Sie sind auf Seite 1von 10

Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153

Contents lists available at SciVerse ScienceDirect

Journal of Non-Newtonian Fluid Mechanics


journal homepage: http://www.elsevier.com/locate/jnnfm

Thermal instabilities in a yield stress fluid: Existence and morphology


Anne Davaille a,⇑, Blandine Gueslin b, Anna Massmeyer a, Erika Di Giuseppe a
a
Laboratoire FAST (CNRS/UPMC/Univ. Paris-Sud), Bat. 502, Rue du Belvedere, Campus Universitaire, 91 405 Orsay Cedex, France
b
Equipe de Dynamique des Fluides Geologiques, IPGP, 1 Rue Jussieu, 75238 Paris Cedex 05, France

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

Article history: We present new laboratory experiments on the development of thermal plumes out of a localized heat
Available online 16 November 2012 source in Carbopol, a yield-stress and shear-thinning fluid. Depending on the yield number Y0, which
compares the thermally-induced stress to the yield stress, three different regimes obtain. For low Y0
Keywords: (<120), no convection develops; while for intermediate values, a small-scale convection cell appears
Convection and remains confined around the heater. For high Y0 (>260), thermal plumes develop. Their morphology
Yield stress differs from the mushroom-shape typically encountered in Newtonian fluids. Combined temperature and
Herschel–Bulkley
velocity field measurements show that a plug flow develops within the plume thermal anomaly, there-
Plumes
fore producing a rising finger-shape. Moreover, light scattering highlights the development of a damaged
zone prior to the plume onset, and the peculiar structure of the gel around the plume as it rises. This
brings new insights into the solid–liquid transition of soft gels.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction Here, we investigate experimentally the development of


thermal instabilities out of a localized heat source in a yield stress
Although non-Newtonian fluids are common in natural and shear-thinning fluid (Carbopol solutions). This set up has been
industrial systems undergoing thermal convection (i.e. food or extensively used to study laminar Newtonian starting plumes
glass processing, lava lakes, planetary mantles, etc.), the occur- [18–23]. Combined Particle Image Velocimetry (PIV) and Thermo-
rence and characteristics of thermal instabilities in such fluids chromic Liquid Crystals (TLC) visualizations allowed us to measure
are still poorly documented and understood. Part of the difficulty simultaneously the velocity and temperature fields [24]. As the
is that the viscosity of these fluids approaches infinity when the heating power and the rheology of the experimental fluid were
shear rate vanishes. Hence, from a theoretical point of view, con- varied, different convective regimes and instability morphologies
vective instabilities cannot grow from a static conductive state were observed. We focus here on their phenomenology and condi-
submitted to infinitesimal perturbations [1–4], contrary to the case tions of existence.
of Rayleigh–Benard convection in Newtonian fluids [5]. In laponite,
where the non-Newtonian effects increase as the fluid ages, con- 2. Experimental setup and fluids
vection is killed after a while [6]. Recent qualitative experiments
in constant yield stress fluids show that thermal convection only 2.1. Setup
started with the help of spurious bubbles, forced mixing or temper-
ature heterogeneities at the top free surface [3]. When we The heat source consists of a Peltier element covered by a cop-
performed similar experiments with well controlled temperature per disk of thickness d = 4 mm and radius R = 12.5 mm. It is placed
conditions and no impurities in the fluid, an unstable linear con- at the center of a square plexiglas tank 20  20  30 cm3 (Fig. 1a).
ductive temperature gradient developed but motion never occured The heater is fixed directly on the bottom surface (Fig. 1b). The
(even after three weeks). Most quantitative studies have therefore upper fluid surface is free but there is a plexiglas lid on the tank
been limited to configurations where the onset of any motion to avoid evaporation and temperature fluctuations induced by
would not require a critical threshold if the fluid was Newtonian: the room. The room is kept at 21.0 ± 1.0 °C (the air-conditioning
creeping motion of solid spheres [7–9] and bubbles [10,11], has a 15–20 min period) and the tank walls are 3 cm thick. This al-
Saffman–Taylor instabilities [12], Poiseuille flow [13,14] or lows to filter the room temperature oscillations and to ensure a
Rayleigh–Poiseuille instabilities [15–17]. quiescent isothermal initial state in the tank. At time t = 0, a con-
stant power Pelec is applied to the heater. The ambient temperature
and the temperature of the heater surface are measured by
⇑ Corresponding author. Tel.: +33 1 69 15 80 56; fax: +33 1 69 15 80 60. thermocouples and monitored through time, as well as the electric
E-mail address: davaille@fast.u-psud.fr (A. Davaille). current voltage and intensity. Because of the non-linear response

0377-0257/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jnnfm.2012.10.008
A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153 145

The TLCs were added once the tank is filled. For each fluid, the
Laser 532 µm
density and thermal expansion coefficient have been measured as
(a) a function of temperature between 15 and 65 °C with a Anton Paar
Optic fiber DMA5000. Given the very small quantity of carbopol added, the
other physical properties are those of the water–glycerin mixtures
[25].
The rheological behavior of the fluids has been measured as a
function of shear rate c_ and temperature with a RS600 (Thermo-
Haake) and a MCR501 (Anton Paar). All measurements lead to a
description of the fluid rheology with a Herschel-Bulkley model
Cylindrical [26–28] (Fig. 2):
lense r ¼ r0 þ K v c_ n ð1Þ
where r is the stress. The values of c_ typically reached with the
rheometer ranged between 104 and 102 s1. The rheological
parameters values given in Table 1 were obtained by fitting the
Heater
descending part of the flow curve measured using a Couette geom-
etry. Because the stresses generated by thermal buoyancy are small
(see next section), we have to use solutions with very low yield
CCD camera stresses, which are delicate to measure. The existence of a real yield
stress at very low values of c_ is even still debated (e.g. [29,26,30]).
(b) Overall, our flow curves become more noisy when c_ < 102 s1,
Copper which could be caused by wall slip and/or transitional effects (e.g.
Peltier element [26]). We therefore kept only the data which was reproducible,
i.e. when c_ > 102 s1. Measurements on the same fluid, but with
Aluminium different apparatus, lead to variations in yield stress r0 of 30%. This
Tank bottom large uncertainty is due to the low values of the yield stress. The
Altuglass exponent n is much more reliable, and varies between 0.50 and
Thermocouples
0.63 (Table 1), which is in the range found in the literature
Fig. 1. (a) Experimental set up. The heater is placed on the center of the bottom [29,25,31]. r0 and n do not show any significant variations with
plate of the tank. At time t = 0, the heater is switched on, and a thermal plume can temperature [25], while the consistency Kv decreases by a factor
develop. Here is shown an image of the thermal plumes obtained in Newtonian of 2 for a temperature increase of 20 °C.
glucose syrups (e.g. [23]). The bright lines represent isotherms (see text). (b) Close
Table 1 summarizes the rheological parameters and the thermal
up on the heater configuration. The cold junction of the Peltier element is placed on
an metallic plate to dissipate the cold, while the hot junction is in contact with the properties of the various studied fluids. The addition of glycerin re-
copper disk. We verified that the temperature on the copper disk was homogeneous sults in increasing the coefficient of thermal expansion, and in
[22] and that the flow was not perturbed by the thermocouples. reducing the yield stress and the consistency compared to water
with a similar quantity of carbopol.

of the Peltier device, the thermal power P delivered into the fluid is 2.3. Parameters
not equal to the electric power supplied to the Peltier [23].
However, for each experiment, an estimate of the thermal power Given our experimental set up and the rheology of the fluid,
actually delivered to the fluid can be computed using the Peltier several parameters are needed to characterize the system (e.g.
element characteristics (see Annex 1). [1,3]):
The fluid is seeded with three types of encapsulated thermo-
chromic liquid crystals (TLC), each reflecting light at a different – the box aspect ratio L/H, where H is the fluid height in the
temperature. A vertical cross-section of the tank is illuminated tank and L its width. Here L/H ranged between 0.83 and
by a 532 nm laser sheet, and images are recorded every second 1.21 depending on the runs.
using a CCD camera (Fig. 1). Depending on the temperature field, – the size aspect ratio between the heater and the box R/H,
several bright lines can be seen on the images (Fig. 1), each of which varied between 13.3 and 19.4 depending on the
which represents a different ‘‘isotherm’’ (for details on the method runs.
see [23]). In our aqueous mixtures of carbopol, the TLCs sometimes – the Reynolds number can be estimated using (e.g. [11]):
also aggregate to form ‘‘particles’’ which scatter light, even without
a temperature field (e.g. Figs. 5 and 6). We therefore use these qRn W 2n
Re ¼ ð2Þ
aggregates as tracer particles and calculate velocity fields using Kv
Particle Image Velocimetry (PIV package from LaVision). where W is the velocity. It never exceeds 1 mm/s in our experi-
ments, so that Re < 0.01 and the instabilities are always in the creep
regime.
2.2. Fluids – the Bingham number Bi compares the yield and viscous
stresses in (1):
The fluids are mixture of carbopol (Noveon ETD2050), distilled r0
water and glycerin (between 0 and 56 wt%, see Table 1). The Bi ¼ ð3Þ
K v c_ n
Carbopol powder (0.4–0.8 g/L) was dispersed in the water-glycerin
solvent for 2 days using a magnetic agitator. Then the solution was When motions are observed, the measured typical shear rates range
neutralized with a sodium hydroxide solution (5 M). It was then between 106 and 0.05 s1. This implies Bi between 0.6 and 1000.
left at rest for 8 days before use to ensure that the gel hydration – the yield parameter Y0 compares the yield stress to the
was complete. stress induced by the thermal density anomaly rth:
146 A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153

Table 1
Fluid properties at 20 °C. Rheological properties: r0 yield stress, n shear-thinning index, Kv consistency. Thermal properties at 20 °C: q density, a thermal expansion, thermal
conductivity k.

Fluid Glycerol (wt%) carbopol (g/l) H (mm) r0 (Pa) n (–) Kv (Pa sn) q0 (kg/m3) a (104 K1) k (W/m K)
J1 56.5 0.4 180 0.037 0.63 0.345 1144 4.69 0.40
M2 0 0.8 180 0.35 0.5 1.57 999 2.10 0.58
M4 51.0 0.6 155 0.16 0.5 0.86 1129 4.89 0.41
J5 56.0 0.65 140 0.15 0.60 0.535 1142 4.61 0.40
J6 56.0 0.625 187 0.106 0.54 0.855 1142 4.61 0.40
S7 56.0 0.6 166 0.10 0.54 0.76 1142 4.62 0.40

1
10 80

PLUME

(Pa)
70

th
60
SHEAR STRESS (Pa)

THERMAL STRESS
0
10 50

40 CELL

30 b

10
-1 20 a
NO MOTION
10

0 0.05 01. 0.15 0.2 0.25 0.3 0.35


10
-2 YIELD STRESS 0 (Pa)
-3 -1 1
10 10 10
Fig. 3. Phase diagram as a function of the yield stress r0 and of the stress of thermal
SHEAR RATE (sec-1) origin rth. Filled symbols stand for the plumes, empty ones for convection confined
around the heater, and crosses for no convection. The solid line represents Yc2 = 260
Fig. 2. Flow curves measured for three fluids, J1 (red stars), J5 (green triangles) and
and the dotted line Yc1 = 120. Stars stand for fluid J1, disks for S7, diamonds for J6,
S7 (magenta disks) and their fits by a Herschel-Bulkley law (solid lines). For the fits, triangles for J5, and squares for M4. The empty black rectangle in the top left corner
only the measurements for shear rates above 102 s1 have been considered (see shows the uncertainty on the stresses. ‘‘a’’ points toward an experiment without
text). (For interpretation of the references to color in this figure legend, the reader is any plume even after four days while experiment ‘‘b’’ generated a plume in less
referred to the web version of this article.) than 30 min.

rth aqgP
Y0 ¼ ¼ ð4Þ 0.5
r0 r0 k
since for a constant power supply, the hot pocket of fluid applies 0.45 120
under the action of gravity a stress which scales as rth  aqg P/k.
p

In a yield stress fluid initially at rest, the hot pocket will be able 105
POISSON RATIO

0.4
to rise only if rth is greater than the yield stress (e.g.
[7,10,11,32,3]). So, Y0 is the key parameter to determine the onset 135
of convection. 0.35
120
Owing to the small values of the thermal expansion coefficient, 0.3
and to the maximum temperature difference that the system can
encounter without generating bubbles, the thermally induced 0.25
stresses are small. Therefore, convective instabilities are expected 120
only for rather small yield stresses (Table 1). Using 6 different flu-
0.2
ids (see Table 1) and varying the electric power input between 0.02
0.5 1 1.5 2 2.5 3
and 6.5 W, we run 62 experiments with rth ranging from 1.5 to 80,
leading to yield parameters between 20 and 1600 (Fig. 3). The typ- ELASTIC MODULUS (Pa)
ical duration of an experiment ranged between 30 minutes and
Fig. 4. Value of Yc1 predicted by Eq. (6) as a function of the elastic modulus and the
four days. Poisson ratio, for three different values of C0 (or set ups) and q = 1142 kg/m3. The
heavy gray lines represents C0 = 1, the thin gray lines C0 = 0.5, and the thick black
lines C0 = 0.2. The numbers indicate the Yc1 values of 120 ± 15.
3. Three different regimes

Once the heat source is turned on, a pocket of hot fluid first
grows by diffusion around the heater (Fig. 6a). Then depending moderate Y0 (Yc1 < Y0 < Yc2, Yc2 = 260(±20)), small scale convection
on the applied thermal stress and the yield stress of the fluid, three cells remain confined around the heater (Fig. 5). For Y0 > Yc2,
different regimes occur. Fig. 3 shows that the transition between plumes develop (Fig. 6). The transition at Y0 = Yc2 is quite sharp:
the different regimes depends primarily on the yield parameter. on Fig. 3, experiment ‘‘a’’ did not show any plume even after four
For Y0 smaller than Yc1 = 120(±15), no motion was detected. For days while experiment ‘‘b’’ generated a plume in less than 30 min.
A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153 147

3.1. No motion where mp is the Poisson ratio, EY = 2(1 + mp)G0 is the Young0 s modu-
lus, G0 is the elastic modulus and C0 is a constant which depends
For low thermal powers, no motion is recorded: a drop of dye on the geometry of the set up.
will remain at its initial position, and the presence of rising bubbles As long as rTE remains smaller than the yield stress, no viscous
neither starts the flow nor perturbs the temperature field. A steady flow can occur. The first critical value of the yield parameter Yc1 there-
state temperature structure develops by conduction and the heater fore corresponds to rTE  r0, which according to (4) and (5) gives:
temperature compared to the initial fluid temperature scales as
DT = P/4kReff [35, p. 215]. Given the geometry of the heater
q:g:ðR þ dÞ 2ð1  2mp Þ
Y c1  : ð6Þ
(Fig. 1b), we can take Reff  R + d. The fluid behaves as a solid and C 0 :G0 ð1 þ mp Þ
due to thermoelasticity, the temperature gradient can produce Eq. (6) shows that the first transition Yc1 depends on the fluid prop-
elastic deformations. Until the heater has reached steady state, erties and the set up geometry. We do not see variations in the
the thermoelastic stresses rTE will build up. According to linear experiments presented here because we are always using the same
thermoelasticity [36], rTE should scale as: set up (same R, d and C0) and fluids which are quite similar. We
a  EY were not able to measure the elastic modulus at the time of the
rTE ¼ C 0  DT  ð5Þ experiments, but since then, rheometry measurements performed
1  2mp
on similar solutions (Massmeyer et al., in prep.) gave us elastic

(a) (b)

24.6

(c) (d)

31.6
24.6 24.6

(e) (f)
f)

24.1

24.6

25.1
31.1

39
40 32.1

Fig. 5. Images for the cellular regime (fluid S7 at P = 1.85 W, Y0 = 247). (a) t = 0. Two dyed drops of fluid have been outlined in orange and green. (b) t = 2000 s. The isotherms
(values indicated in black) appear as bright linear zones and their thickness reflects the local temperature gradient; (c) t = 6000 s; (d) t = 12,000 s ; (e) t = 3.2  105 s.
Superimposed is the analytical solution for the steady state in pure conduction [35]. The values of the isotherms are indicated in °C. The orange drop is now stretched in an
horizontal bar, and the green drop has been stretched in a filament that we cannot distinguish from the isotherm anymore. (f) Velocity field obtained for (e) with PIV (cross-
correlation of images 250 s apart). The maximum velocity is 1.3  106 m/s. The fluid colder than isotherm 24.1 °C remains unyielded. Significant velocity remains confined
around the heater. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
148 A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153

modulus estimates between 0.5 and 3 Pa, similar to values obtained and the particles record no motion (t < 200 s on Figs. 7 and 9). Then
in pure water–carbopol solutions [28,37]. The Poisson ratio for Car- slow motions can be detected within the hot fluid pocket for
bopol solutions is unknown, but we expect it to be close to an 200 < t < 500 s, although the growth of the latter is still dominated
incompressible liquid, i.e. between 0.4 and 0.5. Fig. 4 then shows by heat conduction (Fig. 6a). It is impossible to determine exactly
that the predicted values of Yc1 are close to the observed ones when the transition from no motion to the small convective cell oc-
(120 ± 15) for constant C0 between 0.2 and 1. curs since the detection of very low velocities requires correlation
between images 50–250 s apart! However the transition to the
3.2. Cellular convection plume formation is very easy to spot. For t > 500 s, a finger of hot
fluid emerges from the hot pocket (Fig. 6b–d). It rises in two
Fig. 5 shows the convective pattern obtained for intermediate phases, quickly at first, and then at a steady slower pace (Figs. 7
Y0: after 3 h of experiment, the temperature structure around the and 8). Close inspection of the isotherms inside the plume shows
heater has reached its steady state (Fig. 5b–e) for the isotherms that their heights oscillate through time before becoming constant.
631 °C. The isotherms closer to ambient temperature take longer And so is the velocity field (Fig. 8). Nevertheless, in these experi-
to stabilize (about a day) but no plume will ever develop (even ments, all the plumes eventually reach the surface.
after four days, Fig. 5e). However, on 30 min time-scale, very slow
3.3.1. Transition from cellular convection to plume
particle motions (typically 2–3 orders of magnitude slower than in
The departure of the hot instability is quite sudden: it
the plume cases) are detectable within the hot pocket around the
corresponds to the fast uplift of the isotherms on the vertical
heater (Fig. 5f). Dyed areas in the fluid are continuously deformed
spatio-temporal diagram (Fig. 7) and to the necking of the
under the action of the viscous flow as seen on Fig. 5. This is radi-
isotherms close to the heat source on the horizontal spatio-
cally different from their behavior in the previous elastic regime,
temporal diagrams (arrows on Fig. 9c and d). It also coincides
and shows that the velocity we computed indeed represents mo-
with the maximum temperature recorded on the heater
tion and not elastic deformation. The very low velocity results in
(Fig. 8a) since heat is suddenly removed from the heater surface
a very low Peclet number (Pe = W  R/j  0.1), which implies that
when the plume starts. As it is the thermal power which is kept
advection is negligible compared to diffusion of heat in the conser-
constant and not the heater temperature, the latter decreases in
vation of energy equation. Therefore, only solving the heat conduc-
response to motion. We define here the instability ‘‘onset time’’ tc
tion equation (ignoring the advection) predicts quite well the
as the necking time. Other definitions based on the isotherms
thermal structure (Fig. 5e): only the 24.6 °C isotherm is broader
uplift or on the heater temperature give slightly different
than predicted by the analytical solution, which is compatible with
measurements (variations of less than 3%) but the overall trend
the temperature homogeneization in the core of a convective cell.
remains unchanged.
Fig. 10 shows the onset time as a function of Y0  Yc2. As ex-
3.3. Plumes development pected, the closer to the Yc2 transition, the larger the onset time.
For each fluid, the data points can be fitted with a power law
Figs. 6–9 show the development of a hot instability when
Y0 > Yc2. First, the hot pocket grows by diffusion around the heater tc ¼ a  ðY 0  Y c2 Þb ð7Þ

(a) (b) (c) (d)

24.6
24.6 31.6
24.6

25 mm

(e) (f) (g) 24.6


(h) 24.6

24.6
31.6
31.6 31.6

24.6

Fig. 6. Development of a plume (fluid S7 at P = 4.15 W, Y0 = 554). The bright lines are the TLCs isotherms: (a) t = 300 s; (b) t = 532 s. The colored arrows show the vertical
component of the velocity field Wz. They are superimposed on the raw image. The same color scale has been applied for images (b), (e) and (f). The maximum Wz (in white) is
0.38 mm/s; (c) t = 560 s; (d) t = 600 s; (e) t = 652 s; (f) t = 1140 s; (g) t = 1700 s; (h) t = 2200 s. The velocity fields were calculated with PIV using cross-correlation of images 5 s
(b) to 1 s (c–h) apart.
A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153 149

(a) 160
(a)
(a) (b)(c) (e) (f) (g) (h) T 0.2
40
(d)
HEIGHT (mm)

120

20 0
80 d T/dt

d
-0.2
40 0
tc
(b)
(b)

(10-1 mm/sec)
6

max(Wz)
160
4
24.6
HEIGHT (mm)

120
2
80
0
(c) tc
40
1

HEIGHT (mm)
150
0

Wz (axis)
100
(c) 160
0.5

50
120
HEIGHT (mm)

0
24.6
80 tc
(d) 1
40 150 z

dWr/dr (axis)
HEIGHT (mm)
Wz

0 100 Wr
0
500 1000 1500 2000
TIME (s) 50
r

Fig. 7. Development of a plume (same as Fig. 6). Light intensity of the pixels line: –1
(a) along the plume axis (r = 0) as a function of time and height. The isotherms have 0
0 200 400 600 800 1000 1200
been highlighted in red (24.6 °C), blue (31.5 °C) and yellow (39.5 °C). We can also
follow the particles as they rise along the axis. After a stage where the heat transfer TIME (s)
in the fluid is mainly by conduction, the plume rises quickly around 550 s. The
white vertical bars correspond to the snapshots of Fig. 6 and the black horizontal Fig. 8. Development of a plume (same as Fig. 6). (a) Temperature difference DT
bars to the horizontal cross-sections in Fig. 9. (b) vertically above the heater edge between the Peltier heater and the bulk fluid (red curve) and its time-derivative as a
(r = 12.5 mm). (c) Vertically above r = 25 mm. The numbers in black indicate the function of time. (b) maximum vertical velocity Wz recorded on the plume axis. (c)
isotherm value. (For interpretation of the references to color in this figure legend, Vertical velocity along the plume axis normalized by its maximum at each time, as a
the reader is referred to the web version of this article.) function of depth and time. (d) dWr/dr along the plume axis normalized by its
maximum at each time, as a function of depth and time. Its maximum corresponds
to the stagnation point at the top of the plume (inset). The two vertical dashed lines
show the onset time and the second uplift event (cf. text). (For interpretation of the
but we could not find any rationale between the coefficients a and b
references to color in this figure legend, the reader is referred to the web version of
and the physical properties of the fluids. this article.)
Yc2 is much higher than the one reported for sedimentation of
solid spheres (Yc = 6.85 [7,32]) or bubbles [10,11]. However, in
the plume completely stops and cools down. Then a second pulse
the latter case, the yield number is defined as Y = gDDq/3r0,
is registered. We could distinguish up to three pulses before the
where D is the sphere diameter, and Dq its density excess. In our
steady-state is reached with time lapse between two pulses rang-
case, buoyancy is produced by the temperature differences and
ing from 140 to 1000 s. Given the long time-scales involved, this is
Dq = aqDTav. Since we know the temperature field for each
not due to the elasticity of the fluid (as seen in falling motions of
experiment, we therefore can calculate at each time the averaged
solid spheres for example, i.e. [32]), but it is due to instabilities
temperature excess DTav contained in the hot pocket of fluid
in the plume conduit. Such instabilities have already been reported
around the heater. We can also calculate the effective radius of a
for Newtonian fluids (e.g. [33,22,23]) and shear thinning fluids (e.g.
sphere that would have the same volume, and so we finally can
[34]) although their origin remains unknown. Here, we notice that
estimate Y. To define what constitutes the ‘‘hot pocket’’, we choose
the velocity of the first pulse is much higher than the steady state
all the fluid with a temperature excess greater than 0.1DT. Then
velocity in the plume conduit, and that the amplitude of the veloc-
the boundary between the cellular convection and the plume for-
ity variations is decreasing through time (Fig. 8b). The steady state
mation occurs for Y = 8.8 ± 0.7. This value now is of the same order
represents the hot flow that the heat source is able to support con-
as Yc. That the two values are different is expected since we have a
tinuously. Since it needs to penetrate unyielded material, the first
bottom rigid boundary condition instead of a free surface (as in the
pulse is the most buoyant (i.e. the hottest) and carries more buoy-
experiments for the sphere) and since the geometry of the rising
ancy than what is produced continuously. Therefore the plume
plume is different from a sphere, indeed, Yc probably depends on
head starts draining the source and almost detaches from it. How-
the geometry [10,11].
ever, as it rises, it looses its heat by conduction (i.e. [39]), until its
local buoyancy cannot overcome anymore the yield stress, which
3.3.2. Episodicity explains its strong slow down, or even arrest (Fig. 11). Meanwhile,
We have already mentioned the strong time-dependence of the the thermal boundary layer around the heater is building again
plumes, whereby the temperature and velocity fields fluctuate be- until a new pulse starts. In the reference frame of the rising plume,
fore reaching steady state (Figs. 7 and 8). In some cases (Fig. 11), its tip is characterized by a stagnation point where dWr/dr is
150 A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153

0
(a) 140
RADIUS (mm)

24.6

HEIGHT (mm)
-50 100

(a) 60
0
RADIUS (mm)

24.6 20
-50
31.6

0 1000 2000 3000


(b) TIME (s)
140
0 (b)

VERTICAL VELOCITY
RADIUS (mm) RADIUS (mm)

1
24.6

HEIGHT (mm)
39.5 0.8
-50 31.6
0.6
0.4
(c) 0.2
0
0
24.6
0
-50
(d) 2000 3000
0 500 1000 1500 2000
TIME (s)
(c) 140
1

HEIGHT (mm)
Fig. 9. Development of a plume (same as Fig. 6). Light intensity of the pixel line 0.5

dWr / dr
along a radius as a function of time at four different depths z: (a) z = 102.1 mm; (b)
z = 66.5 mm; (c) z = 12.1 mm; (d) z = 3.2 mm. The black arrows indicate the 0
‘‘necking’’ time used to define the plume onset time (see text). The insets show
where is the cross-section relative to the tank. -0.5

104
b=-0.54 0
2000 3000
b=-0.55 TIME (s)
b=-0.81
ONSET TIME tc (s)

b=-0.62 Fig. 11. Strong plume episodicity (fluid J5 at P = 3.9 W, Y0 = 333). (a) Light intensity
of the pixel line along the plume axis (r = 0) as a function of time and height. The
b=-1.12 isotherms have been highlighted in red (24.6 °C), blue (31.5 °C) and yellow
(39.5 °C). (b) Vertical velocity along the plume axis normalized by its maximum
103
at each time, as a function of depth and time. (c) dWr/dr along the plume axis
normalized by its maximum at each time, as a function of depth and time. After the
conduction stage, the plume rises quickly at 2200 s until it stops completely and
cools down. Then a second pulse is recorded, starting around 3100 s. (For
interpretation of the references to color in this figure legend, the reader is referred
to the web version of this article.)

102
1 10 102 103 104 Fig. 1), while cold more viscous plumes are finger-shaped
Y0 [18,19]. The morphology of the hot and less viscous plumes ob-
c2
served in our experiments is therefore significantly different from
Fig. 10. Plume dimensionless onset time as a function of the yield parameter. The the Newtonian case. On the other hand, intrusion of fluid in elastic
same symbols as in Fig. 3 have been used for the different fluids. To each fluid (and matrices like gelatin or rocks are known to produce columnar or 2-
set of rheological parameters), corresponds one of the colored lines. They D dikes shapes (i.e. [38]).
correspond to Eq. (7) and their exponent b is given in the inset. A close inspection of the velocity field (Fig. 6e and f) reveals that
maximum (Fig. 8d, [23]). We see clearly on Fig. 8c and Fig. 11c that the upwelling material is entirely confined within the thermal
the second pulse also bears this characteristic signature and starts anomaly. Moreover, the velocity radial profiles around the plume
from the bottom of the tank. However, as it is following the conduit axis are quite flat (Fig. 6f). As a measure of the shear rate, we can
previously built, it needs to carry less buoyancy than the first pulse, calculate the second invariant of the strain rate tensor from the
therefore its velocity is smaller, closer to what the heat source can velocity field (Wr, Wz). For an axisymmetric configuration, it writes:
sustain continuously. The system therefore will approach steady " 2  2  2 #1=2
state by a series of pulses decreasing in amplitude. dW r W2 dW z 1 dW z dW r
c_ 0 ¼ þ 2r þ þ þ ð8Þ
dr r dz 2 dr dz
3.3.3. Plume morphology and plug flow
In Newtonian fluids, hot less viscous plumes entering a colder Fig. 12b shows the map of the characteristic shear rate for t = 2000 s:
and more viscous environment are mushroom-shaped (e.g. the shear rate in most of the plume axis region is zero (within the
A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153 151

uncertainties on the shear rate field). So the fluid there is not sheared nally the sudden rise of a hot finger, with strong shear localization.
but moves up as a plug. This shear localization is characteristic of a This behavior is reminiscent of the sequence of events reported in
yield stress fluid and is commonly observed in pipe flows (e.g. [14]). recent rheometry creep experiments using carbopol [41,31]. Those
Moreover, even when the velocity along the axis has reached its experiments were investigating the transition to fluidization of a
steady state, the plume conduit thermal anomaly enlarges slowly sample when a constant stress was applied. In our case of heating
with time (Fig. 9a and b): this is due to the progressive erosion by at constant power, the applied stresses, due to the rising tempera-
the plume flow of the solid part of the fluid. As seen in rheometry ture, are increasing through time till the system reaches steady
measurements [40,31], we observe the slow outward propagation state. However, in both systems, fluidization involves an episode
of the interface between the two phases (Fig. 6g-h, Fig. 9a and b). of strong shear localization and/or shear banding. Divoux et al.
However, a region of plug flow always remains along the plume axis [31] observed total slip on the wall of the apparatus, prior to the
(Fig. 12), and the plume shape remains columnar. establishment of steady shear, while Caton and Baravian [41]
suggested the presence of fractures.
In Figs. 7 and 9, horizontal bright lines represent particles at rest
4. From solid to flowing behavior
(i.e. unyielded areas), and continuous oblique lines moving particle
trajectories (yielded areas). However, in between the two zones,
The development of the thermal instabilities reported here al-
that is just before onset and around the plume thermal anomaly,
ways goes through three phases: first thermo-elastic deformation,
there exists a more chaotic zone, which propagates as the plume
then very slow creep where heat transfer is still dominated by
rises. Fig. 13 presents a zoom on the particles trajectories during
conduction and the Bingham number (Eq. (3)) remains high, and fi-

(a) VELOCITY (mm/s) (b) SHEAR RATE (1/s)

0.045

0.04

0.035

0.03

0.025

0.02

0.015

0.01

0.005

RADIUS (mm) RADIUS (mm)

(c) VISCOSITY (Pa.s) (d)


50.1 2.4

39.8
2.2

1.8

1.6

1.4

1.2

RADIUS (mm) RADIUS (mm)

Fig. 12. Vertical cross-section of the plume structure at t = 2000 s (same plume as Fig. 6). (a) Vertical velocity, (b) shear rate according to Eq. (8), (c) viscosity calculated with
g ¼ r=c_ 0 , (d) shear stress normalized by the yield stress. Dark blue areas are for shear rates below the detection level and are therefore probably unyielded. (For interpretation
of the references to color in this figure legend, the reader is referred to the web version of this article.)
152 A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153

University of Tokyo. B. Gueslin benefited from a post-doctoral fel-


(a) 55
lowship from CNRS/INSU. A. Massmeyer was supported by the Ini-
tial Training Network (ITN) Crystal2Plate, a FP7-funded Marie
HEIGHT (mm)

50
Curie Action under Grant Agreement Number PITN-GA-2008-
215353. This work was supported by CNRS/INSU through program
45
SEDIT, and grants BEGDY and PTECTO from the French ANR.
40
Appendix A. Determination of the thermal power delivered by
35 the Peltier heater
450 500 550 600 650 700
A Peltier element develops a temperature difference between
(b) its two surfaces, DT = Thot  Tcold, when it is fed with a current of
50 intensity I. The thermal power P delivered on the hot surface then
HEIGHT (mm)

writes:
45
1
P ¼ Sp  I  T hot þ I2  Rp  K p  DT ð9Þ
40 2
where Sp is the Seebeck coefficient, Rp is the electric resistance, and
35
Kp the thermal conductivity. Since we measure the temporal evolu-
450 500 550 600 650 700
tion of the voltage U and intensity I, we can determine for each
TIME (s) experiment Sp and Rp through:

U ¼ R p  I þ Sp  DT ð10Þ
Fig. 13. Development of a plume (fluid S7 at P = 4.15 W, Y0 = 554). Zoom on the
vertical spatio-temporal evolution off axis: (a) r = 12.5 mm, and (b) r = 25 mm. Note
and
that the offset is towards the top for (a) which means that the material is rising,
while it is downwards for (b), which means that the downwards return flow is close
U  Sp  DT
to the plume. Rp ¼ ð11Þ
I
At the end of each experiment, the intensity is switched off but be-
the stage of fast uplift of the plume: at the 0.1 mm scale, the trajec- cause the temperature difference takes some time to vanish, a resid-
tories of the particles around the hot thermal anomaly are not con- ual voltage is measured, which decreases through time, and which
tinuous but show sudden offsets, as if a damaged zone was is indeed proportional to DT. According to Eq. (A.2), this gives us a
developing around the plume. Studies of the microscopic structure measure of Sp, which is found to be constant through all the exper-
of Carbopol solutions show that they are concentrated suspensions iments: Sp = 0.023 V/K.
of swollen polymer ‘‘sponges’’ with typical size of 5–20 lm [26,37]. Then, using Eq. (A.3), the resistance is determined for each time.
So the fact that we are able to see at the 0.1–0.5 mm scale the It is found that Rp slightly varies with temperature, in agreement
intermittent sudden disruptions suggests that the transition from with the manufacturer.
solid to flowing behavior may occur through intermittent plastic
events correlated over a finite size area, as proposed by recent
References
models (e.g. [42]). However, we are lacking resolution, both in time
and space, to tell if those events are fractures or slip of one gel [1] J. Zhang, D. Vola, I.A. Frigaard, Yield stress effects on Rayleigh–Benard
particles aggregate onto another one. convection, J. Fluid Mech. 566 (2006) 389–419.
[2] V.S. Solomatov, A.C. Barr, Onset of convection in fluids with strongly
temperature-dependent, power-law viscosity: 2. Dependence on the initial
5. Conclusions perturbation, Phys. Earth Planet. Int. 165 (2007) 1–13.
[3] N.J. Balmorth, A.C. Rust, Weakly nonlinear viscoplastic convection, Non-Newt.
Fluid Mech. 158 (2009) 36–45.
We have conducted experiments on thermal instabilities in [4] A. Vikhansky, Thermal convection of a viscoplastic liquid with high Rayleigh
yield stress fluids. The yield parameter, the ratio of stresses of and Bingham numbers, Phys. Fluids 21 (2009) 103103.
thermal origin over the yield stress, is shown to be the key param- [5] S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability, Dover, 1961.
[6] L. Bellon, M. Gibert, R. Hernandez, Coupling between aging and convective
eter to predict the convective regime. Given the small value of the motion in a colloidal glass of Laponite, Eur. Phys. J. B 65 (2007) 101–107.
thermal expansion coefficient, leading to weak thermal stresses, [7] A.N. Beris, J.A. Tsamopoulos, R.C. Armstrong, R.A. Brown, Creeping motion of a
plumes will develop only for very low yield stresses and will sphere through a Bingham plastic, J. Fluid Mech. 158 (1985) 219–244.
[8] B. Gueslin, L. Talini, Y. Peysson, B. Herzhaft, C. Allain, Flow induced by a sphere
involve weak shear rates (62  102 s1). However, their morphol-
settling in an aging yield-stress fluid, Phys. Fluids 18 (2006) 103101.
ogy and dynamics are markedly different from those of a Newto- [9] A.M.V. Putz, T.I. Burghelea, I.A. Frigaard, D.M. Martinez, Settling of an isolated
nian plume, and unyielded areas coexist with flowing regions. spherical particle in a yield stress shear thinning fluid, Phys. Fluids 20 (2008)
033102.
Further work is now under way to quantify the characteristics of
[10] N. Dubash, I.A. Frigaard, Conditions for static bubbles in viscoplastic fluids,
plume dynamics, the damage zone extent and the transition from Phys. Fluids 16 (2004) 4319–4330.
solid to flowing behavior. [11] N. Dubash, I.A. Frigaard, Propagation and stopping of air bubbles in Carbopol
solutions, J. Non-Newt. Fluid Mech. 142 (2007) 123–134.
[12] P. Coussot, Saffman–Taylor instability in yield stress fluids, J. Fluid Mech. 380
Acknowledgments (1998) 363–376.
[13] I.A. Frigaard, S.D. Howison, I.J. Sobey, On the stability of Poiseuille flow of a
bingham fluid, J. Fluid Mech. 263 (1994) 133–150.
This work benefited from discussions with Veronique Lazarus,
[14] C. Gabard, J.-P. Hulin, Miscible displacement of non-Newtonian fluids in a
Neil Ribe, Kei Kurita, Fabien Mahaut and François Boulogne, and vertical tube, Eur. Phys. J. 21 (2003) 388–393.
help in the lab of Floriane Touitou, Alban Aubertin, Raphael Pidoux [15] I.A. Frigaard, C. Nouar, On three-dimensional linear stability of Poiseuille flow
and Christian Borget. The manuscript was improved thanks to two of bingham fluids, Phys. Fluids 15 (2003) 2843.
[16] D. Martinand, P. Carriere, P.A. Monkewitz, Three-dimensional global instability
anonymous reviewers. A. Davaille was able to treat the data, modes associated with a localized hot spot in Rayleigh–Benard–Poiseuille
thanks to a sojourn at the Earthquake Research Institute, convection, J. Fluid Mech. 551 (2006) 275–301.
A. Davaille et al. / Journal of Non-Newtonian Fluid Mechanics 193 (2013) 144–153 153

[17] C. Metivier, C. Nouar, On linear stability of Rayleigh–Bnard–Poiseuille flow of [29] G.P. Roberts, H.A. Barnes, New measurements of the flow-curves for Carbopol
viscoplastic fluids, Phys. Fluids 20 (2008) 104101. dispersions without slip artefacts, Rheol. Acta 40 (2001) 499–503.
[18] J.A. Whitehead Jr., D.S. Luther, Dynamics of laboratory diapir and plume [30] P.C.F. Moller, J. Mewis, D. Bonn, Yield stress and thixotropy: on the difficulty of
models, J. Geophys. Res. 80 (1975) 705–717. measuring yield stresses in practice, Soft Matter 2 (2006) 274–283.
[19] P. Olson, H. Singer, Creeping plumes, J. Fluid Mech. 158 (1985) 511–531. [31] T. Divoux, D. Tamarii, C. Barentin, S. Manneville, Transient shear banding in a
[20] I.H. Campbell, R.W. Griffiths, Stirring and structure in mantle starting plumes, simple yield stress fluid, Phys. Rev. Lett. 104 (2010) 20830.
Earth Planet. Sci. Lett. 99 (1990) 79–93. [32] H. Tabuteau, P. Coussot, J.R. de Bruyn, Drag force on a sphere in steady motion
[21] E. Kaminski, C. Jaupart, Laminar starting plumes in high-Prandtl-number through a yield-stress fluid, J. Rheol. 51 (2007) 125–137.
fluids, J. Fluid Mech. 478 (2003) 287–298. [33] P. Olson, G. Schubert, C. Anderson, Structure of axisymmetric plumes, J.
[22] J. Vatteville, P. van Keken, A. Limare, A. Davaille, Starting laminar plumes: Geophys. Res. 98 (1993) 6829–6844.
comparison of laboratory and numerical modeling, Geochem. Geophys. [34] P.E. van Keken, Evolution of starting mantle plumes: a comparison between
Geosyst. 10 (2009) Q12013, http://dx.doi.org/10.1029/2009GC002739. laboratory and numerical models, Earth Planet. Sci. Lett. 478 (1997) 1–14.
[23] A. Davaille, A. Limare, F. Touitou, I. Kumagai, J. Vatteville, Anatomy of a laminar [35] H.S. Carslaw, J.C. Jaeger, Conduction of Heat in Solids, Oxford Univ. Press, 1959.
starting thermal plume at high Prandtl number, Exp. Fluids, 2011. doi:http:// [36] L.D. Landau, E.M. Lifshitz, Theory of Elasticity, Pergamon Press, 1986.
dx.doi.org/10.1007/s00348-010-0924. [37] I.A. Gutowski, D. Lee, J.R. de Bruyn, B.J. Frisken, Scaling and mesostructure of
[24] A. Davaille, A. Limare, Laboratory studies on mantle convection, in: D. Carbopol dispersions, Rheol. Acta 51 (2012) 441–450.
Bercovici, G. Schubert (Eds.), Treatise of Geophysics, vol. 7, Elsevier, [38] T. Menand, S.R. Tait, A phenomenological model for precursor volcanic
Amsterdam, 2007, p. 89165. eruptions, Nature 411 (2001) 678–680.
[25] J. Peixinho, Contribution experimentale al etude de la convection thermique [39] G.K. Batchelor, Heat convection and buoyancy effects in fluids, Q. J. R. Met. Soc.
en regime laminaire, transitoire et turbulent pour un fluide a seuil en 80 (1954) 339–358.
ecoulement dans une conduite, PhD thesis, Universit Henri Poincare, 2004. [40] J.C. Baudez, P. Coussot, Abrupt transition from viscoelastic solidlike to
[26] J.M. Piau, Carbopolgels: elastoviscoplastic and slippery glasses made of liquidlike behavior in jammed materials, Phys. Rev. Lett. 93 (2004) 128–302.
individual swollen sponges: meso- and macroscopic properties, constitutive [41] F. Caton, C. Baravian, Plastic behavior of some yield stress fluids: from creep to
equations and scaling laws, J. Non-Newt. Fluid Mech. 144 (2007) 1–29. long-time yield, Rheol. Acta 47 (2008) 601–607.
[27] P. Coussot, L. Tocquer, C. Lanos, G. Ovarlez, Macroscopic vs. local rheology of [42] G. Picard, A. Ajdari, F. Lequeux, L. Bocquet, Slow flows of yield stress fluids:
yield stress fluids, J. Non-Newt. Fluid Mech. 158 (2009) 85–90. complex spatiotemporal behavior within a simple elastoplastic model, Phys.
[28] F.K. Oppong, J.R. de Bruyn, Microrheology and jamming in a yield-stress fluid, Rev. E 71 (2005) 010501.
Rheol. Acta 50 (2011) 317–326.

Das könnte Ihnen auch gefallen