square sidewalls are heated/cooled. The global response
characteristics like the Nusselt number Nu and the Reynolds number Re collapse if the side
height Lis used as the length scale rather than the distance Hbetween heated and cooled
vertical plates. These results are obtained for various Rayleigh numbers 5 ×103RaH
108(based on H) and the aspect ratios L/H=1,2,3 and 5. Furthermore, we present a
novel method to extract the wind-based Reynolds number, which works particularly well
with the experimental Doppler-velocimetry measurements along vertical lines, regardless
of their horizontal positions. The extraction method is based on the two-dimensional
autocorrelation of the time–space data of the vertical velocity.
Key words: convection in cavities
1. Introduction
The understanding of turbulent thermal convection is of great importance for astrophysics,
geophysics, climate research and engineering purposes alike. Although convection has
been under investigation for centuries, even today researchers struggle to unveil its
complex nature. Usually, to study turbulent thermal convection, a model system is
considered, where the fluid is confined between horizontal plates heated from below
and cooled from above, commonly known as Rayleigh–Bénard convection (RBC)
(Bodenschatz, Pesch & Ahlers 2000; Ahlers, Grossmann & Lohse 2009; Chillà &
Schumacher 2012). In this work we study a different model system, where the fluid
Email address for correspondence: lukas.zwirner@ds.mpg.de
© The Author(s), 2021. Published by Cambridge University Press. This is an Open Access article,
distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.
org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium,
provided the original work is properly cited. 932 A9-1
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
is confined between vertical heated/cooled plates, also known as convection inside a
differentially heated enclosure, side-heated convection or vertical convection. Throughout
this work, we will refer to this system as vertical convection (VC).
RBC and VC are two limiting cases of the more general type of inclined convection
(Daniels, Wiener & Bodenschatz 2003; Chillà et al. 2004; Sun, Xi & Xia 2005; Ahlers,
Brown & Nikolaenko 2006; Weiss & Ahlers 2013; Shishkina & Horn 2016; Zwirner &
Shishkina 2018; Zwirner et al. 2020a) and therefore are of particular importance. Note,
that VC, unlike RBC, is inherently unstable even for the smallest temperature difference
between the plates.
Batchelor (1954) was one of the first to investigate VC for very small temperature
differences, and, in the beginning, the main interest was in studying the heat transport
through double layer windows. Therefore, the studied geometry was of very large aspect
ratio Lz/H, where Lzis the height of the convection cell and His the distance between
the heated and cooled vertical plates. Fujii et al. (1970) performed a detailed experimental
study on the evolution of boundary layers (BLs) and local heat transport in VC, using
two concentric cylinders, where the inner one was heated and the outer one cooled, and
using oil and water as working fluids. Belmonte, Tilgner & Libchaber (1995) used a cubic
cell and different gases of Prandtl number Pr 0.7 for Rayleigh numbers RaH1011
and found by shadowgraph visualization a stably stratified bulk, while the temperature
fluctuations were mainly observed close to the boundaries. Additionally, they measured
a scaling of the thermal BL thickness Ra0.29. Koster, Seidel & Derebail (1997)used
X-ray radiography to measure and visualize the density distribution inside a narrow VC
cell filled with liquid gallium. This was a milestone in the visualization of liquid metal
flows, however, it is only applicable to narrow cavities and also no conclusions about
turbulent convection could be drawn. Braunsfurth et al. (1997) conducted experiments
with liquid gallium inside a long VC cells of aspect ratios L/H=1/3and1/4, for small
Grashof numbers (Ra/Pr <5×104) and compared the results with two-dimensional
numerical simulations. Here, we investigate flow at much larger Ra and large aspect ratios.
As in RBC (Shishkina 2021), in VC, the aspect ratio of the container influences the
flow in a finite fluid layer confined between two differently heated plates (Batchelor 1954;
Bejan 1980,1985,2013; Paolucci 1994). However, the effects of the container’s size and
the distance between its walls have not been investigated separately. Understanding this
phenomenon is of special significance in convection, especially in a small-Pr fluid, as both
these length scales influence the flow field. In particular, this sort of flow configuration is
typically encountered in liquid metal batteries (Kelley & Weier 2018) – a technology with
potential in grid-scale storage. In this paper, we attempt to disentangle the aspect ratio
dependency, and shift the focus onto the two predominant length scales using numerical
and experimental data across several aspect ratios. Additionally, we also present a robust
new technique which is capable of computing the Reynolds number based on the wind
velocity.
There are only a few examples of ongoing investigations of VC and in addition to
Churchill & Chu (1975); Graebel (1981); Tsuji & Nagano (1988); Chen & Pearlstein
(1989); Paolucci (1990); Versteegh & Nieuwstadt (1999); Pallares et al. (2010); Kis &
Herwig (2012)and Wang et al. (2021); there are even more studies, but most of them
use air, oil or water with relatively large Prandtl numbers, Pr 0.7. However, recently,
liquid metals with very small Prandtl numbers, Pr 1, have become a particular focus
of investigations in the thermal convection community (King & Aurnou 2013; Scheel
& Schumacher 2016; Schumacher et al. 2016; Aurnou et al. 2018; Vogt et al. 2018a;
Zwirner & Shishkina 2018; Zürner et al. 2019; Zwirner et al. 2020a; Zwirner, Tilgner &
932 A9-2
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
Shishkina 2020b). The aim of the present work is to shed more light on VC of
low-Prandtl-number fluids.
One important step that led to a better understanding of RBC was the development
of the scaling theory by Grossmann & Lohse (2000,2002) (GLT). This theory is based
on the exact relationships between the heat transport, represented by the dimensionless
Nusselt number Nu, and the kinetic and thermal dissipation rates. The GLT assumes
various scaling regimes (Nu Raγ) depending on which part of the flow determines the
scalings of the dissipation rates: the BLs or the bulk. This theory also concludes that there
is no simple scaling law applicable for the entire range of Rayleigh numbers, but that the
exponent γchanges smoothly within these regimes. A similar approach has been applied
by Ng et al. (2015)toVC, however, the difficulty in this case is a non-closed term in the
relationship between Nu and the kinetic energy dissipation rate. Nevertheless, Ng et al.
(2015) concluded that a similar approach to the GLT is applicable to VC, when a suitable
closure model for that specific term is found.
Although one might deem the question about which scales determine the flow settled,
there is still active research on the most suitable scales. For example,Wei(2020) found
the proper scale for the Reynolds shear stress in a differentially heated vertical channel is
a mixed scale of the friction and the maximum mean velocity.
Scaling relations for heat transport and BL thicknesses in vertical convection have been
investigated in the past (Batchelor 1954; Gill 1966; Saville & Churchill 1969;Ostrach
1972). A recent approach agrees upon the same exponents for vertical convection in the
laminar regime, there the exponents were derived from BL theory and confirmed by direct
numerical simulations (DNS) in a theoretical work (Shishkina 2016).
It was found that for Pr 1 the Nusselt number and the wind-based Reynolds number
scale with respect to the Rayleigh and Prandtl numbers as
Nu Pr1/4Ra1/4and Rewind Pr1/2Ra1/2,(1.1a,b)
respectively. Note that the scaling relations are valid regardless of the length scale chosen
and therefore the indices Hand Lare omitted. However, the pre-factor depends on this
choice and in the following we show that the data collapse for different aspect ratios when
the scale is chosen appropriately.
The remaining article is structured as follows: in §2we introduce the experimental
and numerical set-up, in § 3we present a detailed comparison of the experimental and
numerical results and discuss the relevant length scale, heat and momentum transport and
finally we conclude with § 4.
2. Experimental and numerical methods
Here, we present our numerical and experimental methods. A sketch of the basic set-up
used in the experiments and the three-dimensional DNS is shown in figure 1.Weusea
rectangular cell with variable aspect ratio, L/H, where the heated and cooled plates are
squares of length L=Ly=Lz. The distance between the heated and cooled plates is H.
2.1. Numerical set-up
The incompressible Navier–Stokes equations in Oberbeck–Boussinesq approximation
Dtu=ν2up+αg(TT0)ˆz,(2.1)
932 A9-3
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
z
x
y
AB
H
Ly
L
z
T+T
g
Figure 1. Sketch of the VC cell with a heated wall at the left side (temperature T+) and cooled wall at the right
side (temperature T). The velocity is measured along two lines parallel to the z-direction, using the Doppler
probes Aand B. The direction of gravity is indicated by g. This sketch represents the set-up for the DNS as well
as the experiments.
DtT=κ2T,(2.2)
∇·u=0 (2.3)
are solved using the high-order finite volume code GOLDFISH (Kooij et al. 2018)in
Cartesian coordinates. Here, Dtdenotes the substantial derivative, u=(ux,uy,uz)the
velocity vector field, pis the reduced kinetic pressure, Tthe temperature, T+the
temperature of the hot vertical plate, Tthe temperature of the cold vertical plate, T0
is the mean temperature (T++T)/2, νis the kinematic viscosity and κthe thermal
diffusivity. The equations are transformed into their non-dimensional representation using
the distance Hbetween the hot and cold plates, the time tfHgHΔ)1/2and the
temperature difference ΔT+T,gis the gravitational acceleration. Thus, our VC
system depends on the following three non-dimensional input parameters:
RaHαgH3/(κν), Pr ν/κ, Γ =L/H,(2.4ac)
which are the Rayleigh number, Prandtl number and aspect ratio, respectively. All DNS
are conducted at Pr =0.03.
The computational mesh, used in the simulations, is clustered near the boundaries and
the largest simulation (RaH=108and Γ=1) has a resolution of 6743points, while the
averaging time is usually tavg150 tf. We start to collect the statistics after the flow
reaches a stationary state as indicated by the fluctuations of Nu(t)around a constant
long-time average. The number of points within the viscous BL is always Nδu4and
due to the small Prandtl number, the thermal BL is much thicker than the viscous one
and therefore both BLs are well resolved. Additionally, we conducted a mesh convergence
study and also ensure that our typical mesh distance hresolves the Kolmogorov scale η=
3u)1/4(where εuis the kinetic energy dissipation rate), i.e. hη, which represents
the smallest scales of the flow. A detailed list of the DNS resolution can be found in the
Appendix,table 1.
Important characteristic output quantities in VC are the Nusselt number, the Reynolds
number and the thermal/viscous BL thicknesses. The Nusselt number represents the heat
transport through the system and is defined as
NuHuxTκ∂x¯
TAyz
κΔ/H,(2.5)
932 A9-4
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
where ·Ayz means averaging over a slice parallel to the hot plate and ¯· averaging over time.
The thickness of the thermal BL is defined as usual: δθH/(2Nu). For the definition of
the Reynolds number, we follow Shishkina (2016) and introduce the wind velocity in a
similar manner as
Uwind max
xuzAyz ,(2.6)
and with this the wind-based Reynolds number
Rewind =LUwind
ν.(2.7)
Furthermore, we define the thickness of the viscous BL using the slope method (Zhou &
Xia 2010)
δuUwind
xuzAyz |x=0.(2.8)
Note that, for laminar flow,δuRe1/2, and using (1.1) we conclude that δuRa1/4.
For an example of the horizontal profile uzwe refer the reader to figure 8(c,d).
2.2. Laboratory set-up
Figure 1 shows a schematic drawing of the experimental set-up. The experiments were
performed in two rectangular vessels with a square vertical cross-section of L2=
200 mm ×200 mm and a distance between the vertical heated/cooled boundaries of H=
40 and H=66 mm, which gives aspect ratios of Γ=L/H=5andΓ3, respectively.
On the two square vertical surfaces of the vessel, the heat is introduced and removed via
two copper plates, which are tempered by a circulating water bath and the temperature is
controlled via thermocouples inside the copper plates.
The maximal thermal power input is P=1500 W in this study. The other sidewalls are
madeof30mmthickPolyvinylchloride. To minimize heat losses, the whole convection
cell is wrapped in a 30 mm thick closed-cell foam, which has a thermal conductivity
of approximately 0.036 W mK1and is 660 times lower that of the liquid metal
(24 W mK1).
The vessel is filled with the eutectic alloy GaInSn that has a melting temperature of Ts=
10.5C. At room temperature, it has the density ρ=6350 kg m3, the thermal expansion
coefficient α=1.24 ×1041K
1, the thermal conductivity λ=24.05 W mK1,
the kinematic viscosity ν=3.38 ×107m2s1and the thermal diffusivity κ=
1.05 ×105m2s1(Plevachuk et al. 2014). The corresponding Prandtl number is
Pr 0.03.
For the majority of the measurements, we keep the mean fluid temperature constant
at approximately 21 C. Only for the highest Ra do we have to increase the mean fluid
temperature up to 35 C. However, the temperature dependence of the material parameters
is comparatively low, thus the Prandtl number changes only by approximately 10%
(2.97 Pr/1023.25). Therefore, we consider possible non-Oberbeck–Boussinesq
effects to have a rather weak influence on the flow.
2.3. Measuring technique
In this study, the flow velocities are measured using ultrasonic Doppler velocimetry (UDV)
which is an established measurement technique for opaque fluids like liquid metals (Aubert
et al. 2001;Britoet al. 2001; Eckert & Gerbeth 2002; Gillet et al. 2007; Nataf et al. 2008;
932 A9-5
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
Vogt, Räbiger & Eckert 2014; Vogt et al. 2018a,b; Zürner et al. 2019,2020; Yang, Vogt &
Eckert 2021; Vogt, Horn & Aurnou 2021a; Vogt et al. 2021b). The used UDV system is a
DOP3010 (from Signal Processing SA, Lausanne) equipped with 8 MHz transducers. The
UDV transducers send a pulsed ultrasonic signal into the liquid metal which is reflected
by microscopic particles such as oxides. The position and velocity of the particles can
be determined from the transit time of the ultrasound and the phase shift of the echo
from subsequent echo pulses. This allows us to determine the beam-parallel velocity
distribution along the ultrasonic beam (Takeda 2012). The two ultrasonic sensors used in
this study measure velocities along the vertical direction (cf. figure 1). The distances from
the sensors to the heated plate are xp/H=0.25 and 0.75 for Γ=5, and xp/H=0.15
and 0.85 for the Γ=3. The measuring system records velocity profiles with a time
resolution of approximately 0.3s. The spatial resolution is approximately 1 mm in the
beam direction and 5 mm in the lateral direction due to the diameter of the ultrasound
emitting piezoelectric transducer.
The experiment is equipped with 22 thermocouples with nine being embedded in each
copper plate to measure the temperature drop across the fluid layer. The remaining four
thermocouples are attached to the water channels that temperate the copper plates, in order
to determine the temperature change of the circulating water (Tin Tout). Together with
the measurements of the water flow rate ˙
V, this enables the calculation of the convective
heat transport, which is expressed non-dimensionally by the Nusselt number
Nu =˙
Qtot
˙
Qcond =ρcp˙
V(Tin Tout)
λL2Δ/H,(2.9)
where cpis the isobaric heat capacity of water. There are small differences between the
measurements at the hot and cold plates due to thermal losses. The Nusselt numbers
presented here are the averages of the hot and cold measurement (Γ=5), or shown
separately (Γ=3).
2.4. Extraction of the Reynolds number
One objective of this work is to compare the simulations with the experiments, and
therefore we need to define a Reynolds number based on a certain velocity that is easily
measurable in both simulations and experiments. In the experiments, only limited velocity
data are available from the UDV probes, while the simulations provide access to the
complete velocity fields. Therefore, we can directly compare the two-dimensional velocity
fields uz(t,z)only at the Doppler probe positions (xp/H=1/4and3/4). Examples of
these data are visualized in figure 2(a,c,e). There are multiple possibilities to define a
characteristic velocity based on these data. Here, we investigate two candidates in detail,
a straightforward one (the maximal velocity) and a more sophisticated one (based on the
autocorrelation) which we define in the following.
The maximal velocity is defined as
Umax max
zuz(t,z),(2.10)
and the corresponding Reynolds number
Remax
LLUmax
ν.(2.11)
932 A9-6
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
(f)
(a)(b)
(c)(d)
(e)
z/Lζ/L
0
0.2
0.4
0
0.2
–0.2
–0.4
0.4
0
0.2
–0.2
–0.4
0.4
ζ/L
ζ/L
0
0.2
–0.2
–0.4 –4 –2
–1
02
01
4
–1.0 –0.5 0 0.5 1.0
0.4
0.6
0.8
1.0
z/L
0
0.2
0.4
0.6
0.8
1.0
z/L
0255075
t/tfτ/tf
Cuz (τ,ζ)uz/Uf
100 125 150
0.2
0.4
0.6
0.8
1.0
Figure 2. Examples of time–space plots of the vertical velocity uz(t,z)for: (a) DNS data at probe A,Γ=5,
RaL=1.25 ×107,(c) experimental data at probe B,RaL=1.05 ×107and (e) noisy experimental data at
probe B,RaL=6.94 ×106.(b,d,e) The autocorrelation function Cuz(τ, ζ ) for (a,c,e) respectively, cf. (2.12).
The dashed lines indicate the characteristic velocity Uobtained from the autocorrelation as described in § 2.4.
A more sophisticated approach is based on the two-dimensional autocorrelation of
uz(t,z), which we calculate in a discrete manner as
Cuz(τ, ζ ) =1
NMσ2
N,M
i,j
u
z(ti,zj)·u
z(tiτ,zjζ), (2.12)
932 A9-7
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
where u
z(t,z)uz(t,z)−uzzand σis the standard deviation of uz.Notethat,in(2.12),
the fraction in front of the sum normalizes Cuzto values between 1 and 1. Examples of
Cuz(τ, ζ ) are shown in figure 2(b,d,f). The next step is to fit a straight line,
ζ(τ) =Uτ(2.13)
through the points
i
i):Cuzi
i)=max
τ
Cuz(τ, ζi).(2.14)
The slope of this line gives the characteristic velocity U. Additionally, we restrict the
domain of the fit to the relevant points around the origin. How this extraction algorithm
works with experimental and numerical data is shown by several examples in figure 2.At
this point, we can define the Reynolds number based on Uas
Re
L=LU
ν.(2.15)
The advantage of evaluating the velocity in this way is that this method provides very
robust results even in the case of low-quality signals. Figure 2(e) shows an example of
a measurement with a very noisy signal. This is due to weak echo amplitudes caused
by an imperfect acoustic contact between the transducer and the liquid metal or an
insufficient number of reflective tracers in the measurement volume. In this case, the local
velocity values are subject to a high measurement uncertainty, in particular with increasing
distance from the sensor. However, the method proposed here uses almost all the available
information from the measurements, and by taking into account the spatio-temporal nature
of the velocity field, the deficit of poor signal quality can be essentially compensated.
3. Results and discussion
In the following, we will present the results and discuss the evidence, that the relevant
length scale in VC is based on the plate size rather than on the distance between the hot and
cold plates. Furthermore, we are going to explore the flow structures and BLs, and finally
we discuss the new method based on the autocorrelation to determine Reynolds numbers
from UDV data. A summary of the measured quantities can be found in the Appendix,
tables 2 and 3, however, this is only for reference, as all the quantities are presented in
various figures.
3.1. Global flow organization
To provide a deeper insight into the general flow organization, we show in figure 3
instantaneous flow fields at different RaLand aspect ratios Γ. We calculate the full heat
transport vector
Ω(uTκT)/(κΔ/H), (3.1)
and then find regions where its magnitude is greater or equal to the Nusselt number
(|Ω|Nu). This prominently shows the large-scale circulation and captures the major
regions, where heat is transported. Here, we choose the magnitude of the full heat flux
vector as the basis and not just the xcomponent. In this way we can capture the regions
of vertical heat transport close to the tempered walls and also the heat transport between
them. Additionally, this surface is coloured by temperature, to distinguish the hot and cold
streams of the flow. This way of visualization helps to overcome two difficulties: on the one
932 A9-8
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
Temperature
T+
T
(a)(b)
(c)
(e)(d)
Figure 3. Instantaneous heat transport and temperature. The shown superstructures enclose regions, where
the instantaneous heat transport |Ω(t)|Nu, and they are coloured with the local temperature. Panels (ac)
are at RaH=5×104and different aspect ratios Γ=1,2,5, respectively. Panels (d,e,c) are at similar RaL=
5×106,4×106,6.25 ×106, respectively.
hand, vector fields are difficult to render in three dimensions and even streamlines appear
rather cluttered, and on the other hand, temperature isosurfaces appear rather smooth due
to the low Prandtl number and therefore hide the vigorous nature of turbulent convection
in liquid metals.
3.2. The characteristic length scale
There are two prominent length scales in a rectangular box with square hot/cold plates:
the distance Hbetween the plates and the height of the side hot/cold square plates L. One
of the questions we are going to answer in this section is: On which length scale (Hor L)
should the scaling dimensionless quantities like Ra and Re be based?
We denote quantities that are based on Lor Hwith the subscript Lor H, respectively,
and they convert as
RaL=RaHΓ3,NuL=NuHΓand ReL=ReHΓ. (3.2ac)
932 A9-9
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
1
2
5
10
20
RaH
N
uH
NuL
NuH RaH
−0.25
Γ = 1 (DNS)
Γ = 2 (DNS)
Γ = 3 (EXP)
Γ = 5 (DNS)
Γ = 5 (EXP)
0.10
0.15
0.20
1
2
5
10
20
104105106107108
RaL
104105106107108
104105106107108
NuL RaL
−0.25
0.12
0.14
0.16
104105106107108
(a)
(b)
Figure 4. Nusselt number, Nu, vs Rayleigh number, Ra, based on length scales: (a)distanceofhotandcold
plate Hand (b) plate length L.ForΓ=3 the Nusselt number is measured at the hot plate (red pluses) and at the
cold plate (blue pluses). The insets show the compensated Nusselt number based on a Ra1/4-scaling (indicated
by dashed lines).
A comparison of the length scales for the Nusselt number is shown in figure 4(a,b)andthe
data collapse for the length scale L, while they do not do so for length scale H. In general,
one can see this collapse for the Reynolds numbers in figure 5(a,b) and the BL thickness
in figure 6. This means, that the relevant length scale in VC is the plate size, L,rather
than the plate distance, H, for global quantities like Nu and Re. Further below, will analyse
these quantities and figures in detail.
The Nusselt number for the Γ=3 experiments is only 70 % of the value observed in
the numerical simulations. Interestingly, this appears to be a systematic difference in liquid
932 A9-10
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
RaL
104
102
103
104
102
103
104
105106107108
RaL
104105106107108
104105106107108
(a)
(b)
Γ = 1, Rewind (DNS)
Γ = 2, Rewind (DNS)
Γ = 5, Rewind (DNS)
Γ = 5, Probe A (DNS)
Γ = 5, Probe B (DNS)
Noisy experiments
Γ = 5, Probe A (EXP)
Γ = 5, Probe B (EXP)
Γ = 3, Probe A (EXP)
Γ = 3, Probe B (EXP)
0.8
1.0
1.2
1.4
ReL
max RaL
−0.5
104105106107108
0.5
1.0
1.5
R
eL
max
ReL
ReL
RaL
−0.5
Figure 5. (a) Reynolds numbers based on the characteristic velocity Re,and(b) Reynolds numbers based on
the maximal vertical velocity vs the Rayleigh number. The data of red/blue triangles (DNS) and crosses/pluses
(experiments) are obtained by probes at position A/B, respectively. The orange background of data points
indicates noisy measurements, cf. figure 2(e). The dashed lines show the theoretical scaling Re Ra1/2
(Shishkina 2016) and the grey band represents the uncertainty margin of ±20 %. The wind-based Reynolds
numbers Rewind from DNS data are shown by grey solid symbols for different aspect ratios Γ=1(circle),2
(square) and 5 (triangle). The insets show compensated plots.
metal simulations and experiments. Zwirner et al. (2020a) compare DNS and experiments
of VC with liquid sodium as fluid inside a cylinder (Γ=1). There, the absolute
measurement of Nu also shows similar differences. Since the DNS results are in agreement
with the literature (Scheel & Schumacher 2016), we think this systematic deviation is due
to some physical effects not considered within the Navier–Stokes equations, and should be
investigated in the future.
932 A9-11
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
RaL
104
103
10–3
10–2
10–1
100
105106107108
108
106
104
108
106
104
δ/L
2.0
2.5
3.0
3.5
0.20
0.30
0.40
δuRaL
0.25/L
δθRa–0.25/L
δθ
δu
Figure 6. Thermal BL thickness δθ=H/(2Nu)(open symbols, black crosses) and viscous BL thickness δu
based on slope criterion (filled symbols, blue crosses, red pluses) vs the Rayleigh number. The dashed and
dash-dotted lines show the theoretical scaling law δRa0.25 for both BLs. The insets show the compensated
data. Experimental data for Γ=3 (pluses) and Γ=5 the (crosses). DNS data (circles, squares, triangles) as
in figure 4.
As next step, we take a closer look at the local flow organization, which reveals a more
complex picture. Here, we focus on the time-averaged heat transport across the central
vertical cross-section ¯
Ωx(figure 7a) and the respective profile along the vertical direction
(figure 7b). The profiles are obtained from the same simulations as the instantaneous
snapshots in figure 3.Note that the three red curves correspond to figure 3(d,e,c)and
have a similar RaL. Although, the slopes of the profiles close to the boundary are similar,
the behaviour in the bulk is qualitatively different for different aspect ratios. The smaller
aspect ratios, Γ=1 and 2, have a steep rise and a steep fall of the heat transport around its
maximum close to the wall, where the heat transport is approximately 5 times larger than
the average Nu, while for Γ=5, the rise and fall are gentle and the maximum is slightly
above the average heat transport. The shape of the heat transport profile for the Γ=5 case
shares some similarities with the profiles at smaller aspect ratios Γ=1 and 2 for similar
RaH(blue curves). Therefore, one may conclude that the local flow organization is strongly
influenced by the aspect ratio and the profiles show a rather complex behaviour. This is
further supported by the shapes of the time-averaged horizontal profiles of the temperature
and vertical velocity component (figure 8).
In figure 9 we show the vertical profiles of the thermal BL thickness at the heated
plate, for the same cases discussed in the previous paragraph. Here, one notices that,
for the same RaH, the profiles are quite different for different aspect ratios (figure 9a),
while, for similar RaL5×106,the thermal BLs are of similar structure (figure 9b). At
the top (z/L=0) and bottom (z/L=1) one can prominently notice the influence of the
adiabatic boundaries. The thermal BL grows from the bottom to top, first
slowly until z/L0.7 and then faster, being influenced by the adiabatic top
boundary.
932 A9-12
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
z
y
0
2
4
6
z/L
Ωx (x = 0.5 H)y/Nu
Γ = 1 RaH = 5 × 104
Γ = 2 RaH = 5 × 104
Γ = 5 RaH = 5 × 104
Γ = 1 RaH = 5 × 106
Γ = 2 RaH = 5 × 105
00.1 0.2 0.3 0.4 0.5
(a)
(b)
Figure 7. (a) Cross-section of ¯
Ωx(x=0.5H)/Nu for Γ=1andRaH=5×106and (b) profiles of the local
heat transport ¯
Ωx(x=0.5H)averaged in the y-direction as functions of vertical coordinate z. For the red lines
RaL5×106and these data are collected from the same simulations as presented in figure 3(ce). All DNS
data.
0 0.1 0.2 0.3 0.4 0.50 0.1 0.2 0.3 0.4 0.5
x/H
0.1 0.2 0.3 0.4 0.5
x/H
0.1 0.2 0.3 0.4 0.5
(a)(b)
(c)(d)
uzAyz /Uf
T0
T+
T
Ayz /Δ
T0
T+
0
0.2
0.4
0.6
0
0.2
0.4
0.6
Γ = 1 RaH = 5 × 106
Γ = 2 RaH = 5 × 105
Γ = 5 RaH = 5 × 104
Γ = 1 RaH = 5 × 104
Γ = 2 RaH = 5 × 104
Γ = 5 RaH = 5 × 104
Figure 8. Time-averaged horizontal profiles of (a,b) the temperature and (c,d) the vertical velocity
component for (a,c)RaH=5×104and (b,d)RaL5×106. All DNS data.
3.3. The Reynolds number in experiments and DNS
In § 2.4, we introduced two different methods to extract the Reynolds number and
in the following we discuss the robustness of these methods and compare the results
from experiments and DNS. We need to take into account that the beam of the
932 A9-13
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
0.1 0.2 0.3 0.4
(a)(b)
δθ(z)/L
0.2 0.4 0.6
δθ(z)/L
0
0.2
0.4
0.6
0.8
1.0
z
/L
0
0.2
0.4
0.6
0.8
1.0
Figure 9. Profiles of the thermal BL δθ/L=−(2xT)1at the hot plate x=0averagedinthe y-direction as a
function of the vertical coordinate z. Note that this definition of δθis analogous to δθ/L=1/(2Nu),andatthe
plate heat is transported exclusively by conduction;(a)Γ=1,2,5andRaH=5×104and (b)Γ=1,2,5at
similar RaL5×106. The legend is identical to figure 7(b). All DNS data.
(a)(b)
0.1 0.2 0.3 0.4 0.5
xp/H
02468
0.98
1.00
1.02
1.04
100 Ap/H2
ReL/ReL (Ap = 0)
ReL/ReL (xp = 0.25)
0
0.5
1.0
1.5
2.0
Figure 10. Dependence of the normalized Reynolds numbers Remax
L(squares) and Re
L(circles) on (a)the
horizontal probe position and (b) the beam thickness (represented by the cross-sectional area Ap) for the DNS
at RaH=104(filled symbols), RaH=5×105(open symbols) and Γ=5.
Doppler-velocimetry probe has a certain horizontal extent of diameter less than H/4.
Analysing the DNS data, we can mimic this horizontal extent by locally averaging over
a certain area. Due to the Cartesian geometry used in the simulations, we average over a
beam with a square cross-section of area ApandasidelengthofuptoH/4 to account for
the finite extent of the Doppler-velocimetry beam.
To verify the robustness of our measurements of the vertical velocity component we
separately analyse the influence of two major aspects: (i) the horizontal position of the
probe and (ii) the thickness of the ultrasonic beam. The Reynolds number Re
Ldepends
weakly on the horizontal position xpof the probe, while Remax
Ldepends strongly on the
horizontal position of the probe (figure 10a). By changing the averaging area of the square
beam from our numerical probes in the DNS, we find that the extend of the probing beam
has a rather small effect of 2 % on the Reynolds numbers (figure 10b).
The Re
Ldata of regular and noisy experimental data agree very well (figure 5a),
but the Remax
Ldata show systematically lower values for noisy experiments (figure 5b).
This demonstrates the second advantage of the proposed method for extracting the
Reynolds number, besides the weak dependence on the probe position. Note that this weak
dependence on the probe position is quite important, as the positions in the experimental
932 A9-14
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
set-ups for Γ=3 and 5 differ. The experimental Re
L-data collapse nicely for both aspect
ratios (figure 5a).
From the discussion above, we conclude that Re
Land thus Uis an appropriate measure
of the characteristic velocity in the system. To finalize this section, we give a physical
interpretation of U. From the definition of U(2.13)andfigure 2(a,c,e) it becomes clear
that Uquantifies the vertical advection component of the velocity fluctuations along the
line of measurement, or in other words: it is the advection velocity of the fluctuations.
3.4. The scaling relations of global Nu and Re
In Shishkina (2016) the scaling relations for heat and momentum transport in laminar
VC were derived based on similarity solutions for the BL equations. Furthermore, the
validity of this theoretical scaling was shown by DNS data for VC inside a cylindrical
domain of aspect ratio one. In figures 4,5and 6the theoretical scalings for Nu Ra1/4
and Re Ra1/2are indicated by dashed lines. These results agree well with our DNS and
experimental data.
A general remark on the uncertainty of the scaling relations: the Prandtl number in the
experiments might be slightly larger than 0.03 and a good estimate is up to 10 %, i.e. 0.033.
To ensure that this does not affect our conclusions, we estimate how it would influence the
data in the double logarithmic plots. The theoretical scaling from Shishkina (2016) agrees
well with our data, hence we use it as the basis for our analysis. The prefactor for the Nu
and Re scalings in our plots thus depends on the Prandtl number, which, for the slightly
larger Prandtl number of Pr =0.033 results in a factor of 1.024 and 1.049, respectively
(cf. (1.1)). From this one may conclude that the experimental data of Nu and Re might be
shifted by 2.4%and4.9 % towards higher Nu and Re, respectively, which would provide a
even better collapse with the DNS data. Although one should keep in mind that this is by
far not the only source of uncertainty in our data, nevertheless this estimation seems to be
a reasonable explanation of the systematic deviation between experiments and DNS in the
Nu and Re data.
One observation from the experiments is that the scaling exponents of Ra are slightly
larger for the Nusselt number in the case of Γ=5, but slightly smaller for the Reynolds
number compared with the expected scaling exponents from theory and also DNS.
The major challenge in the experiments is the measurement of very small temperature
differences, especially at small Ra. Therefore, the determination of Nu in this range
is highly error prone. From this, we noticed that the uncertainty of Nusselt number is
especially large for low Rayleigh numbers and, together with the fact that the experiments
cover just one order of magnitude in Rayleigh number, this could lead to a large uncertainty
of the scaling exponent.
The deviations of the Reynolds number in the experiments compared with the theoretical
scalings one can interpret as follows. With increasing thermal forcing, not only does the
amplitude of the velocity increase, but also the flow structure changes in the convection
cell. Thus, the location, where the maximum vertical velocity is achieved, moves towards
the tempered plates and away from the beam of the UDV measurements. Although the
measurement of Re
Lshows only a weak dependence on the probe position (especially
for Γ=5), the evaluation of the DNS data for all aspect ratios suggests that the scaling
exponent of Re
Lbecomes slightly smaller as the velocity is measured further away from
the tempered plates. The measuring position of the sensor, however, remains unchanged.
It seems to be also plausible that this effect is more strongly manifested for Γ=3
than for Γ=5. Unfortunately, in the experiments, we cannot change the sensor position
932 A9-15
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
ReL
DNS Γ = 1, ReL
wind
DNS Γ = 2, ReL
wind
DNS Γ = 5, ReL
wind
DNS Γ = 5, ReL
104
103
102
0.2
0.3
0.4
0.5
0.6
0.7
a = δuReL/L
Figure 11. The parameter afrom the relation δu/L=aRe (cf. Prandtl 1905) vs the Reynolds number. All
data from DNS at different aspect ratios Γfor the Reynolds number based on the wind velocity (2.7)andRe
L.
easily, but scenarios are conceivable in which the measurements at a fixed position
underestimate the increase in Re
Lwith Ra and therefore underestimates the scaling
exponent.
The thickness of the thermal and viscous BLs can be obtained from δθ/L=1/(2NuL)
and δu/L=a/ReL(Prandtl 1905), respectively. In figure 6 we show the respective BLs,
and, while the thermal BL thickness is obtained straightforwardly, the viscous BL needs
an estimation of the parameter awhich is widely accepted in case of RBC to be 0.482 for
a cylindrical cell of unit aspect ratio (Grossmann & Lohse 2002), but in general dependent
on Γ. From the DNS data, using Re
L, we can estimate this parameter for Γ=5tobe
a0.38 and then apply this value of ato the experimental data for both aspect ratios
Γ=3 and 5. Thus, we calculate the average ratio of Re
L1/2and δu, which is obtained
by (2.8). The estimate of aworks well for both aspect ratios Γ=3and5usedbythe
experiments (figure 6). This gives an estimate that the viscous BL at RaL107has a
thickness of approximately 1 mm, while the distance between the plates is H=40 mm
(Γ=5).
This procedure can also be applied to the different aspect ratios, and by using Rewind
instead of Reas the Reynolds number. A full comparison is shown in figure 11.Atfirst
glance these data show no conclusive trend, however,one needs to take into account the
above discussed uncertainty margin of ±20 % for the Reynolds numbers. With this in
mind, figure 11 is similar to the compensated plots shown as insets of figure 5(a,b)and
these data show similar scatter.
4. Conclusions
We studied VC of a liquid metal with low Prandtl number (Pr 0.03), more
precisely, the eutectic alloy GaInSn, using the complementary results from experimental
measurements and DNS. The agreement between DNS and experiments is reasonable and
deviations are within the acceptable range. Furthermore, we extensively discussed these
deviations.
932 A9-16
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
Our study shows quantitatively, with respect to the global heat and momentum transport,
and qualitatively by the instantaneous heat transport superstructures (figure 3), that the
plate size L, rather than the distance between the plates H, is the relevant length scale in
VC.
Furthermore, we found that the bulk of the flow contributes only little to the global
heat transport. The heat is mainly transported by horizontal layers close to the top/bottom
boundary of the convection cell, while the bulk heat transport is far below the mean heat
transport or even slightly negative. These heat transporting layers have a thickness in the
range of approximately 0.1Lto 0.25 Leach.
Most importantly, we reported a novel method to extract the wind velocity
from experimental data, in particular from time–space plots of Doppler-velocimetry
measurements, using the two-dimensional autocorrelation defined in (2.12). With the
support from the DNS, we show that the Reynolds numbers obtained by our novel
method are very similar to the wind-based Reynolds numbers, while they are only weakly
dependent on the horizontal position of the probe. Additionally, we showed that even for
noisy experimental data our method delivers reasonable Reynolds numbers. Note that
this cannot be achieved in direct measurements of the velocity, since they are strongly
influenced by the probe position and lead to systematically smaller Reynolds numbers.
Further, we report a value of a0.38 to calculate the viscous BL thickness via δu=
aL/ReLfor VC.
Although the results are obtained only at a single value of Pr 0.03, one would expect
similar behaviour for fluids with Pr 1 in general. Altogether, we gained important
knowledge on the turbulent wind extraction, in particular in liquid metals. This is
important for accurate calculations of the Reynolds number from experimental data and
also because the scaling theories (Grossmann & Lohse 2000; Shishkina 2016) are built
upon the notion of wind. One challenge for further theoretical studies is that, in VC, unlike
RBC, there is a non-closed term in the exact relation between the global kinetic energy
dissipation rate and the vertical convective heat transport (Ng et al. 2015; Zwirner &
Shishkina 2018). Also, for the future, a deeper investigation of the local flow organization
is necessary, because it, unlike the global quantities, strongly depends on the aspect
ratio.
Acknowledgements. The authors acknowledge the Leibniz Supercomputing Centre (LRZ) for providing
computing time.
Funding. This work is funded by the Deutsche Forschungsgemeinschaft (DFG) under the grants Sh405/7
(Priority Programme SPP 1881 ‘Turbulent Superstructures’), VO 2331/1-1 and VO 2331/4-1.
Declaration of interests. The authors report no conflict of interest.
Author ORCIDs.
Lukas Zwirner https://orcid.org/0000-0002-8805-7292;
Sanjay Singh https://orcid.org/0000-0002-5305-7524;
Sven Eckert https://orcid.org/0000-0003-1639-5417;
Tobias Vogt https://orcid.org/0000-0002-0022-5758;
Olga Shishkina https://orcid.org/0000-0002-6773-6464.
Appendix
The following tables contain relevant experimental and numerical data.
932 A9-17
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
DNS resolution
L/HRa
HNxNyNzTavg/tfNδuNδθ
15×103114 114 114 25 2 7 41
11×104114 114 114 251 6 34
15×104114 114 114 25 2 4 22
11×105170 170 170 292 4 27
15×105226 226 226 335 4 25
15×106338 338 338 161 5 31
11×107450 450 450 186 5 35
11×108674 674 674 66 7 47
25×103114 142 142 326 6 41
21×104114 142 142 328 5 35
25×104114 170 17 0 5 31 5 26
21×105170 226 226 332 5 26
25×105226 338 338 205 5 26
55×103114 226 226 286 6 48
51×104114 226 226 286 6 43
55×104170 338 338 299 7 49
51×105226 450 450 320 6 49
52×105226 450 450 162 6 40
55×105226 450 450 219 5 30
Table 1. Simulation parameters: the aspect ratio (L/H), Rayleigh number (RaH), grid size (Nx×Ny×Nz),
averaging time (Tavg) and number of points in the viscous (Nδu)andthermalBLs(Nδθ). The Prandtl number is
fixed to Pr =0.03.
DNS
L/HRa
HRaLNuHRemax
LRe
LRewind δu/Hδθ/H
15×1035×1031.42 144 0.0467 0.352
11×1041×1041.75 218 0.0401 0.286
15×1045×1042.76 481 0.0250 0.181
11×1051×1053.30 659 0.0198 0.151
15×1055×1055.05 1407 0.0122 0.099
15×1065×1069.05 4391 0.0062 0.055
11×1071×10710.82 6294 0.0052 0.046
11×1081×10819.82 19 874 0.0027 0.025
25×1034×1041.42 380 0.0395 0.352
21×1048×1041.69 553 0.0333 0.296
25×1044×1052.70 1437 0.0238 0.185
21×1058×1053.35 2237 0.0212 0.149
25×1054×1064.85 4662 0.0141 0.103
55×1036.25 ×1051.20 1257 1908 1281 0.0441 0.417
51×1041.25 ×1061.35 1851 2525 1858 0.0377 0.371
55×1046.25 ×1061.99 4270 5638 4339 0.0255 0.252
51×1051.25 ×1072.44 6305 7849 6373 0.0217 0.205
52×1052.50 ×1073.03 8557 10 858 9475 0.0186 0.165
55×1056.25 ×1074.02 12 686 14 974 16 522 0.0152 0.124
Table 2. Summary of the measured quantities in the DNS for Pr =0.03: the aspect ratio (L/H), the Rayleigh
number (Ra) based on Land H, the Nusselt number (Nu), the various Reynolds numbers (Re)denedby(2.11),
(2.15)and(2.7) and the viscous and thermal BL thicknesses (δu,δθ). Note that Re
Land Remax
Lare the averages
of probes Aand B.
932 A9-18
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
Experiments
L/HRa
HRaLNuHRemax
LRe
L
31.30 ×1043.60 ×105— 2911 2836
34.51 ×1041.25 ×106—37453542
39.29 ×1042.59 ×106— 4661 4355
32.31 ×1056.42 ×106— 7405 5488
34.72 ×1051.31 ×107—6817 6758
37.09 ×1051.97 ×107— 7489 7904
39.49 ×1052.64 ×107— 8007 8590
31.68 ×1064.68 ×107—8312 9148
31.19 ×1063.32 ×107—8755 9948
31.44 ×1064.01 ×107 9101 10 850
31.93 ×1065.36 ×107 9401 11 025
53.30 ×1044.13 ×106— 2740 3982
55.53 ×1046.91 ×1061.63 2898 4889
55.55 ×1046.94 ×1061.54 1924 5471
55.80 ×1047.26 ×1061.65 3625 5197
56.55 ×1048.19 ×1061.80 3939 5375
56.59 ×1048.24 ×1061.72 2199 5610
57.11 ×1048.89 ×10 61.69 —
58.40 ×1041.05 ×107— 4468 6236
59.97 ×1041.25 ×1072.02 4011 6419
59.99 ×1041.25 ×1071.93 2425 6015
51.46 ×1051.83 ×1072 .41 3189 7557
51.60 ×1052.00 ×1072.59 4443 7545
52.04 ×1052.55 ×1072.88 6640 8468
52.06 ×1052.57 ×1072.82 4370 8358
52.42 ×1053.03 ×1073.12 7075 9129
52.58 ×1053.23 ×1073.21 6647 9092
53.29 ×1054.12 ×1073.51 7754 9848
53.31 ×1054.14 ×10 73.41 6705 19 784
Table 3. Summary of the measured quantities in the experiments: the Rayleigh number (Ra) based on Land
H, the Nusselt number (Nu), and the two Reynolds numbers (Re)denedby(2.11)and(2.15). Note that Re
L
and Remax
Lare the averages of probes Aand B.Rowsmarkedbyinvolve noisy UDV data. The experiments are
conducted at Pr 0.03 and the aspect ratios L/H=3 and 5.
932 A9-19
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
Experiments Γ=3
RaHRaLNuH(hot) NuH(cold) RaHRaLNuH(hot) NuH(cold)
5.31 ×1041.43 ×1061.74 — 2.36 ×1056.38 ×106—2.31
5.41 ×1041.46 ×106—1.412.41 ×1056.51 ×1062.28 —
5.76 ×1041.55 ×106—1.322.44 ×1056.60 ×106—2.29
6.02 ×1041.62 ×1061.65 — 2.50 ×1056.75 ×106—2.31
6.11 ×1041.65 ×1061.82 — 2.58 ×1056.98 ×1062.32 —
6.15 ×1041.66 ×106—1.472.60 ×1057.02 ×1062.33 —
6.65 ×1041.79 ×1061.76 — 2.66 ×1057.18 ×1062.39 —
6.96 ×1041.88 ×106—1.492.74 ×1057.39 ×106—2.36
7.43 ×1042.01 ×106—1.592.81 ×1057.58 ×1062.37 —
7.66 ×1042.07 ×106—1.523.00 ×1058.11 ×106—2.39
7.94 ×1042.14 ×1061.81 — 3.09 ×1058.35 ×1062.43 —
8.32 ×1042.25 ×106—1.583.21 ×1058.67 ×1062.47 —
8.52 ×1042.30 ×106—1.683.53 ×1059.54 ×106—2.48
8.63 ×1042.33 ×106—1.723.62 ×1059.79 ×1062.50 —
9.22 ×1042.49 ×106—1.773.82 ×1051.03 ×1072.54 —
9.26 ×1042.50 ×106—1.754.05 ×1051.09 ×107—2.56
9.27 ×1042.50 ×1061.86 — 4.13 ×1051.12 ×1072.63 —
9.51 ×1042.57 ×106—1.704.56 ×1051.23 ×107—2.62
9.64 ×1042.60 ×106—1.814.74 ×1051.28 ×1072.69 —
9.88 ×1042.67 ×1061.57 — 4.98 ×1051.34 ×107—2.68
1.04 ×1052.82 ×106—1.835.09 ×1051.37 ×107—2.68
1.06 ×1052.86 ×1061.91 — 5.22 ×1051.41 ×1072.76 —
1.07 ×1052.89 ×1061.88 — 7.42 ×1052.00 ×107—2.90
1.07 ×1052.90 ×1061.96 — 7.71 ×1052.08 ×107—2.92
1.08 ×1052.91 ×1061.79 — 7.77 ×1052.10 ×1073.00 —
1.08 ×1052.92 ×106—1.781.00 ×1062.71 ×107—3.08
1.19 ×1053.23 ×1061.91 — 1.03 ×1062.78 ×1073.19 —
1.20 ×1053.23 ×106—1.871.04 ×1062.82 ×107—3.08
1.32 ×1053.57 ×1061.98 — 1.26 ×1063.41 ×107—3.23
1.43 ×1053.86 ×1062.03 — 1.29 ×1063.50 ×1073.34 —
1.45 ×1053.90 ×106—1.981.32 ×1063.55 ×107—3.23
1.56 ×1054.22 ×1062.06 — 1.52 ×1064.12 ×107—3.34
1.58 ×1054.25 ×1062.13 — 1.56 ×1064.21 ×1073.47 —
1.64 ×1054.42 ×106—2.171.58 ×1064.27 ×107—3.34
1.68 ×1054.53 ×1062.10 — 1.79 ×1064.82 ×107—3.44
1.68 ×1054.53 ×1062.11 — 1.85 ×1065.00 ×107—3.46
1.79 ×1054.85 ×1062.14 — 1.85 ×1065.01 ×107—3.46
1.97 ×1055.33 ×106—2.191.86 ×1065.02 ×1073.59 —
2.08 ×1055.61 ×1062.27 — 1.89 ×1065.12 ×1073.59 —
2.11 ×1055.69 ×1062.23 — 1.97 ×1065.31 ×107—3.49
2.22 ×1056.00 ×106—2.231.97 ×1065.32 ×107—3.50
2.30 ×1056.20 ×1062.27 — 2.02 ×1065.47 ×107—3.51
Table 4. Experiments with Γ=3: the Rayleigh number (Ra) based on Land H, the Nusselt number (Nu)
based on Hmeasured at the hot or cold plate.
932 A9-20
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
REFERENCES
AHLERS,G.,BROWN,E.&NIKOLAENKO, A. 2006 The search for slow transients, and the effect of imperfect
vertical alignment, in turbulent Rayleigh–Bénard convection. J. Fluid Mech. 557, 347–367.
AHLERS,G.,GROSSMANN,S.&LOHSE, D. 2009 Heat transfer and large scale dynamics in turbulent
Rayleigh–Bénard convection. Rev. Mod. Phys. 81, 503–537.
AUBERT,J.,BRITO,D.,NATAF,H.-C.,CARDIN,P.&MASSON, J.-P. 2001 A systematic experimental study
of rapidly rotating spherical convection in water and liquid gallium. Phys. Earth Planet. Inter. 128 (1–4),
51–74.
AURNOU, J.M., BERTIN,V.,GRANNAN, A.M., HORN,S.&VOGT, T. 2018 Rotating thermal convection in
liquid gallium: multi-modal flow, absent steady columns. J. Fluid Mech. 846, 846–876.
BATCH EL OR, G.K. 1954 Heat transfer by free convection across a closed cavity between vertical boundaries
at different temperatures. Q. Appl. Maths 12, 209–233.
BEJAN, A. 1980 A synthesis of analytical results for natural convection heat transfer across rectangular
enclosures. Intl J. Heat Mass Transfer 23 (5), 723–726.
BEJAN, A. 1985 Mass and heat transfer by natural convection in a vertical cavity. Intl J. Heat Fluid Flow 6(3),
149–159.
BEJAN,A.2013Convection Heat Transfer, 4th edn. John Wiley & Sons.
BELMONTE,A.,TILGNER,A.&LIBCHABER, A. 1995 Turbulence and internal waves in side-heated
convection. Phys. Rev. E51, 5681–5687.
BODENSCHATZ, E., PESCH,W.&AHLERS, G. 2000 Recent developments in Rayleigh–Bénard convection.
Annu. Rev. Fluid Mech. 32, 709–778.
BRAUNSFURTH, M.G., SKELDON, A.C., JUEL,A.,MULLIN,T.&RILEY, D.S. 1997 Free convection in
liquid gallium. J. Fluid Mech. 342, 295–314.
BRITO,D.,NATAF,H.-C.,CARDIN,P.,AUBERT,J.&MASSON, J.-P. 2001 Ultrasonic Doppler velocimetry
in liquid gallium. Exp. Fluids 31 (6), 653–663.
CHEN,Y.-M.&PEARLSTEIN, A.J. 1989 Stability of free-convection flows of variable-viscosity fluids in
vertical and inclined slots. J. Fluid Mech. 198, 513–541.
CHILLÀ,F.,RASTELLO,M.,CHAUMAT,S.&CASTAING, B. 2004 Long relaxation times and tilt sensitivity
in Rayleigh–Bénard turbulence. Eur. Phys. J. B40 (2), 223–227.
CHILLÀ,F.&SCHUMACHER, J. 2012 New perspectives in turbulent Rayleigh–Bénard convection. Eur. Phys.
J. E35, 58.
CHURCHILL,S.W.&CHU, H.H.S. 1975 Correlating equations for laminar and turbulent free convection from
a vertical plate. Intl J. Heat Mass Transfer 18, 1323–1329.
DANIELS, K.E., WIENER,R.J.&BODENSCHATZ, E. 2003 Localized transverse bursts in inclined layer
convection. Phys. Rev. Lett. 91, 114501.
ECKERT,S.&GERBETH, G. 2002 Velocity measurements in liquid sodium by means of ultrasound Doppler
velocimetry. Exp. Fluids 32 (5), 542–546.
FUJII,T.,TAKEUCHI,M.,FUJII,M.,SUZAKI,K.&UEHAR, H. 1970 Experiments on natural-convection
heat transfer from the outer surface of a vertical cylinder to liquids. Intl J. Heat Mass Transfer 13, 753–770.
GILL, A.E. 1966 The boundary-layer regime for convection in a rectangular cavity. J. Fluid Mech. 26, 515–536.
GILLET,N.,BRITO,D.,JAU LT,D.&NATAF, H.C. 2007 Experimental and numerical studies of convection
in a rapidly rotating spherical shell. J. Fluid Mech. 580, 83–121.
GRAEBEL, W.P. 1981 The influence of Prandtl number on free convection in a rectangular cavity. Intl J. Heat
Mass Transfer 24, 125–131.
GROSSMANN,S.&LOHSE, D. 2000 Scaling in thermal convection: a unifying theory. J. Fluid Mech. 407,
27–56.
GROSSMANN,S.&LOHSE, D. 2002 Prandtl and Rayleigh number dependence of the Reynolds number in
turbulent thermal convection. Phys. Rev. E66, 016305.
KELLEY,D.H.&WEIER, T. 2018 Fluid mechanics of liquid metal batteries. Appl. Mech. Rev. 70 (2), 020801.
KING, E.M. & AURNOU, J.M 2013 Turbulent convection in liquid metal with and without rotation. Proc. Natl
Acad. Sci. USA 110, 6688–6693.
KIS,P.&HERWIG, H. 2012 The near wall physics and wall functions for turbulent natural convection. Intl
J. Heat Mass Transfer 55, 2625–2635.
KOOIJ, G.L., BOTCHEV, M.A., FREDERIX, E.M.A., GEURTS, B.J., HORN,S.,LOHSE,D.,VA N D E R
POEL, E.P., SHISHKINA,O.,STEVENS, R.J.A.M. & VERZICCO, R. 2018 Comparison of computational
codes for direct numerical simulations of turbulent Rayleigh–Bénard convection. Comput. Fluids 166, 1–8.
KOSTER, J.N., SEIDEL,T.&DEREBAIL, R. 1997 A radioscopic technique to study convective fluid dynamics
in opaque liquid metals. J. Fluid Mech. 343, 29–41.
932 A9-21
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
L. Zwirner and others
NATAF,H.-C.,ALBOUSSIERE,T.,BRITO,D.,CARDIN,P.,GAGNIÈRE,N.,JAULT,D.&SCHMITT,D.
2008 Rapidly rotating spherical Couette flow in a dipolar magnetic field: an experimental study of the
mean axisymmetric flow. Phys. Earth Planet. Inter. 170 (1–2), 60–72.
NG, C.S., OOI,A.,LOHSE,D.&CHUNG, D. 2015 Vertical natural convection: application of the unifying
theory of thermal convection. J. Fluid Mech. 764, 349–361.
OSTRACH, S. 1972 Natural convection in enclosures. Adv. Heat Transfer 8, 161–227.
PALLARES,J.,VERNET,A.,FERRE,J.A.&GRAU, F.X. 2010 Turbulent large-scale structures in natural
convection vertical channel flow. Intl J. Heat Mass Transfer 53 (19), 4168–4175.
PAOLUCCI, S. 1990 Direct numerical simulation of two-dimensional turbulent natural convection in an
enclosed cavity. J. Fluid Mech. 215, 229–262.
PAOLUCCI, S. 1994 The differentially heated cavity. Sadhana 19 (5), 619–647.
PLEVACHUK,Y.,SKLYARCHUK,V.,ECKERT,S.,GERBETH,G.&NOVAKOVI C, R. 2014 Thermophysical
properties of the liquid Ga–In–Sn eutectic alloy. J. Chem. Engng Data 59 (3), 757–763.
PRANDTL, L. 1905 Über Flüssigkeitsbewegung bei sehr kleiner Reibung. In Verhandlungen des III. Int. Math.
Kongr., Heidelberg, 1904, pp. 484–491. Teubner.
SAVILLE,D.A.&CHURCHILL, S.W. 1969 Free Convection at low Prandtl numbers in laminar boundary
layers. Ind. Engng Chem. Fund. 8(2), 329–332.
SCHEEL,J.D.&SCHUMACHER, J. 2016 Global and local statistics in turbulent convection at low Prandtl
numbers. J. Fluid Mech. 802, 147–173.
SCHUMACHER,J.,BANDARU,V.,PANDEY,A.&SCHEEL, J.D. 2016 Transitional boundary layers in
low-Prandtl-number convection. Phys. Rev. Fluids 1, 084402.
SHISHKINA, O. 2016 Momentum and heat transport scalings in laminar vertical convection. Phys. Rev. E93,
051102.
SHISHKINA, O. 2021 Rayleigh–Bénard convection: the container shape matters. Phys. Rev. Fluids 6, 090502.
SHISHKINA,O.&HORN, S. 2016 Thermal convection in inclined cylindrical containers. J. Fluid Mech. 790,
R3.
SUN,C.,XI, H.-D. & XIA, K.-Q. 2005 Azimuthal symmetry, flow dynamics, and heat transport in turbulent
thermal convection in a cylinder with an aspect ratio of 0.5. Phys. Rev. Lett. 95, 074502.
TAKEDA,Y.2012Ultrasonic Doppler Velocity Profiler for Fluid Flow.Springer.
TSUJI,T.&NAGANO , Y. 1988 Characteristics of a turbulent natural convection boundary layer along a vertical
flat plate. Intl J. Heat Mass Transfer 31 (8), 1723–1734.
VERSTEEGH, T.A.M. & NIEUWSTADT, F.T.M. 1999 A direct numerical simulation of natural convection
between two infinite vertical differentially heated walls scaling laws and wall functions. Intl J. Heat Mass
Transfer 42, 3673–3693.
VOGT,T.,HORN,S.&AURNOU, J.M. 2021aOscillatory thermal–inertial flows in liquid metal rotating
convection. J. Fluid Mech. 911,A5.
VOGT,T.,HORN,S.,GRANNAN,A.M.&AURNOU,J.M.2018aJump rope vortex in liquid metal convection.
Proc. Natl Acad. Sci. USA 115, 12674–12679.
VOGT,T.,ISHIMI,W.,YANAGISAWA,T.,TASAKA,Y.,SAKURABA,A.&ECKERT,S.2018bTransition
between quasi-two-dimensional and three-dimensional Rayleigh–Bénard convection in a horizontal
magnetic field. Phys. Rev. Fluids 3, 013503.
VOGT,T.,RÄBIGER,D.&ECKERT, S. 2014 Inertial wave dynamics in a rotating liquid metal. J. Fluid Mech.
753, 472–498.
VOGT,T.,YANG,J.-C.,SCHINDLER,F.&ECKERT, S. 2021bFree-fall velocities and heat transport
enhancement in liquid metal magneto-convection. J. Fluid Mech. 915, A68.
WANG,Q.,LIU,H.-R.,VERZICCO,R.,SHISHKINA,O.&LOHSE, D. 2021 Regime transitions in thermally
driven high-Rayleigh number vertical convection. J. Fluid Mech. 917,A6.
WEI, T. 2020 Scaling of Reynolds stresses in a differentially heated vertical channel. Phys. Rev. Fluids 4,
051501.
WEISS,S.&AHLERS, G. 2013 Effect of tilting on turbulent convection: cylindrical samples with aspect ratio
γ=0.50. J. Fluid Mech. 715, 314–334.
YANG, J.C., VOGT,T.&ECKERT, S. 2021 Transition from steady to oscillating convection rolls in
Rayleigh–Bénard convection under the influence of a horizontal magnetic field. Phys. Rev. Fluids 6(2),
023502.
ZHOU,Q.&XIA, K.-Q. 2010 Measured instantaneous viscous boundary layer in turbulent Rayleigh–Bénard
convection. Phys. Rev. Lett. 104, 1043 01.
ZÜRNER,T.,SCHINDLER,F.,VOGT,T.,ECKERT,S.&SCHUMACHER, J. 2019 Combined measurement of
velocity and temperature in liquid metal convection. J. Fluid Mech. 876, 1108–1128.
932 A9-22
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
Vertical convection of liquid metals
ZÜRNER,T.,SCHINDLER,F.,VOGT,T.,ECKERT,S.&SCHUMACHER, J. 2020 Flow regimes of
Rayleigh–Bénard convection in a vertical magnetic field. J. Fluid Mech. 894,A21.
ZWIRNER, L., KHALILOV,R.,KOLESNICHENKO,I.,MAMYKIN,A.,MANDRYKIN,S.,PAVLINOV,A.,
SHESTAKOV,A.,TEIMURAZOV,A.,FRICK,P.&SHISHKINA, O. 2020aThe influence of the cell
inclination on the heat transport and large-scale circulation in liquid metal convection. J. Fluid Mech. 884,
A18.
ZWIRNER,L.&SHISHKINA, O. 2018 Confined inclined thermal convection in low-Prandtl-number fluids.
J. Fluid Mech. 850, 984–1008.
ZWIRNER, L., TILGNER,A.&SHISHKINA, O. 2020bElliptical instability and multiple-roll flow modes of
the large-scale circulation in confined turbulent Rayleigh–Bénard convection. Phys. Rev. Lett. 125, 054502.
932 A9-23
Downloaded from https://www.cambridge.org/core. IP address: 181.215.6.255, on 03 Dec 2021 at 02:36:15, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2021.977
... This setup makes DHVC an invaluable tool in investigating thermal patterns, fluid motion, and their underlying interaction, thereby attracting significant attention from experimental, theoretical, and numerical researchers for decades. [1][2][3][4] The scientific inquiry into DHVC began with Batchelor's pioneering work, 1 which first addressed the steady-state heat transport across the double-glazed windows. Subsequent studies by Elder 5,6 explored laminar and turbulent free convection in vertical slots, providing key insights into flow dynamics and heat transport mechanisms. ...
... This study has extended our comprehension of DHVC dynamics by confirming scaling relations of Nu $ Ra 0:25 and Re $ Ra 0:37 in laminar DHVCs, particularly within a Ra range of 10 8 -10 10 . Recently, Zwirner et al. 2 numerically and experimentally explored turbulent thermal convection of liquid metal with small Pr % 0:03 in a boxshaped container with various aspect ratios. They developed a method to accurately derive the wind-based Re values from experimental Doppler-velocimetry data, utilizing 2D autocorrelation of vertical velocity time-space data. ...
Article
Full-text available
This study explores the non-Oberbeck–Boussinesq (NOB) effects on hydrodynamics and heat transport in two-dimensional glycerol-filled differentially heated vertical cavity (DHVC). The simulations span Rayleigh numbers (Ra) from 2 × 10 3 to 5 × 10 9 and temperature difference ( Δ θ ̃) up to 50 K at a Prandtl number (Pr) of 2547. We showed the emergence of stratified flow structures, delineated the NOB effects on temperature distribution symmetry, and analyzed the scaling behaviors of the Nusselt number (Nu), Reynolds number (Re), and thermal boundary layer (BL) thicknesses ( λ ¯ h θ and λ ¯ c θ) against Ra. For R a ≥ 3 × 10 5, the stratification number (S) shows reduced sensitivity to changes in Ra, stabilizing around 0.5. Additionally, the center temperature ( θ cen) appears to be unaffected by Ra and increases linearly with Δ θ ̃ for R a > 10 6, satisfying θ cen ≈ 2.99 × 10 − 3 K − 1 Δ θ ̃. Our results also revealed that N u ∼ R a γ Nu and R e ∼ R a γ Re with 0.2649 ≤ γ Nu ≤ 0.2654 and 0.3633 ≤ γ Re ≤ 0.3643, respectively, where γ Nu and γ Re exhibit a monotonic decrease as NOB effects intensify. For all investigated Ra values, N u NOB / N u OB < 1 and R e NOB / R e OB > 1 hold consistently, with deviations from OB predictions capped at 6.38% and 2.63% for R a ≥ 10 8, respectively. The analysis of thermal BL thickness reveals distinct scaling behaviors, characterized by λ ¯ h , c θ ∼ R a γ λ ¯ h , c, with scaling exponents ranging from −0.2690 to −0.2669 for both OB and NOB scenarios. Notably, it reveals a divergence from water-based DHVC trends, showing linear decreases in the hot wall's scaling exponent and increases for the cold wall.
... With this configuration, research is still concerned to identify proper scaling laws from the setup's characteristic Rayleigh number Ra to the heat transport and boundary layer dimensions. In Zwirner et al. [48] it was shown by simulations and experiments, that the plate size is a much more dominating factor than the aspect ratio in contrast to the horizontal setup. ...
... Interestingly, for Ra below 1 × 10 5 the velocity profile in the viscous boundary layer can be detected. According to Zwirner et al. [48] the boundary layer should be in the range of 2.5 to 1.5 mm, of which ULM is here shown to detect. For increased Ra the boundary layer is further pushed to the wall and the velocity gradient increases. ...
Preprint
Full-text available
Convection of liquid metals drives large natural processes and is important in technical processes. Model experiments are conducted for research purposes where simulations are expensive and the clarification of open questions requires novel flow mapping methods with an increased spatial resolution. In this work, the method of Ultrasound Localization Microscopy (ULM) is investigated for this purpose. Known from microvasculature imaging, this method provides an increased spatial resolution beyond the diffraction limit. Its applicability in liquid metal flows is promising, however the realization and reliability is challenging, as artificial scattering particles or microbubbles cannot be utilized. To solve this issue an approach using nonlinear adaptive beamforming is proposed. This allowed the reliable tracking of particles of which super-resolved flow maps can be deduced. Furthermore, the application in fluid physics requires quantified results. Therefore, an uncertainty quantification model based on the spatial resolution, velocity gradient and measurement parameters is proposed, which allows to estimate the flow maps validity under experimental conditions. The proposed method is demonstrated in magnetohydrodynamic convection experiments. In some occasions, ULM was able to measure velocity vectors within the boundary layer of the flow, which will help for future in-depth flow studies. Furthermore, the proposed uncertainty model of ULM is of generic use in other applications.
... 16 Since then, VC has become a prevalent topic in experimental, theoretical, and numerical studies. [17][18][19][20][21][22][23][24][25][26][27] The VC system is usually applied to sidewall heating or cooling situations, for example, ice melting and reactor cooling. 28,29 The study of VC systems can also help us understand more complex thermal convection phenomena. ...
Article
Full-text available
We employ the direct numerical simulation to study the heat transfer behavior and flow structures in a vertical convection system with rough sidewalls. The parameters are chosen with Rayleigh number spanning the range of 1×108≤Ra≤3×1010 and Prandtl number fixed at 1.0. The results reveal that the impact of rough walls on the Nusselt number Nu and the Reynolds number Re is influenced by the height of the rough element h. When h is not sufficiently high, the roughness impedes the flows within the boundary layer and traps massive heat between rough elements, and both Nu and Re are lower than those in the smooth-wall case. However, the extent of the Nu and Re reduction regimes decreases as Ra increases. For sufficiently large Ra, the reduction regime for both Nu and Re may vanish, and roughness breaks up the limitation of the thermal boundary layer and facilitates the eruption of thermal plumes from roughness tips, resulting in the enhancement of both Nu and Re. Based on these results, the critical heights hc for Nu and hcr for Re are identified. Both exhibit similar scaling behavior with Ra, with hc consistently being larger than hcr for the same value of Ra.
... However, it is recognised widely that the mean velocity profile in wall shear flows cannot be applied to describe vertical convection (VC), due to the effect of buoyancy (Ke et al. 2020). The VC is the buoyancy-driven flow along a vertical heated wall, and it usually includes three configurations (Hölling & Herwig 2005;Howland et al. 2022): (a) a single heated vertical plate immersed in an ambient fluid (Tsuji & Nagano 1988a,b;Wells & Worster 2008;Ke et al. 2020), abbreviated as 'plate VC' hereafter; (b) a closed cavity with two opposite sidewalls heated at different temperatures (Shishkina 2016;Jiang et al. 2019;Wang et al. 2021;Zwirner et al. 2022), abbreviated as 'cavity VC' hereafter; (c) a vertical infinite channel with periodic boundary conditions in the wall-parallel directions (Pallares et al. 2010;Ng et al. 2013Ng et al. , 2015Howland et al. 2022;Ching 2023), abbreviated as 'channel VC' hereafter. This study focuses on the last configuration of VC, as shown in figure 1(a). ...
Article
In this study, mean velocity and temperature profiles for turbulent vertical convection (VC) confined in an infinite channel are investigated theoretically. The analysis starts from the governing equations of the thermal flow, with Reynolds shear stress and turbulent heat flux closed by the mixing length theory. Employing a three-sublayer description of the mean fields, the mean velocity and temperature profiles are found to be linear laws near the channel wall (viscosity-dominated sublayer), and they follow power laws close to the channel centre (turbulence-dominated sublayer). The characteristic scales of velocity, temperature and length in the present profiles arise naturally from the system normalisation, rather than from scaling analyses, thus ensuring a sound mathematical description. The derived profiles are verified fully via various literature data available in the classical regime; further, they are compared with the reported profiles, and the results indicate that the present profiles are the only ones with the ability to interpret data accurately from different sources, demonstrating much better versatility. Meanwhile, we provide analytical arguments showing that in the ultimate regime, the mean profiles in VC may remain in power laws, rather than the log laws inferred by analogy with Rayleigh–Bénard convection (RBC) systems. The power profiles recognised in this study are induced by the effect of buoyancy, which is in parallel with the mean flow in VC and contributes to the streamwise momentum transport, whereas in RBC systems, buoyancy is perpendicular to the mean flow, and does not influence the streamwise momentum transport, resulting in log profiles, being similar to the case of wall shear flows.
Article
The three-dimensional stability of two-dimensional natural convection flows in a heated, square enclosure inclined to the horizontal is investigated numerically. First, the computational procedure is validated by comparison of base flow solutions to results reported in literature across a range of inclinations. A bi-global linear stability analysis is then conducted to investigate the stability of these two-dimensional base flows to infinitesimal three-dimensional perturbations, and the effect that buoyancy forces (defined by a buoyancy number RNR_N ) and enclosure inclination θ\theta have on these stability characteristics. The flow is first observed to become three-dimensionally unstable at buoyancy number RN=213.8R_N = 213.8 when θ\theta is 180180^\circ ; this increases to RN=2.54×104R_N = 2.54 \times 10^4 at inclination θ=58\theta =58^\circ . It is found that the two-dimensional base flow is more unstable to three-dimensional perturbations with the critical RNR_N corresponding to three-dimensional instability being significantly lower than its two-dimensional counterpart across all considered inclinations except 83θ8883^\circ \leq \theta \leq 88^\circ , where the most unstable mode is a two-dimensional oscillatory mode that develops in the boundary layers along the conducting walls. Eight different leading three-dimensional instability modes are identified, with inclinations 58θ<8858^\circ \leq \theta < 88^\circ transitioning through an oscillatory mode, and inclinations 88θ18088^\circ \leq \theta \leq 180^\circ transitioning through a stationary mode. The characteristics of the primary instability modes corresponding to inclinations 88θ17988^\circ \leq \theta \leq 179^\circ indicate the presence of a Taylor–Görtler instability.
Article
Thermally driven turbulent flows are omnipresent in nature and technology. A good understanding of the physical principles governing such flows is key for numerous problems in oceanography, climatology, geophysics, and astrophysics for problems involving energy conversion, heating and cooling of buildings and rooms, and process technology. In the physics community, the most popular system to study wall-bounded thermally driven turbulence has been Rayleigh-Bénard flow, i.e., the flow in a box heated from below and cooled from above. The dimensionless control parameters are the Rayleigh number Ra (the dimensionless heating strength), the Prandtl number Pr (the ratio of kinematic viscosity to thermal diffusivity), and the aspect ratio Γ of the container. The key response parameters are the Nusselt number Nu (the dimensionless heat flux from the bottom to the top) and the Reynolds number Re (the dimensionless strength of the turbulent flow). While there is good agreement and understanding of the dependences Nu(Ra,Pr,Γ) up to Ra∼1011 (the “classical regime”), for even larger Ra in the so-called ultimate regime of Rayleigh-Bénard convection the experimental results and their interpretations are more diverse. The transition of the flow to this ultimate regime, which is characterized by strongly enhanced heat transfer, is due to the transition of laminar-type flow in the boundary layers to turbulent-type flow. Understanding this transition is of the utmost importance for extrapolating the heat transfer to large or strongly thermally driven systems. Here the theoretical results on this transition to the ultimate regime are reviewed and an attempt is made to reconcile the various experimental and numerical results. The transition toward the ultimate regime is interpreted as a non-normal–nonlinear and thus subcritical transition. Experimental and numerical strategies are suggested that can help to further illuminate the transition to the ultimate regime and the ultimate regime itself, for which a modified model for the scaling laws in its various subregimes is proposed. Similar transitions in related wall-bounded turbulent flows such as turbulent convection with centrifugal buoyancy and Taylor-Couette turbulence are also discussed.
Article
This research article delves into the natural convection of liquid metal in a three-dimensional cavity with varying magnetic fields and wall conductivities using direct numerical simulation. Two opposite sidewalls are heated/cooled, while the magnetic field is perpendicular to the main circulation. Our primary focus is on examining flows within the Grashof number, Gr⩽108, the Hartmann number, Ha⩽400, and the wall conductance ratio, Cw=0.01–1. It is found that weakly conducting walls experience a significant enhancement in convection within a specific range of magnetic field strength, whereas highly conducting walls exhibit pronounced flow attenuation. The applied horizontal magnetic field alters the plume's topology and dynamics, generating a more coherent and energetic large-scale flow structure, while it weakens convection by consuming buoyant potential energy through Joule dissipation. This results in a competition between the rectifying effect and the damping effect to determine whether the magnetic field has a positive or negative feedback on heat transfer, with the quasi-two-dimensional state serving as a critical point. Additionally, varying wall conductivities transform the current distribution within the parallel layer, influencing the flow's response to changes in field strength. The formation of corner vortices can be considered by the curvature of the boundary layer that undergoes turning at corners. Furthermore, the effects of wall shear and plume transport on heat transfer are systematically investigated. The analysis reveals that while the plume area remains almost constant, the condensation of coherent structures facilitates greater horizontal heat transport per unit area of the plume, contributing significantly to the overall heat transfer enhancement. Finally, the computed Nusselt number Nu and Reynolds number Re can be correlated as functions of Ha/Gr1/3. The critical conductance ratio referring to the complete suppression of convection conforms to the scaling of 2.31(Ha/Gr1/3)−5, where heat transfer occurs solely by thermal conduction once it exceeds this value.
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
Thermally driven vertical convection (VC)-the flow in a box heated on one side and cooled on the other side, is investigated using direct numerical simulations with Rayleigh numbers over the wide range of 10 7 ≤ Ra ≤ 10 14 and a fixed Prandtl number Pr = 10 in a two-dimensional convection cell with unit aspect ratio. It is found that the dependence of the mean vertical centre temperature gradient S on Ra shows three different regimes: in regime I (Ra 5 × 10 10), S is almost independent of Ra; in the newly identified regime II (5 × 10 10 Ra 10 13), S first increases with increasing Ra (regime II a), reaches its maximum and then decreases again (regime II b); and in regime III (Ra 10 13), S again becomes only weakly dependent on Ra, being slightly smaller than in regime I. The transition from regime I to regime II is related to the onset of unsteady flows arising from the ejection of plumes from the sidewall boundary layers. The maximum of S occurs when these plumes are ejected over approximately half of the area (downstream) of the sidewalls. The onset of regime III is characterized by the appearance of layered structures near the top and bottom horizontal walls. The flow in regime III is characterized by a well-mixed bulk region owing to continuous ejection of plumes over large fractions of the sidewalls, and, as a result of the efficient mixing, the mean temperature gradient in the centre S is smaller than that of regime I. In the three different regimes, significantly different †