You are on page 1of 16

International Journal of Thermal Sciences 108 (2016) 1e16

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


journal homepage: www.elsevier.com/locate/ijts

Numerical simulations of constant-inux gravity currents in conned


spaces: Application to thermal storage tanks
E. Kaloudis a, *, D.G.E. Grigoriadis b, E. Papanicolaou a
a
Solar and Other Energy Systems Laboratory, National Center for Scientic Research Demokritos, 153 10, Aghia Paraskevi, Attikis, Greece
b
Computational Sciences Laboratory UCY-CompSci, Department of Mechanical and Manufacturing Engineering, University of Cyprus, 75 Kallipoleos
Avenue, P.O. Box 20537, Nicosia, 1678, Cyprus

a r t i c l e i n f o a b s t r a c t

Article history: This study uses Computational Fluid Dynamics (CFD) to investigate numerically the ow phenomena in
Received 8 September 2015 the intrusion region of a thermal storage water tank during discharge. The early times of the discharging
Received in revised form process have signicant effect on the thermal mixing and the associated energy losses. A detailed time-
13 April 2016
evolution of the ow and temperature elds inside the tank was obtained for a range of relevant Froude
Accepted 14 April 2016
Available online 3 May 2016
(Fr) numbers. Parameters such as the thermocline thickness (d), the entropy generation rate S_g and the
thermal mixing factor (k) were calculated to quantify the mixing mechanism in the tank. Authors chose
several alternative discharging scenarios where the Froude number (Fr) varied between 0.05 and 2.00, a
Keywords:
Gravity currents
range which corresponds to typical discharging conditions in real applications involving water storage
Thermal storage tanks. The gravity currents (GCs) developing as the incoming cold uid ows along the oor of the tank,
Buoyancy driven ows their subsequent reection on the opposite vertical wall and the interaction between the reverse ow
Thermal mixing and the incoming ow were analyzed and correlated to d, S_g and k.
2016 Elsevier Masson SAS. All rights reserved.

1. Introduction thinner thermocline and thus less mixing. Considering a wide


range of inlet Froude number Frin 0.7  14.5 in their experimental
Thermal energy storage tanks permit excess thermal energy to study, they demonstrated the importance of the densimetric inlet
be stored until energy demand exceeds the immediately available Froude number Frin on the thermocline thickness. They concluded
solar energy supply. The dynamic process of retrieving the stored that Frin must be of the order of 1 or less for optimal design. Ji and
energy is called discharging. In order to achieve the best perfor- Homan [4] reported that the maximum entropy production occurs
mance (low mixing) during the discharging process, the thickness early in the discharging process when the cold uid moving along
of the intermediate region separating the hot and cold water layers the bottom in the form of a gravity current (GC) is reected on the
(thermocline) should be kept as small as possible (Dincer and Rosen opposite wall (Fig. 1). For the aforementioned reasons, the aim of
[1]). The initial moments of the discharging process play a signi- the present study is to identify and analyze the main mixing
cant role in thermal mixing in storage tanks (Kaloudis et al. [2]) due mechanisms that occur during this particular time period, in a tank
to the hydrodynamic and thermal disturbances caused by the of initially uniform temperature, through high-resolution, two-
interaction of the incoming and outgoing streams of water. In dimensional numerical simulations.
particular, the mixing-rate (quantied in terms of thermocline The GC ow conguration in the discharging of a storage tank is
thickness) is 67% higher in the initial period, in contrast to the later characterized by complex hydrodynamic phenomena, as may be
times of the process where diffusion is the dominating mixing seen in Fig. 1. The GC ow propagates with almost a steady velocity
mechanism. Yoo et al. [3] pointed out the importance of designing (uf) across the tank oor (a phase called rst pass by some authors
inlet diffusers in such a way that the water ows along the oor of 5) until it reaches the opposite wall (Fig. 1a). After the impact, it
the tank in the form of a density (gravity) current, which leads to a penetrates vertically into the tank up to a maximum height (Fig. 1b)
where it gets deected horizontally and a reverse ow sets in
(Fig. 1c). This reverse ow (second pass) interacts with the GC ow
* Corresponding author. in the rst pass, generating a shear layer which is subject to in-
E-mail addresses: skaloudis@ipta.demokritos.gr (E. Kaloudis), grigoria@ucy.ac.cy stabilities and potentially increased mixing. Depending on the ow
(D.G.E. Grigoriadis), elpapa@ipta.demokritos.gr (E. Papanicolaou).

http://dx.doi.org/10.1016/j.ijthermalsci.2016.04.018
1290-0729/ 2016 Elsevier Masson SAS. All rights reserved.
2 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16

enclosures such as storage tanks which is the subject of interest


here. Among the few relevant studies belong the aforementioned
ones by Nakos [5], Ji and Homan [4], as well as Homan and Soo [14].
In his study, Nakos [5] also conrmed the importance of the hy-
drodynamics of gravity currents in the efcient storage of chilled
water. Homan and Soo [14] studied numerically an orthogonal tank
and the internal wave phenomena occurring by the impact of the
GC to the opposite wall, with his results being limited to Frin 1. A
more recent study by Nabi and Flynn [15] showed that similar
situations may also arise in building ventilation problems. The
present work aims at further advancing the study of GCs in
conned spaces, by considering their interactions with solid
boundaries in storage tanks and their implications with regard to
mixing, by using Direct Numerical Simulations (DNS).
In the sections to follow, the mathematical formulation is rst
presented in Section 2, as well as the criteria used to quantify the
mixing process. The numerical code and the choice of computa-
tional domain and grid is presented in Section 3. Section 4 contains
the presentation of results and discussion, followed by the con-
clusions in Section 5.

2. Mathematical formulation

2.1. Governing equations

Using the Boussinesq's approximation, i.e. a linear variation of


Fig. 1. A schematic depiction of the investigated ow-conguration and the main ow
density r, with temperature T according to r  ro/ro b(T  To), the
patterns: GC propagation along the bottom (1st pass), wall impact and reection,
reverse ow (2nd pass). The outow boundary is the entire top surface of the domain, non-dimensional equations of motion and energy for mixed con-
placed at a sufcient vertical distance, here H* 10. vection problems become:

vUi
0 (1)
parameters, the reverse ow may also reach the inlet (Fig. 1d) vxi
where it is subjected to a second reection causing new distur-
bances in the developing thermocline. !
vUi vUi Uj vp 1 v vUi Gr 
GCs have been studied over the past three decades mainly due    q diy (2)
to their regular occurrence in the elds of geophysics, oceanog- vt  vxj vxi Re vxj vxj Re2
raphy, meteorology etc. In the relevant investigations, the density

!
rather than temperature was used as the transport variable. Typical
vq vUj q 1 v vq
applications involved problems such as lock-exchange ows (Hartel (3)
vt  vxj RePr vxj vxj
et al. [6], Ooi et al. [7], Ghasemi et al. [8]), entrainment of two-
dimensional gravity currents (Hallworth et al. [9]) and propaga-
tion of a gravity current into a two-layer stratied ambient uid where U is the dimensionless velocity, * denotes the rest of the
(White and Helfrich [10]). In the majority of cases, the domain of dimensionless variables and diy is the Kronecker's delta.
interest is characterized by an unlimited extent in the direction of These equations have been non-dimensionalised using the
the GC propagation. Quite a few studies have considered the height of the inlet diffuser Hin as the characteristic length scale, the
interaction of a two-layer ow with solid obstacles mounted at the inlet velocity uin as the characteristic velocity and the ratio Hin/uin as
bottom of the channel, with heavier uid lying below a lighter one. the characteristic time scale. The initial temperature of the tank To
For this problem, and by means of an inviscid hydraulic theory, and the inlet water temperature Tin are used to dene the dimen-
several regimes of ow (subcritical or supercritical, partially sionless temperature as q* (T  Tin)/(To  Tin). The Reynolds and
blocked or completely blocked ow) have been identied, Grashof dimensionless numbers are dened as Re uinHin/n and
3 T  T =n2 . The ratio Gr/Re2 also denes the Archi-
Gr g bHin
depending of the value of the Froude number upstream of the o in

obstacle (Rottman et al. [11], Rottman and Simpson [12], Lane-Serff medes number Ar which expresses the relative importance of free
et al. [13]). The present situation has similar characteristics to the to forced convection.
completely blocked ow described in the aforementioned problem,
as the presence of the vertical wall opposite the inlet does not allow
2.2. Methods to assess mixing
any possibilities other then the full reversal of the ow direction of
the GC, irrespective of the Froude number value. The completely
One of the most popular methods to quantify the energy losses
blocked ow over an obstacle normally gives rise to an internal
associated with mixing during the charge/discharge phases, is to
bore, which is an abrupt jump in the level of the interface between
analyze the evolution of the thermocline thickness (Haller et al.
the two uid layers, moving in the direction opposite to that of the
[16]) presented in Section 2.2.1. A more detailed view of the thermal
GC 12. Essentially, the second-pass of the GC mentioned above is
mixing and its spatial and temporal variation can also be obtained
marked by this internal bore movement, which also plays an
by other indicators such as the entropy production rate (Ji and
important role in the ensuing mixing in the tank.
Homan [4]) and the thermal mixing factor (Homan and Soo [17])
Few studies have considered GC ows in completely conned
which are presented in the following sections.
E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16 3

2.2.1. Thermocline thickness


 2
Entrainment is one of the most important features of gravity   Pe d
k td  (8)
currents, increasing the total volume of incoming uid and simul- td 4c1
taneously increasing its temperature (Cheong and Han [18]). The
net entrained volume is dened as Ve Vt  Vo, where Vt is the total where the coefcient c1 is calculated as
R
volume of cold uid and Vo Q_ in dt is the volume uxed through c1 1:46  0:3633:2Te  1:01=2 , Te is an edge temperature
the entrance (Cheong and Han [18]). This index quanties the dened by the user (here Te 0.1) and the time td is non-dimen-
volume of water that has been subjected to thermal mixing and is sionalized with the tank's lling time, td t=tfill .
closely related to the thermocline thickness index. The thermocline
thickness d* is dened as the distance (or volume) where 3. Numerical methods
q 0:1 qh and q 0:9 qh , with qh the highest temperature
encountered in the hot volume of water. In the present study, the 3.1. Computational code
evolution of the thermocline thickness d* will be used to quantify
mixing and it will be compared with the variation reported by The numerical code used in the present study was based on a
relevant studies in the literature. In particular, d*(t) will be calcu- fully-explicit, fractional-step approach and nite-differences on
lated as the volume of uid with temperature in the range of staggered, Cartesian grids. The solid boundaries are not described
0:1 qh < q < 0:9 qh divided by the length (L) of the cavity. via boundary conforming grids but using the immersed boundary
method 22. Hence at the vicinity of solid boundaries, the solution is
locally reconstructed, since the surfaces of the tank do not coincide
2.2.2. Volumetric entropy generation rate with the grid lines. A forcing procedure is followed to dynamically
Entropy production in thermouid processes, characterizes the mimic the existence of solid boundaries as proposed by Balaras
degradation of thermal and mechanical energy. For the present [23]. Using orthogonal Cartesian grids and a staggered variable
study this undesirable effect corresponds to thermal losses which arrangement, this code uses a fast, direct solver for pressure solu-
can be determined with the help of the second law of thermody- tion. Time advancement was performed using a fully explicit sec-
namics (Bejan [19]). The volumetric entropy generation rate can be ond order Adams-Bashforth scheme, and the momentum equation
calculated by the following equation: was discretised with a central, nite difference scheme. The tem-
perature equation was discretised using the higher order HLPA
000
k m
S_g 2 VT2 F; (4) scheme (Zhu [24]). The choice of the HLPA scheme for simulating
T T mixed convection ows in water storage tanks has been justied in
The volumetric entropy generation rate in its dimensionless previous studies, such as the ones by Papanicolaou and Belessiotis
form is given by the following equation (Ji and Homan [4]): [25] and Kaloudis et al. [26].
The efciency of the resulting solver is such that less than
  2
*000 t2 Vq Ec F 100 Mb of physical memory are required per million computational
S_g c   ; (5)
RePr 1 tc q 2 Re 1 tc q nodes. In DNS mode, the running performance allows more than
100 millions degrees of freedom to be updated every second on low
cost personal computers (e.g. on six-core CPU's such as an AMD
where Ec u2in =Cp To is the Eckert number, tc DT/To is the tem-
processor clocked at 3.2 GHz). A detailed description of this nu-
perature difference ratio and the volumetric entropy generation
*000 merical code can be found in Grigoriadis et al. [22], Grigoriadis et al.
rate S_g has been non-dimensionalised with rCpuin/Hin. The rst
[27]. This numerical code has been already tested and validated in a
term in Eq. (5) reects the entropy generation due to thermal
wide range of computational test cases, such as mixed convection
diffusion whereas the second term is the entropy generation due to
in storage tanks (Kaloudis et al. [26], Kaloudis et al. [2]), magne-
viscous effects, with F* being the dimensionless dissipation
tohydrodynamic ows (Grigoriadis et al. [27]), turbulent ows over
function:
rippled beds (Grigoriadis et al. [28], Grigoriadis et al. [29]) and
 2  2  2 turbulent ows past obstacles (Grigoriadis et al. [22]).
vU vV vV vU
F 2 (6)
vx vy vx vy
3.2. Simulation cases
For the purposes of the present study, it was also useful to
calculate the dimensionless cumulative entropy generation In this study, the authors chose six different discharging sce-
Sg;cum t  , which represents the total entropy produced within a narios where the inlet Froude number (Frin) varied between 0.05
given volume U and up to a specic time t*: and 2.00, corresponding to typical discharging conditions of water
storage tanks. Keeping the inlet temperature in the tank constant
Zt   000
 for all simulation cases (Tin 15  C or qin 0) the resulting sets of
1 _* dx dy dt 
Sg;cum t   S (7) simulation parameters can be found in Table 1.
U g
0
3.3. Boundary conditions

For the ow variables, no-slip conditions were imposed along


the tank walls. At the inlet slot, a uniform velocity prole was
2.2.3. Thermal mixing factor
considered. Along the outlet boundary, a convective boundary
Another approach to quantify the evolution of the thermocline
condition (Ferziger and Peric [30]) was imposed, namely:
thickness is to consider a mixing factor which reects the impact of
the multi-dimensional uid motion on the evolution of vf vf
d* (Tarawneh and Homan [20]). Ji and Homan [21] dened the uc 0; (9)
vt vx
thermal mixing factor (k), based on the thermocline thickness (d*),
as: where the variable f stands for any velocity component of velocity
4 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16

Table 1
Parameters and dimensionless numbers for each simulation case.

Case 1 Case 2 Case 3 Case 4 Case 5 Case 6

Tin[ C] 15.00 15.00 15.00 15.00 15.00 15.00


To[ C] 66.38 52.25 42.00 32.65 27.80 24.28
tc (Eq. (5)) 0.1783 0.1293 0.0937 0.0613 0.0444 0.0322
Frin 0.05 0.10 0.20 0.50 1.00 2.00
Pr 4.260 4.961 5.592 6.287 6.704 7.034
Rein 133.7 184.2 262.9 441.9 675.5 1056.5
Gr 7.15$106 3.39$106 1.73$106 7.81$105 4.56$105 2.79$105
Ec 1.00$1011 2.49$1011 6.28$1011 2.18$1010 5.71$1010 1.52$109

and uc is the convection speed velocity computed by the ow rate at In order to select a suitable grid resolution, the test Case with
the outlet. Adiabatic temperature boundary conditions were the highest Grashof number was considered, namely Case 1. To
considered along all solid boundaries of the computational domain, guarantee a grid-independent solution, four different grid resolu-
whereas at the outlet the normal derivative was set to zero. tions were constructed and tested (grids G1f, G2f, G3f, G4f), as
shown in Table 2. A comparison of the results using those grids can
been found in Fig. 3, where the grid convergence for the time
3.4. Computational domain e grid selection evolution of the thermocline thickness and entropy generation
rates is demonstrated. For a quantitative comparison also, the
Based on several preliminary calculations (not presented here), relative differences of each grid compared to the nest one are also
the length of the domain L*, was chosen as L* 10, so that for all reported in Table 3. Based on those differences, the numerical
values of Frin considered, the gravity current is allowed to develop resolution of grid G3f was considered as sufcient for the rest of the
and travel for a fair amount of time before reaching the opposite simulations presented hereafter. The average relative difference in
wall of the cavity. d* of G3f to the nest grid G4f was found below 1.4%. The corre-

In order to calculate the minimum domain height H* which sponding difference in S_g was 6.9% which was also considered
leaves the solution independent (at least for the initial times of the acceptable keeping in mind the high sensitivity in ow and

discharging process and the range of dynamical parameters temperature gradients. Similar values of relative error in S_g have
considered here), three different domain heights were considered, also been accepted in past studies, such as the one by Ji and Homan
namely G1H, G2H, G3H (Table 2). The numerical resolution of those [4].
three different grids was identical up to H* 6 close to the lower
wall, while extra computational nodes were added only to take into
4. Results and discussion
account the increased domain height for grids G2H and G3H. All
grids employ grid-clustering close to the wall regions. The param-
4.1. Flow description
eters that are used to compare the results, are the time variation of
the thermocline thickness d* and the total entropy generation rate
 The basic ow phenomena occurring during the initial times of
S_g . The latter, takes into account gradients of the ow and the
the discharging process were identied and described in Section 1.
temperature eld (Eq. (5)) making it suitable for grid-
In this section, the simulation results will be presented in three
independency tests. It should be noted that in order to obtain
parts, each one corresponding to a different phase: (i) GC propa-
comparable results among different grids, the spatial integration of
 gation (1st pass, Fig. 1a), (ii) GC impact on opposite wall (Fig. 1b)
S_g was only performed for 0 < H* < 6.
and (iii) GC reection, which is manifested through an internal bore
The test case considered to select the height of the computa-
development and its upstream propagation (2nd pass, Fig. 1c).
tional domain was Case 6 with Fr 2.0 where the maximum
penetration of the GC was expected. According to the simulation

results for d* and S_g , presented in Fig. 2, it is safe to conclude that 4.1.1. Gravity current propagation (1st pass)
none of the three domain heights affect the accuracy of the solu- Fig. 4 shows the detailed view of the ow eld near the inlet
tion. Even though the choice of H* 6 would be less computa- opening at the very beginning of the discharging process. Heavy,
tionally demanding, H* 10 was selected in order to capture the cold uid enters through the inlet and a GC starts to form and move
maximum penetration height of Case 6 also (see Fig. 17). Thus, for along the oor of the tank, initially unobstructed. Contrary to the
the rest of the simulations in this study the computational domain case of the lock-exchange problem, e.g. Hartel et al. [6], which is
was set to (L*, H*) (10,10). characterized by generation of GCs due to a xed volume release
(also called discontinuous GCs 31), here the GC is one of constant
inux release. As such, the present situation bears more similarities
Table 2 to the problem studied by Shringarpure et al. [32], who also
Data for grid and computational domain extent in each direction.
considered cases with return and no-return ow above the GC, the
Grid name Number of Computational Number of nodes at inlet lock-exchange problem being a typical example of the former case.
grid nodes domain Here, during the 1st pass, signicant horizontal motion takes place
NX NY L/Hin H/Hin only in one direction (no-return ow), with only the bottom-
G1H 256 198 10 6 78 located, high density (low-temperature) uid moving from left to
G2H 256 223 10 8 78 right. A return ow situation arises only during the 2nd pass, as it
G3H 256 248 10 10 78 will be discussed below.
G1f 182 143 10 10 40 This early stage of heavy uid release into a volume of lighter
G2f 256 248 10 10 78 uid is known as the slumping phase 33 and a characteristic
G3f 384 384 10 10 110 quantity is the fractional depth f h/H, where h is the depth of the
G4f 616 571 10 10 152
GC, measured at a specic location, and whose value changes
E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16 5

Fig. 2. Effect of domain height on the time variation of the thermocline thickness (a) and total entropy generation rate (b) for Case 6.


Fig. 3. Effect of numerical resolution on the time evolution of the thermocline thickness d* (a) and entropy generation rate S_g (b) for Case 1.

Table 3 propagates along the oor, the recirculation in the billows, typically
Grid comparison in terms of relative difference (%) for the thermocline thickness and present in Kelvin-Helmholtz instability and rolling up in the region
entropy generation rate.
of velocity shear immediately behind the head of the GC, has been
G1f/G4f G2f/G4f G3f/G4f identied as the main mixing mechanism. For both Cases 3 and 4
d*(avg) % 11.4 5.9 1.4 (Frin 0.2 and Frin 0.5 respectively), a big recirculation zone in the
 form of a billow at the back of the GC head can clearly be distin-
S_g avg % 45.9 21.3 6.9
guished in Figs. 5 and 6, followed by one or two more, smaller ones.
A special mention should be made here of Case 5 (Frin 1) in which,
based on the temperature and vorticity contours of Figs. 7e9, the
rapidly during this period and H the total depth (here the tank
second billow develops almost at the beginning of discharge
height). As can be seen in Fig. 4, this value is strongly dependent on
(t* 4.4), merges with the head at later times (t* 8) and separates
the inlet Froude number and in the present situation, by consid-
again at t* 11.
ering an estimate for the depth h at the location where this attains
In the most widely studied case so far in the GC literature, the
its minimum after the release (shown with the arrows in Fig. 4), it is
lock-exchange problem, and during the initial slumping phase
found to increase from a value roughly equal to f 0.01 for
mentioned above, the GC has been found to move at constant speed
Frin 0.05, to f 0.03 for Frin 0.2 and to f z 0.095 for Frin 1.0.
and to maintain a constant depth 33. This may possibly be suc-
The rst two values, when compared against the criterion intro-
ceeded by an inertial phase before eventually the GC enters its nal
duced by 33 with a cutoff value of f 0.075, are found to lie in the
viscous phase. Some authors report also of an initial phase pre-
range characterizing a deep-ambient (small f) conguration,
ceding all the above, the acceleration phase 35, whereby the front
whereas the value for Frin 1.0 is shown to lie in the large-
velocity increases rapidly to a maximum value, before dropping
fractional depth range. This point will be further discussed below,
again to an almost constant value characterising the slumping
for a later GC development phase, after a clearly dened head, front
phase.
and body of the GC have been established.
In view of the above, and with the understanding that in the
Proceeding into the next stage of development of the GC, the
present case the GC is generated under different conditions (con-
temperature and vorticity contours are shown in Fig. 5 for an in-
stant inux), the front location Xf for several values of the Froude
termediate value of Frin 0.2 (Case 3 in Table 1). Those are very
number is plotted in Fig. 10a on a logarithmic scale, as a function of
illustrative in distinguishing the main characteristics of a GC, such
a time non-dimensionalized by tb Hin/ub and over an interval
as the head, main body and nose (Simpson [34]). As long as the GC
corresponding to the rst pass. On the same plot, respective results
6 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16

Fig. 5. Temperature and vorticity contours after the creation of the GC for Frin 0.20.

Fig. 4. Effect of Froude number at very early times of the cold uid release. Detailed
view at the inlet region, with colors corresponding to uid temperature. (For inter-
pretation of the references to color in this gure caption, the reader is referred to the
web version of this article.)

from recent representative studies of lock-exchange problems from


Ooi et al. [7] and Paik et al. [31], are plotted for comparison. In order
to relate to those studies, some additional quantities characterizing
a GC, such as the front speed Uf and height hf , fractional depth f
based on hf, as well as the front Froude number Frh based on two Fig. 6. Temperature and vorticity contours after the creation of the GC for Frin 0.50.
alternative denitions, have been calculated and shown on Table 4.
The rst denition of Frh, Frh,i, is needed since the above mentioned
studies use the height of the lock in their analysis i.e. a xed chosen dimensionless value of temperature, here taken equal to
quantity and analogous to the height of the inlet slot in the present 0.9. Based on this value, the respective fractional depths f are
case. On the other hand, the second denition, also very common in computed, which may be slightly different than those estimated
the GC literature, is based on the front depth hf, which is a quantity earlier just downstream of the inlet slot, as shown in Fig. 4. How-
that varies along the distance travelled by the GC and is also ever, the values taken at this stage of development of the GC are the
dependent on Frin. The front depth is computed according to most representative ones for f, in line with what is common
Huppert and Simpson [33], i.e., at a horizontal location just behind practice in the literature.
the head and, more specically, at a point where the interface be- What may be observed in Fig. 10a for all values of Frin is that after
tween GC body and ambient uid rst becomes horizontal. The an initial period, the slope of the lines changes and only for the two
value of hf is taken as the vertical location corresponding to a intermediate Froude numbers considered, Frin 0.1 and 0.2, may
E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16 7

Fig. 9. Temperature and vorticity contours after the creation of the GC (t* 11) for
Fig. 7. Temperature and vorticity contours after the creation of the GC (t* 4.4) for Frin 1.00: billow reseparation.
Frin 1.00: two distinct billows.

according to Refs. 7 and 31, both of which associated the value of 1


with the slumping phase of the gravity current, characterized by a
constant front speed. In the present case however, and as can be
better observed in Fig. 10b, the front speed does not attain a con-
stant value at large times. Here only data corresponding to the GC
development are shown (t/tw < 0.8  0.9) and the last time interval,
as the GC approaches the end wall is shown separately below. It
should also be noted that in the plot of Fig. 10b the front velocity is
non-dimensionalized by the sum of ub and uin rather than alone, to
make it comparable to the energy conserving limit of 0.5 predicted
by theory (Benjamin [36]). Unlike lock-exchange ows, which only
carry potential energy upon release, here there is an incoming ow
carrying an amount of kinetic energy and this component should be
taken into account.
What may be clearly observed is an acceleration phase at early
times, at the end of which uf reaches a peak, then drops to a local
minimum value, followed by a new acceleration and slow decay.
Regarding this phenomenon of an acceleration phase and the
subsequent deceleration, limited evidence has been reported in the
literature, such as in the studies by Paik et al. [31] and Cantero et al.
[35]. In fact, the latter study has obtained a non-monotonic
behavior of the front speed at early times and attributed this
observation to the behavior of two-dimensional GCs which exhibit
strong coherent vortices. In the present study, this is also the case,
as for Fri 0.2 for instance, the time period corresponding to the
deceleration (t* 0.5  1.0 or t  =tw
 0:1  0:2 in Fig. 10b) is exactly

the period of the interface roll-up (see inset in Fig. 10b). It should be
noted that the fact that the roll-up vortices are more prominent in
Fig. 8. Temperature and vorticity contours after the creation of the GC (t* 8) for 2D simulations such as the present ones do not render the results
Frin 1.00: billow merging. less realistic for the present practical problem of interest, the
rectangular storage tanks. In such congurations, the inlet-outlet
diffusers should be designed in such a manner as to maintain 2D
the slope be assumed to approach the value of 1 and only for short rather than 3D ow features and to delay the onset of spanwise
times. The non-dimensionalization of front speed shown here is
8 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16

xed volume release. The time variation of this quantity is shown in


Fig. 11a and a self-similarity type of relationship has been sought,
yielding an exponent of 1/4 for the best t. This shows a milder
deceleration in the present case, compared to the 1/3 decay in the
lock exchange problem (Cantero et al. [35]), whereby the inertial
phase is associated with a reected bore catching up with the GC
front and both start travelling at a common speed, gradually
reducing (deceleration).
To further examine whether the inertial phase is succeeded by a
viscous phase in our case, we followed the analysis of Hallez and
Magnaudet [37], who presented, among others, specic transition
criteria to the buoyancy-viscous phase for lock-exchange congu-
rations, distinguishing between short (SR) and long reservoirs (LR)
respectively. The latter are more comparable to the present case of
constant inux GCs and the same special normalized front position

Fig. 10. Time evolution of: a) the front location and b) the front speed for different
inlet Froude numbers. The inset in (b) shows the ow structures (vorticity contours)
near the inlet, responsible for the non-monotonic variation of uf at early times.

Table 4
Characteristic dimensionless quantities of the gravity current during the 1st pass:
front speed (Uf), front height hf , fractional depth (f), front Froude numbers (based
on inlet and front heights), time for impact on vertical wall.

Frin Uf hf f Frh,i Frh,f tw

Case 1 0.05 5.05 0.186 0.019 0.254 0.585 1.7


Case 2 0.10 3.63 0.263 0.026 0.365 0.708 2.7
Case 3 0.20 2.44 0.350 0.035 0.489 0.824 4.1
Case 4 0.50 1.39 0.505 0.050 0.695 0.977 7.2
Case 5 1.00 0.93 0.805 0.081 0.929 1.034 11.3
Case 6 2.00 0.60 1.183 0.118 1.203 1.105 16.9

instabilities in order to achieve reduced mixing and better storage


efciency (Yoo et al. [3], Papanicolaou and Belessiotis [25]). After
this period, the GC head, along with a clearly discernable horizontal
interface between GC and ambient uid behind the head, have
formed as seen in Figs. 4b and 5 and the phase with GC propagation
at a nearly constant-speed (slumping phase) should follow.
As already mentioned, however, here the front speed does not
attain a constant value at later times but rather a value that grad-
ually decays. This is an indication that the GC ow may be tran-
sitioning into a different regime. To investigate such a
development, i.e., a potential transition to a buoyancy-inertia
phase, the non-dimensional front velocity (uf/ub) is scaled here
with the respective incoming ow rate (ui/ub) to the 1/3 power, and
a dimensionless group is formed analogous to the one dened by
Cantero et al. [35]. Those authors have scaled the velocity by the Fig. 11. Front velocity vs. dimensionless time and search for characteristic phases: a)
volume of uid release instead, given that their problem was one of inertial (self-similar) phase (Ref. [35]) and b) transition to a viscous phase (as in
Ref. [37]).
E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16 9

and time variables, b x f and bt respectively dened by the above au- due to constant GC front speed, a disturbance (decrease of tem-
thors are also employed for this purpose. For this non- perature) at t  =tw
 0:5 for all cases is observed indicating the rst

dimensionalization, length and time scales dened as a2ReFr1 pass. It is evident that for larger Frin the temperature disturbance
and a2ReFr2 respectively are used, where a 0.0922 for two has a longer duration. This phenomenon is associated with the
dimensional geometries, such as in the present case. By recasting respective increase in size of the GC head.
our own Xf vs. time data for all Froude numbers with respect to In Fig. 14 temperature contours for all simulation cases are
those variables we obtained the picture shown in Fig. 11b. The b xf presented shortly before the GC impact on the vertical wall,
bt line corresponds to a constant velocity regime, whereas the b xf whereby the form and size of GC front and body are still clearly
bt 1=2 line to the viscous regime. The intersection b x f ; bt 1; 1 is the dened. It is obvious that for increasing Frin the size of the GC (body
transition point to the viscous regime and as can be seen, with the and head height) grows accordingly. This is due to smaller buoy-
present data, this point is barely reached and this only for the two ancy effects, since the temperature difference for higher Frin be-
highest Froude numbers. Therefore, our data do not show a formal comes smaller. Also, as seen on Table 4 and in Fig. 14, the
transition to the viscous regime for the chosen values of Re and dimensionless time at which the GC reaches the opposite wall
whatever viscous effects are present must act only in the preceding differs signicantly between cases. This is a result of the different
inertial phase. As will be discussed below however, the presence of front velocities (uf) for each GC.
the vertical end wall is expected to be an additional factor for In the literature there is very limited information regarding the
deceleration and the reduced values of Frh obtained. inuence of a solid boundary on the development of GC, particu-
The above effects are also the reason why the values of the front larly in relation to the front Froude number. Ermanyuk and Gavrilov
Froude number Frh computed by both denitions shown on Table 4 [38] report some relevant information in terms of the variation of
are lower than the theoretical ones. The inviscid theory and ex- Froude number with the fractional depth for the GC ow over an
periments in lock-exchange generated GCs with an unlimited obstacle. They show that as f tends to 0, there is a steep drop in Frh
horizontal domain of propagation show that for small values of the to values below 0.5, rather than an increase towards the theoretical
fractional depth f such as the ones shown on Table 4 (deep limiting value of 2. This is in very close agreement to the present
ambient) the front Froude number should tend to 2. This is also the ndings shown in Table 4, which show a decreasing trend in the
case for constant-inux GCs (or starting ows), for which the values of Frh with decreasing f.
variation of Frh with f is not as strong 34. As will be discussed below The variation of front velocity as it approaches the end wall is
however, the presence of the vertical end wall is expected to be an shown in detail in Fig. 15 for various Froude numbers and the same
additional factor for deceleration and the reduced values of Frh non-dimensionalization used in Fig. 10b for consistency. It may be
obtained. observed that for most Froude numbers the deceleration starting
In Table 4, Uf has been listed for each case and as seen in Fig. 12a, roughly for t/tw > 0.9 is fairly steep, with the exemption of the
the Frin variation of Uf could be approximated by a power-law curve lowest value (Fr < 0.05), where it is milder.
of the form Uf 0:923 Frin 0:583 , making possible to estimate t  for
w
various domain lengths. Again, higher values of Uf for lower Frin are
4.1.2. Gravity current impact on opposite wall
a result of higher buoyancy forces which suppress the cold uid,
The next phase in the GC development follows after the GC has
forcing it to move sideways and minimizing any vertical
propagated along the entire tank oor, reaching the opposite wall
movements.
for impact. As can be seen in the temperature contours of Fig. 16 for
Fig. 12b also indicates that a Frin correlation exists for the GC
Frin 1, the GC approaches the opposite wall (t* 2.5), crashes and
body height (hf). Specically, a power-law curve
penetrates vertically into the tank up to a maximum height (t* 3).
hf 0:793 Frin 0:493 can approximate the increase in GC height h
f After the impact, it gets reected (t* 3.5) and a reverse ow sets in
for the range of Frin tested. The values of hf are also presented in
(t* 3.8). This situation is very similar to the one arising in the
Table 4.
completely blocked GC ow described by Rottman et al. [11], who
In Fig. 13 temperature values are presented over time at
characterize it as a hydraulic jump. The Q-criteria contours
different vertical locations in the tank mid-length (x* 5), for each
(Q 1=2Uij Uij  Sij Sij > 0, where Uij is the ltered vorticity tensor
computational case. In the abscissa, time has been non-
[39]) plotted in Fig. 16 help us identify the destruction of the
dimensionalized with the time needed for the GC to reach the
 , shown for all cases on Table 4. As expected, and recirculation in the GC head. Nevertheless, the temperature-eld
opposite wall tw
disturbances created at the back of the GC head are still present.

Fig. 12. Gravity current velocity (a), body height (b), for all Froude numbers (1st and 2nd pass).
10 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16

Fig. 14. Temperature contours for all simulation cases at the rst pass. The form and
size of gravity currents for all values of Frin is presented at the moment immediately
before their impact onto the opposite wall.

ows 7. So the terms internal bore or reected GC may be


here used interchangeably and are the main phenomenon charac-
terizing the second pass. The interaction of this reverse ow with
Fig. 13. Temperature values over time for a location in the middle of the cavity length the continuous GC ow in the rst pass, generates a shear layer and
(x* 5) at different heights for each computational case (1st pass).
depending on Frin the bore propagation could also bear some of the
GC characteristics.
Fig. 17 shows the temperature contours of GC for all cases In Fig. 19, temperature values are presented over time for
reaching a maximum penetration height hmax . It is clear that different locations at the cavity mid-length (x* 5) and at different
larger Frin values correspond to bigger hmax due to smaller buoy- heights for each computational case. A clear disturbance (localized
ancy forces. In Table 5, hmax is reported and according to Fig. 18 the drop in temperature) indicates the second pass. Nevertheless, it is
variation of hmax with respect to Frin, could be also approximated by only for lower values of Frin that this disturbance has similarities to
a power-law curve hmax 5:031 Frin 0:384 . the GC rst pass, as already presented in Fig. 13.
In Table 5 the values of reected ow velocity (Uf,2) and bore
height hb;2 are listed. A comparison with the GC front speed (Uf)
4.1.3. Reection of the gravity current e upstream internal bore and its body height hf in the rst pass is presented in Fig. 12a and
propagation b respectively. As expected, Uf,2 has lower values than Uf since the
The third distinct phase of the discharge process, as already momentum of the rst pass has decreased due to the impact onto
been mentioned in Section 1, is the propagation of the reected GC the vertical wall, opposite the inlet. Similar to Uf, Uf,2 can also be
towards the inlet wall. This phenomenon may also be characterized approximated
  with a power law equation, i.e.
as the motion of an internal bore 12, which is essentially a moving   0:638 . According to Fig. 12b the body height of the
Uf ;2  0:532 Frin
hydraulic jump and is also encountered in many lock-exchange
E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16 11

entrance region at t* 3.1. This pattern resembles in form the


undular bores observed in geophysical ows, as discussed, among
others, by Crook and Miller [40] and Cummins and Armi [41].
Furthermore, two re-circulations (Kelvin-Helmholtz instabilities)
could be observed in the cold uid region. Similar behavior was
observed for Frin 0.1 (Fig. 21) and Frin 0.2 (not shown here).

4.2. Thermal mixing

In order to quantify mixing, the dimensionless thermocline


thickness (d*), the dimensionless volumetric entropy generation

rate S_g and the thermal mixing factor (k) were calculated. Fig. 22
shows the time evolution of the aforementioned parameters for
each simulation case. Also, for each case the time at which the GC
reaches the opposite wall tw  , dening the rst characteristic

period (rst pass), has been marked with a vertical dashed line (1st
one to the left).
For the rst-pass period t  0  tw  , the calculation of the b

coefcient is performed, following Homan [42] suggestion of a


relationship of the form d* ~ t*b. Results are shown in Table 5. An
average value of b 1.22 is obtained, which is signicantly higher
Fig. 15. Time evolution of the front speed for different inlet Froude numbers showing
compared to later times of a discharging process (b 0.33  0.43)
the deceleration as the GC approaches the end wall (reached at t tw).
(Homan [42], Kaloudis et al. [2]) where diffusion is the main mixing
mechanism. This enhanced mixing caused by the GC movement

reected ow (bore) has larger values than the GC front depth hf , compared to later times is also illustrated by the behavior of S_g in
_ 
which is expected since the volume of cold uid is constantly Fig. 22a (green lines). Sg values show an ascending behavior until a
maximum value is attained near tw  . For smaller Fr , the time of
increasing. Nonetheless, the power law approximation 
in
hb;2 4:162 Frin0:616 shows that h is increasing at a faster rate than maximum S_g coincides with tw  but as Fr takes greater values the
in
b;2 
hf . rst S_g maximum is shifted to later times. This leads to the
It is worth mentioning that for Frin 2, Uf,2 and hb;2 could not be conclusion that the GC impact on the opposite wall affects mixing
determined, since there is no distinct reverse ow for a 2nd pass. strongly for smaller Frin, since at larger Frin, the wall cannot inhibit

This is in accordance to Yoo et al. [3] study, where it is mentioned the further increase of S_g values.

that for Frin > 2, the ow characteristics show jet-like rather than Another pattern that characterizes the time evolution of S_g is the
buoyancy-driven behavior. appearance of a secondary maximum (marked with a dashed line at
a dimensionless time t2nd  in Fig. 22). In order to further investigate
Another characteristic of this phase is the wavy pattern obtained
by the interface between reected GC and incoming cold uid. In and explain this behavior, the temperature contours for all simu-

lation cases at time t2nd are presented in Fig. 23. For all cases, ow
Fig. 20 temperature and vorticity contours are presented showing
this pattern extending from the middle of the tank towards the elds share common characteristics. More specically, the reected
GC is travelling backwards, towards the inlet diffuser (second pass)

Fig. 16. Temperature (top) and Q-criterion contours (bottom) for Frin 0.10. The destruction of the GC head recirculation can be observed.
12 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16

Fig. 17. Temperature contours for all cases at the moment of maximum penetration height.

Table 5 mentioned. This is the result of the shear layer forming between the
Characteristic quantities of the reected gravity current (internal bore) relating to two ow streams (GC and incoming water) and the inuence of the
the wall impact and the 2nd pass: maximum penetration hmax , reected GC speed, opposite wall. The inuence of Frin is also apparent in Fig. 23, as for
height (Uf,2, hb;2 ), exponent b in the expression d* ~ t*b for all cases.
the lower values of Frin, a greater number of vortices appear in the
Frin hmax Uf,2 hb;2 b aforementioned region and a more distinct secondary maximum
Case 1 0.05 1.65 3.64 0.73 1.19
appears (Fig. 22).

Case 2 0.10 2.06 2.21 0.96 1.15 Another characteristic of the S_g values at all times is that they
Case 3 0.20 2.66 1.54 1.36 1.21 become smaller as Frin increases (Fig. 22). In Fig. 24 the maximum
  ) and the cumulative S_
Case 4 0.50 3.65 0.83 2.67 1.28 S_g;max (at t  t2nd  
g;cum (until t t2nd ) values
Case 5 1.00 5.08 0.52 4.51 1.27
are plotted against Frin numbers and the decrease of both param-
Case 6 2.00 6.80 e e 1.24
eters could be approximated by power law curves
 1:380 ; S_
S_g;max 6:7  107 Frin 5 0:732 .
g;cum 1:3  10 Frin Since
smaller values of Frin indicate bigger temperature differences, the
buoyancy forces (rst term of Eq. (5)) dominate, affecting the en-
tropy production more than the viscous effects (second term of Eq.
(5)).
The time evolution of thermal mixing factor k is shown in Fig. 22
(blue lines). At a closer look, a correlation with d* vs. time evolution
is observed. In particular, both parameters are increasing until
 , and afterwards, when the growth rate of d* decreases, the
t  t2nd
values (not the rate) of k decreases. The above transient results are
in accordance to Homan and Soo [17] study, where the time traces
of k show similar behavior: zero k values at the beginning of the
discharging process, reaching a maximum value at early times and
then decrease and remain steady to lower values of k. According to
the same study this behavior should be expected, since larger k
values indicate enhanced mixing occurring at the very early times
of the discharging process (until the end of the second pass).
Fig. 25 shows k at t  t2nd
 for all simulation cases, in compar-
ison with k results of uniform (Cole and Bellinger [43]) and non-
uniform (Zurigat et al. [44]) diffusivity model according to Ji and
Homan [21] calculations and Wildin and Sohn [45] experimental
data. k 1 represents the optimal discharging process with the
maximum achievable efciencies (Homan and Soo [17]). Both
models predict k 1 until Frin 2 Pe 400  1000 and then k increases

Fig. 18. Maximum penetration height after the GC's impact on the opposite wall, for all exponentially. The present simulation results of k (just as the
Froude numbers.
experimental data of Wildin and Sohn [45]) show a similar trend
but bigger values from the diffusivity models. The reason for this
difference is due to enhanced mixing of the early stages of the
having already covered the rst half of the distance and a group of
process.
vortices are created ahead of the GC and above the inlet as already
E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16 13

Fig. 20. Temperature and vorticity contours showing a wavy pattern in the entrance
region during the GC second pass (Fr 0.05).

Fig. 19. Temperature values over time for a location in the middle of the cavity (x* 5)
at different heights for each computational case (2nd pass).
Fig. 21. Temperature and vorticity contours showing a wavy pattern in the entrance
region during the GC second pass (Fr 0.10).
5. Conclusions

The creation and development of gravity currents during the propagation has been demonstrated and compared with results for
discharging process of rectangular storage tanks and the associated unobstructed GC development from the literature. Also, it was
thermal mixing were studied numerically with the use of the DNS during the rst pass that the highest mixing rate (in terms of d*)
method. The interest lies in the interaction of GCs with solid was observed. The impact on the opposite wall was observed to
boundaries in a water storage tank and its effect on the efciency of have a less pronounced effect on mixing, since the GC mixing
storage. mechanisms have already died out on impact. After the impact on
Gravity current characteristics during the rst pass have been the opposite wall, a reected GC (internal bore) is created propa-
analyzed and the effect of various transitions in the GC develop- gating towards the inlet diffuser (second pass). A secondary

ment and the obstruction due to the restricted domain of maximum of S_g values is produced before the reected GC reaches
14 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16


Fig. 22. Time evolution of dimensionless thermocline thickness (d*), dimensionless volumetric entropy generation rate S_g and thermal mixing factor (k).


Fig. 23. Temperature contours during the second pass of the GC for all simulation cases. The time instant for all cases coincides with a secondary maximum of S_g .

the side wall on the inlet side due to the creation of a group of studied and it was shown that a parametric analysis is possible
vortices. since the main parameters that characterize the ow and mixing

Six different cases (different respective Frin numbers) were (Uf, hmax , S_g ) could be well approximated by using Frin-depended
E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16 15

Nomenclature

Ar Archimedes number Ar g bH DT=u2in Gr=Re2


Cp specic heat under constant pressure [J kg1 K1]
Ec Eckert number Ec u2in =Cp To p
Frin inlet Froude number Frin uin = g 0 =H
Frh,f Froude number
qof the front, based on front depth
Frh;f uf = g 0 hf
Frh,i Froude number of the front, based on inlet height
g magnitude of the gravitational acceleration [m s2]
0 0
g reduced gravity g g (Dr/ro) [m s2]
Gr Grashof number based on the height of the tank
Gr g bHin3 DT=n2

H height of the tank [m]


Hin height of the inlet diffuser slot [m]
hf depth of the gravity current at the front [m]
k thermal conductivity [W m1 K1]
L length of the tank [m]
n coordinate in the direction normal to the wall (n x, y or
z)
p pressure [N m2]
Pr Prandtl number of water Pr n/a

Fig. 24. Volumetric entropy generation rate S_g for all cases. Q_ in volumetric ow rate, Q_ u h W, [m3 s1]
in in in
Re Reynolds number based on the height of the tank
Re uinHin/n
000
S_g volumetric entropy generation rate [W m3 K1]
t physical time [s]
tll ll time U=Q_ in , [s]
T local temperature [K]
To initial temperature of the uid in the tank [K]
x,y coordinate distance along the length and height of the
tank [m]
u,v Cartesian velocity components [m s1] p
ub buoyancy velocity based on the inlet height: ub g 0 Hin
1
uf velocity of the gravity current front [m s ]
uin inlet velocity [m s1]
U,V dimensionless Cartesian velocity components U u/uin,
V y/uin

Greek symbols
a thermal diffusivity [m2 s1]
b coefcient of thermal expansion of water b 1/U(vU/
vT)p, [K1]
d thermocline thickness [m]
d* dimensionless thermocline thickness d* d/H
DT temperature scale DT Tin  To, [K]
Fig. 25. Thermal mixing factor k for all simulation cases, in comparison with k results Dr characteristic density difference Dr rin  ro, [kg m3]
of uniform (Cole and Bellinger [43]) and non-uniform (Zurigat et al. [44]) diffusivity q* dimensionless temperature q* (T  Tin)/DT
model according to Ji and Homan [21] calculations.
n kinematic viscosity of water [m2 s1]
r density [kg m3]
power-law equations. The maximum penetration height hmax f fractional depth of gravity current f hf/H

increases as Frin increases. On the other hand, Uf and S_g are U tank volume [m3]
inversely proportional to Frin due to smaller buoyancy forces uz vorticity component along the spanwise direction [s1]
associated with higher Frin numbers. Results of thermal mixing
factor are in a qualitative agreement with mixing models showing
Subscripts
enhanced mixing for the starting times of the discharging process.
avg averaged value
in inlet related parameter
max maximum value
Acknowledgment o initial temperature in the tank

Author E. Kaloudis acknowledges the nancial support from the


graduate fellowship program of the National Center for Scientic Superscripts
Research Demokritos. * dimensionless variables
16 E. Kaloudis et al. / International Journal of Thermal Sciences 108 (2016) 1e16

References [22] Grigoriadis D, Bartzis J, Goulas A. LES of the ow past a rectangular cylinder
using the immersed boundary concept. Int J Numer Methods Fluids 2003;41:
615e32.
[1] Dincer A, Rosen M. Thermal energy storage, systems and applications. John
[23] Balaras E. Modeling complex boundaries using an external force eld on xed
Wiley and Sons, Ltd; 2010.
cartesian grids in large-eddy simulations. Comput Fluids 2004;33:375e404.
[2] Kaloudis E, Grigoriadis D, Papanicolaou E, Panidis T. Large eddy simulation of
[24] Zhu J. A low-diffusive and oscillation-free convection scheme. Commun Appl
thermocline ow phenomena and mixing during discharging of an initially
Numer Methods 1991;7:225e32.
homogeneous or stratied storage tank. Eur J Mech B Fluids 2014;48:94e114.
[25] Papanicolaou E, Belessiotis V. Transient development of ow and temperature
[3] Yoo J, Wildin M, Truman C. Initial formation of a thermocline in stratied
elds in an underground thermal storage tank under various charging modes.
thermal storage tanks. ASHRAE Trans 1986;92:280e92.
Sol Energy 2009;83:1161e76.
[4] Ji Y, Homan K. Transition from gravity- to inertia-dominated behavior
[26] Kaloudis E, Grigoriadis D, Papanicolaou E, Panidis T. Large eddy simulations of
computed for the turbulent stably-stratied lling of an open enclosure. Int J
turbulent mixed convection in the charging of a rectangular thermal storage
Heat Fluid Flow 2006;27:490e501.
tank. Int J Heat Fluid Flow 2013;44:776e91.
[5] Nakos J. Prediction of velocity and temperature proles in gravity currents for
[27] Grigoriadis D, Kassinos S, Votyakov E. Immersed boundary method for the
use in chilled water storage tanks. J Fluids Eng Trans ASME 1994;116:83e90.
MHD ows of liquid metals. J Comput Phys 2009;228:903e20.
[6] Hartel C, Meiburg E, Necker F. Analysis and direct numerical simulation of the
[28] Grigoriadis D, Dimas A, Balaras E. Large-eddy simulation of wave turbulent
ow at a gravity-current head. Part 1. Flow topology and front speed for slip
boundary layer over rippled bed. Coast Eng 2012;60:174e89.
and no-slip boundaries. J Fluid Mech 2000;418:189e212.
[29] Grigoriadis D, Balaras E, Dimas A. Coherent structures in oscillating turbulent
[7] Ooi S, Constantinescu G, Weber L. 2D large-eddy simulation of lock-exchange
boundary layers over a xed rippled bed. Flow Turbul Combust 2013;91:
gravity current ows at high Grashof numbers. J Hydraul Eng 2007;133:
565e85.
1037e47.
[30] Ferziger J, Peric M. Computational methods for uid dynamics. Springer;
[8] Ghasemi V, Firoozabadi B, Mahdinia M. 2D numerical simulation of density
2002.
currents using the SPH projection method. Eur J Mech B Fluids 2013;38:
[31] Paik J, Eghbalzadeh A, Sotiropoulos F. Three-dimensional unsteady RANS
38e46.
modeling of discontinuous gravity currents in rectangular domains. J Hydraul
[9] Hallworth M, Huppert H, Phillips J, Sparks R. Entrainment into two-
Eng 2009;135:505e21.
dimensional and axisymmetric turbulent gravity currents. J Fluid Mech
[32] Shringarpure M, Lee H, Ungarish M, Balachandar S. Front conditions of high-
1996;308:289e311.
Re gravity currents produced by constant and time-dependent inux: an
[10] White B, Helfrich K. A general description of a gravity current front propa-
analytical and numerical study. Eur J Mech B Fluids 2013;41:109e22.
gating in a two-layer stratied uid. J Fluid Mech 2012;711:545e75.
[33] Huppert H, Simpson J. The slumping of gravity currents. J Fluid Mech 1980;99:
[11] Rottman J, Simpson J, Hunt J, Britter R. Unsteady gravity current ows over
785e99.
obstacles: some observations and analysis related to the phase II trials.
[34] Simpson J. Mixing in the front of a gravity current. Acta Mech 1986;63:
J Hazard Mater 1985;11(C):325e40.
245e53.
[12] Rottman J, Simpson J. The formation of internal bores in the atmosphere: a
[35] Cantero M, Lee J, Balachandar S, Garcia M. On the front velocity of gravity
laboratory model. Quart J R Meteorol Soc 1989;115(488):941e63.
currents. J Fluid Mech 2007;586:1e39.
[13] Lane-Serff GF, Beal LM, Hadeld TD. Gravity current ow over obstacles.
[36] Benjamin TB. Gravity currents and related phenomena. J Fluid Mech 1968;31:
J Fluid Mech 1995;292:39e53.
209e48.
[14] Homan K, Soo S. Model of the transient stratied ow into a chilled-water
[37] Hallez Y, Magnaudet J. A numerical investigation of horizontal viscous gravity
storage tank. Int J Heat Mass Transf 1997;40:4367e77.
currents. J Fluid Mech 2009;630:71e91.
[15] Nabi S, Flynn M. The hydraulics of exchange ow between adjacent conned
[38] Ermanyuk E, Gavrilov N. Interaction of internal gravity current with an
building zones. Build Environ 2013;59:76e90.
obstacle on the channel bottom. J Appl Mech Tech Phys 2005;46:489e95.
[16] Haller M, Cruickshank C, Streicher S, Harrison W, Andersen E, Furbo S.
[39] Haller G. An objective denition of a vortex. J Fluid Mech 2005;525:1e26.
Methods to determine stratication efciency of thermal energy storage
[40] Crook N, Miller M. A numerical and analytical study of atmospheric undular
processes, review and theoretical comparison. Sol Energy 2009;83:1847e60.
bores. Quart J R Meteorol Soc 1985;111:225e42.
[17] Homan K, Soo S. Laminar ow efciency of stratied chilled-water storage
[41] Cummins P, Armi L. Upstream internal jumps in stratied sill ow: observa-
tanks. Int J Heat Fluid Flow 1998;19:69e78.
tions of formation, evolution, and release. J Phys Oceanogr 2010;40:1419e26.
[18] Cheong H, Han Y. Numerical study of two-dimensional gravity currents on a
[42] Homan K. Integral solutions for transient temperature proles in stably-
slope. J Oceanogr 1997;53:179e92.
stratied open enclosures. J Heat Transf 2003;125:273e81.
[19] Bejan A. Entropy generation minimization: the method of thermodynamic
[43] Cole R, Bellinger F. Natural thermal stratication in tanks: nal report. Phase
optimization of nite-size systems and nite-time processes. CRC Press; 1996.
1. Argonne National Laboratory; 1982.
[20] Tarawneh C, Homan K. Measurements of density prole evolution during the
[44] Zurigat YH, Liche PR, Ghajar AJ. Inuence of inlet geometry on mixing in
stably-stratied lling of an open enclosure. Int J Heat Fluid Flow 2008;29:
thermocline thermal energy storage. Int J Heat Mass Transf 1991;34:115e25.
1113e24.
[45] Wildin M, Sohn C. Flow and temperature distribution in a naturally stratied
[21] Ji Y, Homan K. On simplied models for the rate- and time-dependent per-
thermal storage tank. Technical Report, USACERL TR FE-94/01. US Army
formance of stratied thermal storage. J Energy Resour Technol 2007;129:
Construction Engineering Research Labotaries; 1993.
214e22.

You might also like