Sunteți pe pagina 1din 12

Journal of Colloid and Interface Science 439 (2015) 42–53

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Regular Article

Does shaking increase the pressure inside a bottle of champagne?


A. Vreme a,b, B. Pouligny a, F. Nadal c, G. Liger-Belair b,⇑
a
Centre de Recherche Paul Pascal, CNRS, 115 avenue Schweitzer, 33600 Pessac, France
b
Equipe Effervescence (GSMA), UMR CNRS 7331, Université de Reims Champagne-Ardenne, BP 1039, 51687 Reims, France
c
Commissariat à l’Énergie Atomique, 33114 Le Barp, France

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

Article history: Colas, beers and sparkling wines are all concentrated solutions of carbon dioxide in aqueous solvents. Any
Received 28 July 2014 such carbonated liquid is ordinarily conditioned inside a closed bottle or a metal can as a liquid–gas 2-
Accepted 9 October 2014 phase system. At thermodynamic equilibrium, the partial pressure of carbon-dioxide in the gas phase
Available online 22 October 2014
and its concentration in the liquid are proportional (Henry’s law). In practical conditions and use (trans-
port, opening of the container, exterior temperature change, etc.), Henry’s equilibrium can be perturbed.
Keywords: The goal of this paper is to describe and understand how the system responds to such perturbations and
Henry’s equilibrium
evolves towards a new equilibrium state. Formally, we investigate the dynamics around Henry’s equilib-
Carbonated beverages
Molecular diffusion
rium of a closed system, through dedicated experiments and modeling. We focus on the response to a
Bubble dynamics sudden pressure change and to mechanical shaking (the latter point inspired the article’s title). Observa-
tions are rationalized through basic considerations including molecular diffusion, bubble dynamics
(based on Epstein–Plesset theory) and chemi-convective hydrodynamic instabilities.
Ó 2014 Elsevier Inc. All rights reserved.

1. Introduction In this article, we want to examine a few simple phenomena


that occur on the rather small scale of laboratory experiments, in
The capacity of carbon dioxide to get dissolved in aqueous volumes on the order of a liter. Most of our experimental work is
media is of paramount importance, in nature and industrial prod- done with pure water and CO2 but the study is relevant to common
ucts and productions. At large scale, CO2 fluxes between the atmo- situations involving carbonated beverages. Beyond the chemically
sphere and oceans are determinant in controlling the general most simple case of a water–carbon dioxide solution (model sys-
temperature of the planet, and solutions for storing huge amounts tem), our experiments and analysis equally apply to real sparkling
of the gas into aquifer reservoirs [1] are currently the matter of beverages such as champagne, beers and colas. The system under
intensive search. How much gas can be stored, in conditions of consideration is a closed bottle of water–carbon dioxide solution
constant temperature T, is ruled by the well-known Henry’s law, (either the model system or a real beverage). We address a few fre-
which states that the equilibrium concentration of dissolved CO2 quently asked and apparently simple questions such as: (i) How
is proportional to the partial pressure of gas P: fast the liquid inside can ingest or release carbon dioxide in a sit-
uation where the thermodynamic equilibrium between vapor
c ¼ kH P ð1Þ and liquid phases is perturbed? (ii) What is the consequence of
shaking the bottle on the pressure inside the bottle and the later
kH in Eq. (1) is Henry’s constant and c is CO2 mass concentration. By occurrence of effervescence and gushing, when the bottle is
definition, Henry’s law only applies to a static configuration, which, opened? As we will see, though the above questions may appear
strictly speaking, requires an infinite equilibration time to be real- childish, the involved phenomena may not be that simple, and
ized. In practical conditions, true equilibrium is not met, meaning some of the answers are not intuitive. It may come as a paradox
that the system is constantly at finite distance from a true equilib- to the reader that such questions have not been thoroughly exam-
rium configuration. Real situations, from global climatic phenom- ined and solved decades ago. Nevertheless and rather surprisingly,
ena down to micrometer-sized living cells, involve CO2 exchanges they have not been, as far as we can tell based on our literature
between phases, a gas phase and a liquid phase in simplest search, particularly in the field of champagne chemical-physics
situations. [2–6]. In this context, the reported research should also have a
didactical interest, beyond the field of sparkling beverages
⇑ Corresponding author. engineering.
E-mail address: gerard.liger-belair@univ-reims.fr (G. Liger-Belair).

http://dx.doi.org/10.1016/j.jcis.2014.10.008
0021-9797/Ó 2014 Elsevier Inc. All rights reserved.
A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53 43

The article is organized as follows. Section 2 is about materials


and methods. We explored the above questions with dedicated
experiments, with specially designed laboratory hardware, some
of them with real champagne bottles. In Section 3 we shortly
review basic concepts of gas solution equilibrium thermodynam-
ics, essentially Henry’s law for carbon dioxide in water. The section
provides quantitative bases on the amount of dissolved carbon
dioxide as a function of both pressure and temperature. Henry’s
law is illustrated in the case of a bottle of champagne that has been
left at rest and constant temperature. In Section 4 we address the
related question of how fast Henry’s equilibrium can be reached,
in practical conditions, and discuss the roles of molecular diffusion
and convection. We first explore the case of super-saturation, i.e.
when the gas pressure on top of the bottle is smaller than Henry’s
pressure. Next, we tackle the case of under-saturation. In both sit-
uations the perturbation is induced by applying a small pressure
step to the system. The objective is to assess the time it takes for
the pressure inside the bottle to come back to the new equilibrium
value. As we will see, the pressure profile measured for carbon
dioxide rejection observed after a negative pressure step (super-
saturated solution), is consistent with a purely diffusive model,
provided the whole system is kept at constant temperature. The
case of a positive pressure step (under-saturated solution) is more
difficult to interpret as the transient situation of an aqueous solu-
tion ingesting carbon dioxide through a free flat interface is
mechanically unstable [7].
Section 5 is dedicated to effects of shaking, starting from a solu-
tion that is initially close to equilibrium. Essentially we report
experiments where the pressure in the bottle neck is measured,
Fig. 1. Sketch of the experimental setups (a) aluminum container filled with water/
in parallel with observations of bubbles produced by the mechan- carbon dioxide solution. (b) Glass bottle filled with champagne. The cork has been
ical perturbation. Contrary to the belief that shaking increases the replaced by a home-made pneumatic connector and a small optical window has
pressure, we show that the main effect is a small transient pressure been fitted to the bottle neck (optical axis in dashed line).
drop. As we explain, the drop may be understood through simple
considerations about bubble kinetics, based on Epstein–Plesset
theory [8]. type). P and T signals are numerized using an acquisition amplifier
Our main results are summarized in the conclusion, together MX 840 A (HBM, Germany), and recorded on a PC computer with
with a few prospects. We end the paper with a short discussion Catman software.
on necessary conditions for the onset of effervescence when the The metal container has a capacity V T ¼ 1:27 L, with a circular
bottle is uncorked. Supplementary information is provided at the cross section S ¼ 63:6 cm2 . The total height (20 cm) is shared
end of the paper. Appendix A is a supplement to Section 4; essen- between a liquid phase of height hL , and a gas phase on top, of
tially the detailed mathematical resolution of the diffusion model height hG . Both heights were varied among different experimental
of pressure kinetics. Useful information on Epstein’s article runs.
about bubbles dynamics in water/gas solutions is reminded in Preparing a carbonated solution is straightforward, as we only
Appendix B. have to inject CO2 under a few bar pressure. While keeping the
solution under constant CO2 pressure (the intake valve remains
open), we just wait for the gas to get dissolved until equilibrium
2. Materials and methods is achieved. Purely diffusive dissolution is extremely slow, as we
discuss in Section 4. Not surprisingly, the process can be much
Experiments about pressure kinetics consist in recording the accelerated if the liquid is stirred, simply by vigorously shaking
pressure P inside a container as a function of time t, for instance the container. In our standard procedure, the container is located
when the bottle is subject to a sudden perturbation of Henry’s inside a water bath, whose temperature T 0 is controlled within
equilibrium. Such experiments can be made using a simple airtight 0.02 K. To stir the solution, the container is pulled out of the bath,
container. We used two different types of container/solution cou- shaken several times over a 12-h period, and the intake valve is
ple: most of the quantitative experiments have been performed closed. The system is then left at rest back in the bath for about
using a cylindrical metal tank (cf. Fig. 1a) filled with pure water/ 48 h. After this preparation, the solution is supposed to be
carbon dioxide binary solution. Some qualitative observations close to equilibrium, though it is never strictly so, as discussed in
has been done on a real glass champagne bottle (cf. Fig. 1b) filled Section 5.
with champagne (Trouillard Elexium). A flat optical window was In the case of the real system, the aqueous solution (cham-
fitted to the champagne bottle neck to view bubbles, see Fig. 1b. pagne) was already saturated. We had to replace the cork of the
In the following, the first system is referred to as model system bottle with a home made pneumatic device (see Fig. 1b), connected
and the second one is referred to as real system. to the same kind of pressure sensor as in the model system. The
The lid of the metal tank is equipped with three ports, for gas original cork was gently removed, to avoid foaming, and was
admission and purge, and for pressure and temperature measure- immediately replaced by the pneumatic device. The loss in carbon
ments, see Fig. 1a. The pressure sensor (Druck Unik 5000, GE, US) dioxide caused by the brief opening of the bottle was negligible, as
is attached to the exterior of the container, and the temperature we could check from the value of the pressure at rest, about 6 bar
is measured inside the gas phase by means of a thermocouple (K at room temperature. The volumes of gas and liquid phases inside
44 A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53

the champagne bottle were V G ¼ 32 cm3 and V L ¼ 750 cm3 , kH ðTÞRT


He ¼ ; ð5Þ
respectively. M
In principle, the experimental system should be perfectly gas-
proof. In reality the pneumatic technology used in our experiments M being the molar mass of carbon dioxide. Eq. (4) gives the pressure
had very small but measurable leaks, which caused drifts in pres- corresponding to Henry’s equilibrium as a function of the total
sure on the order of a millibar per hour, in typical conditions of amount of carbon dioxide in the container.
6 bar-CO2 pressure. The leak comes as a limitation in long-term Note that P depends on the temperature, directly and indirectly
observations, but, fortunately enough, has no visible influence on through the dependence of kH on T. Henry’s constant monoto-
the scale of few-hour recordings; which turned out enough for nously decreases with T following a Van’t Hoff-like equation:
our purposes, as we will see.
A few complementary experiments were dedicated to viewing   
0 DH D 1 1
fluid convection driven by CO2 dissolution, see Section 4. These kH ðTÞ ¼ kH exp   ; ð6Þ
experiments were carried out with a specially designed cubic con-
R T T0
tainer equipped with flat transparent windows, see Ref. [9] for
0
technical details. In this case a minute amount of fluorescein dye where kH is Henry’s constant of CO2 at T 0 ¼ 298 K, and DHD is the
(1.6  106 M L1) was added to the water solution as a marker dissolution enthalpy of CO2 in the liquid of interest (in J mol1).
of CO2 dissolution within the fluid. The method exploits the fact Illustrative values of parameters entering the Vant’ Hoff for-
that the fluorescence intensity of the dye strongly decreases when mula can be obtained from Agabaliantz data [13] about champagne
0 1
the solution pH becomes acid [10–12], as in the case of carbonated and other sparkling wines: kH ’ 1:21 g L1 bar ,
1
solutions [11,10,12]. The effect is very intense with fluorescein; DHD ¼ 24800 J mol . The temperature dependences of kH and P
CO2 rich regions show up as dark zones in the fluorescence image within a typical 75 cL champagne bottle is shown in Fig. 2 (data
of the solution, even under CO2 pressures as small as 0.1 bar. from ref [14]). The concentration of dissolved gas can be deduced
Examples of such images are given in Section 4.4. from Fig. 2 using Eq. (1). It is worth noting that the concentration
of dissolved carbon dioxide is only slightly temperature-depen-
3. Thermodynamic equilibrium dent, whereas the pressure is much more sensitive (about
200 mbar K1).
In this section, we address the ideal situation of thermodynamic
equilibrium of a carbonated solution. Though we deal with well-
known concepts, we want to set out basic relations prior to moving 4. Henry’s equilibrium recovery
to non equilibrium phenomena. We consider a closed container
partially filled with water, and derive the expression of the partial We now suppose that the equilibrium of the liquid–gas system,
pressure of carbon dioxide in the gaseous phase, given the total initially at pressure P 0 and at c0 ¼ kH P 0 concentration, is suddenly
number of CO2 molecules. We start from Eq. (1), where c is the perturbed by quickly changing the pressure to a new value,
mass concentration of dissolved CO2 molecules, P is the partial P1 ¼ P0 þ DP. The perturbation is achieved either by injecting or
pressure of carbon dioxide in the vapor phase. kH , the Henry’s con- removing a small amount of gas, through the purge-intake port
stant (usually in g L1 bar1) is a function of temperature, and of the container (see Fig. 1). According to whether DP > 0 or
depends on the particular composition of the liquid. DP < 0, the liquid becomes under- or super-saturated, respectively,
In the whole paper, we suppose that the gas phase is composed of and the system starts evolving towards a new Henry equilibrium.
carbon dioxide only, such that the total pressure P and the carbon We want to know the characteristics of the final state and how
dioxide partial pressure P CO2 are considered as identical (this is not long it takes for the system to reach it.
rigorously true since the gas phase is a mix of all the components In paragraph Section 4.1, we give the expression of the final
present in the liquid phase, each of them being at saturation pressure. Next, we present a simple diffusion-based model for
pressure). the pressure relaxation, within the simplifying assumption that
We consider now a closed cylindrical container with a volume there is no convection (see Section 4.2). In Section 4.3, results of
V G of gas on top of a volume V L of liquid, with a flat horizontal experiments concerning the dynamics of dissolution (or release)
interface, of surface S, in between. For simplicity, we suppose that of carbon dioxide in water are reported.
both volumes remain constant (we neglect the minute changes of
the liquid volume due to gas dissolution or rejection). The total
number of moles of carbon dioxide nT is a conserved quantity that 3.0
decomposes into nG moles in the gaseous phase and nL moles in the
8
liquid:
2.5
kH (g L−1 bar−1 )

nT ¼ nG þ nL : ð2Þ
7
PCO2 (bar)

In the pressure range of interest (a few bars), we may safely


suppose that equilibriums in gaseous phases are ruled by the ideal 2.0
gas law. Then:
6
PV G ¼ nG RT: ð3Þ
1.5
where R is the ideal gas constant (8.31 J K1 mol1). Combining Eqs.
5
(1)–(3), we obtain:
1.0
nT RT 5 10 15 20
P¼ ; ð4Þ
V G þ HeV L T (◦ C)
In the above equation, He stands for ‘‘Henry number’’, defined Fig. 2. Typical temperature dependences of kH and P within a 75 cL champagne
by: bottle. Pressure: solid line and left ordinate axis. Henry’s constant: dashed line and
right ordinate axis.
A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53 45

4.1. New equilibrium state ph2G ~2


h
s¼ 2
¼ ; ð14Þ
ðHeÞ D D
The final asymptotic value of the pressure after equilibrium
recovery is denoted by P 1 . Following the same reasoning as in Sec- We recall that, in Eq. (12), hG ¼ V G =S is the height of the gaseous
phase. The equivalent liquid height is defined as h ~ ¼ pffiffiffiffi
phG =He. Then
tion 3, we obtain (for t ! 1Þ:
s, given by (14), may be thought of as the time taken by CO2 to dif-
  1 ~
VL fuse in the liquid over the distance h.
P1 ¼ P0 þ DP 1 þ He : ð7Þ Let us put numbers in Eq. (14) to get some useful orders of mag-
VG
nitude. The Henry number He ¼ 0:83. With D ¼ 1:8  109 m2 s1
Note that, when the volume of liquid is much larger than the (see Refs. [15–18]), kH ’ 1:5  105 kg m3 Pa1 (see Refs.
volume of gas, the pressure returns back to its initial value [19,20]), and hG ’ 5 cm, we obtain ~ ’ 0:11 m
h and
(P 1 ¼ P 0 ). s ’ 6:3  106 s. We thus come out with a very long time (about
73 days!), but this number may be misleading in suggesting that
4.2. Dynamics of recovery the diffusive process is extremely slow. In fact, the square root
term in Eq. (12) is singular at t ¼ 0, corresponding to an infinite
We now elaborate a simple model based on the following time derivate. The pressure in fact is predicted to very rapidly vary
assumptions: at short time. The short time limit, according to (13), corresponds
to t 6 7  103 s.
(a) The gas phase is permanently at equilibrium, meaning that
the ideal gas law is verified at all times; 4.3. Response to decompression
(b) The temperature T is uniform and constant;
(c) Dissolution only proceeds by molecular diffusion (convec- In this paragraph, we examine the effects of a small negative
tive effects in the liquid phase will be discussed further on); pressure step (DP < 0), meaning that we decrease the pressure
(d) The growth time of the pressure step is much smaller than inside the container down to a prescribed value by opening the
any other characteristic time of the system (in other words, purge valve for a short while dt, around t ¼ 0, conventionally. We
the pressure step is achieved instantaneously on the time observe the response in P to this perturbation. In some of the
scale of diffusion); experiments we record both pressure and temperature responses,
(e) Saturation is immediately reached at the liquid–gas inter- see Fig. 3 for illustration. The figure indicates that the system
face. Consequently, Henry’s law is permanently satisfied at responds in two steps. The first one is a short-lived response, in
z ¼ 0, and the shift in CO2 mass concentration, Dc ¼ c  c0 , the form of negative peaks in both P and T, on the order of a minute
is related to the shift in pressure, DP ¼ P  P0 , by the in duration. In the second step, the temperature is back to the con-
equality stant value imposed by the thermostat (T0 = 20.35 °C), and the
pressure linearly increases as a function of t1=2 . Qualitatively, the
Dc ¼ kH DP on Sðz ¼ 0Þ: ð8Þ latter response is in line with the prediction of the diffusion model,
Dc obeys the classical one-dimensional (1D) linear diffusion see Eq. (12).
equation The drops in P and T are very likely due to the direct thermal
response of the gas phase to the decompression. If dt is short
@ t Dc ¼ D@ zz Dc; ð9Þ enough, the decompression is close to adiabatic. The resulting tem-
perature drop dT may be estimated as:
where D is the diffusivity of carbon dioxide in the liquid of interest.
The diffusion equation must be completed by the boundary condi-
tion (8) – assumption (e) – and a condition of vanishing flux at
the bottom of the tank:
P0 =5.4 bar
@ z Dc ¼ 0 at z ¼ hL ; ð10Þ
5.12
During the relaxation process, the molecules of carbon dioxide P2
diffuse through the interface from the gas phase into the liquid 5.10
(under-saturation, DP > 0) or in the opposite direction (over-satu-
P (bar)

ration, DP < 0). Because the system is closed, dissolution or rejec- 5.08
tion of CO2 directly modifies the pressure in the gas phase through δPe
a feed back process. The system formed by Eqs. (9), (8) and (10) is 5.06
closed when adding the conservation equation:
5.04
dDnG dDnL P1
¼ for t > 0þ : ð11Þ
dt dt 20.5
In Appendix A, we offer a full resolution of the above system, in
T (◦ C)

the limit of a semi-infinite liquid phase. Experimental conditions 20.0


δTe
are such that only the short time limit of the general solution is
19.5
of practical interest. We find:
h i 0 20 40 60 80 100
P ¼ P0 þ DP 1  ðt=sÞ1=2 ; ð12Þ t1/2 (s1/2 )

which is valid in the limit Fig. 3. Pressure and temperature response to a small negative pressure step
(experiment #6 in Table 1). P and T are plotted versus t1=2 for direct comparison
4  102 with the diffusion model, Eq. (12). The figure shows pressure values between 5.03
t6 s; ð13Þ and 5.13, to focus on details of the pressure response at ‘‘short times’’ (Eqs. (12) and
p2
(13)). The arrow on top indicates the initial pressure, P 0 ¼ 5:4 bar, which is well
In (12) and (13) s is a diffusion time, given by: above the represented interval.
46 A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53

"  ð1cÞ=c #
P1 0.05
dT ¼ T 0 1  ; ð15Þ
P0
0.04

with a concomitant pressure drop given by: 0.03

(P − P2 )/ΔP
"  ð1cÞ=c #
P1 0.02
dP ¼ P1 1 : ð16Þ
P0
0.01
Estimates from the above equations, with c ¼ 1:28, are in line
with the amplitudes of the short time responses in experiments 0.00
where decompression was operated ‘‘brutally’’, meaning that dt
was on the order of a few seconds. In the example shown in -0.01
Fig. 3, decompression was operated more gently, in about 2 min. 0.01 0.02 0.03 0.04
The process then was intermediate between adiabatic and isother- (t/τe )1/2
mal, and the drops in P and T are inferior to values (0.08 bar and
4 K, respectively) estimated with Eqs. (15), and (16). Fig. 4. Dimensionless pressure ðP  P 2 Þ=DP as a function of ðt=se Þ1=2 (se being the
An experimental characteristic time, denoted as se , can be experimental characteristic time), for the experiments listed in Table 1. Note that
the time has been rescaled using a different se for each experiment (cf. Table 1).
deduced from the slope of the linear response in Fig. 3. Because
of the short-lived thermal response, the amplitude of the pressure
perturbation to be retained for comparison with the diffusion The source of convection is most likely attributable to the ther-
model is not the pressure drop imposed by the experimenter, mal perturbation, which causes transient gradients of the fluid
DP ¼ P1  P0 , but the drop corrected for the thermal effect, density. Density gradients have a vertical component, which may
DP ¼ P2  P0 (see Fig. 3). Results from experiments carried out with be either stabilizing or destabilizing (as in the classical Rayleigh-
different values of the pressure drop are gathered in Table 1. The Bénard-Taylor problems [22,23]), and horizontal components.
table indicates the measured se next to the theoretical diffusion Because the bottle and the liquid inside have finite thermal con-
time calculated using Eq. (14), and the corresponding DP. Using ductivities, temperature variations (dT), whatever their origins,
the parameters listed in the table, we may transform the time cannot be uniform in horizontal sections of the system. This situa-
and the pressure into dimensionless variables, ðt=se Þ1=2 and tion, involving a horizontal temperature gradient, causes a hori-
ðP  P 2 Þ=DP, respectively. Results from the whole set of experi- zontal gradient in density of the liquid, which is unstable to
ments can then be gathered into a single graph, as shown in convection, whatever the amplitude and the sign of dT.
Fig. 4. Note that all data merge onto a unique curve, meaning that The amplitude of the convection may be hardly perceptible if
the experimental records for pressure relaxation are indeed well decompression has been operated ‘‘mildly’’, meaning that cooling
represented by Eq. (12). However the agreement is only qualitative has been minimal, resulting in very small gradients. Of course
because se in general is less than what we expect based on Eq. (14), the thermal response might be about eliminated if decompression
see Table 1. Quantitative agreement is met only for the smallest is operated very gently (dt ! 1). But the pressure step has to be
amplitude of the pressure drop and mild operation, as in Fig. 3. fast on the scale of the time limit set by Eq. (13) for the t 1=2
When the perturbation is large and sudden, the pressure response response to be observed. Due to the latter requirement, dt cannot
still seemingly follows the diffusion law, Eq. (12), but with an be larger than about 2 min. Completely eliminating convection in
anomalously short characteristic time. The anomaly suggests the real experiments may then turn elusive.
existence of a convective flow, that inevitably accelerates the gas
exchange through the interface. However, in the case of decom- 4.4. Response to compression – chemi-convection
pression, diffusion should not be destabilizing, essentially because
the density of carbonated water increases with the concentration We now turn to the response to a positive pressure step, mean-
of CO2 , see Ref. [21]. In response to the pressure drop, molecular ing that a finite amount of gas is injected through the intake valve
diffusion has the effect of decreasing the amount of CO2 in the of the container (Fig. 1). Experimental records of the pressure
layer below the interface. The density of this layer is then less than responses in this case show complex behaviors, far from the rather
that of the bulk solution underneath. Therefore the diffusive flux of simple picture of the former paragraph on decompression. Com-
CO2 , from the liquid to the gas phase on top, results in a density pression first heats the gas phase, through the same mechanism
profile that monotonously decreases from the bottom to the top as that described before, but with an opposite sign. Heating of
of the liquid. Such a profile is stable[22,23], meaning that it cannot the liquid through the interface now leads to lowering the density
promote convection. on top of the liquid phase. Adiabatic – or nearly so – compression
then has a stabilizing effect, and cannot cause convection.
In the compression experiment, carbon dioxide is added into
Table 1
Table of the characteristic time se measured for each experiment. The experimental
the system and progressively gets dissolved in the liquid phase.
and theoretical adiabatic pressure drops (denoted by dPe and dP resp.), together with The process starts by molecular diffusion, creating a diffuse layer
the expected theoretical time s and the experimental adiabatic temperature drop that is heavier than the fluid underneath. The resulting unstable
dT e ¼ T 1  T 0 are given for information. density profile is the source of convection [24,25], a process known
# DP (bar) se (s) s (s) dP e (bar) dP (bar) dT e (K) as ‘‘chemi-convection’’. Contrary to the case of decompression, the
source of chemi-convection is intrinsic to carbon dioxide dissolu-
1 0.092 2:3  10 6
4:6  106 0.0078 0.0217 –
tion, and then subsists as long as complete dissolution has not been
2 0.165 2:1  106 4:6  106 0.0349 0.0432 –
achieved [7,9]. In practical conditions, the heavy diffuse layer gets
3 0.250 1:9  106 4:6  106 0.0512 0.0643 –
destabilized within a few seconds; consequently the diffusive
4 0.106 9:6  106 9:64  106 0.0032 – 0.107
5 –
regime of pressure relaxation is not observable.
0.215 7:4  106 9:64  106 0.0026 0.122
6 – Convective motion of the fluid cannot be directly viewed
0.414 5:1  106 9:64  106 0.0021 0.113
through the opaque metal container, and even not in the cham-
A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53 47

pagne bottle, which has not enough optical quality. In this case we 5.05
used the special cubic container [9] mentionned in Section 2. We
observed a sample of water inside a parallelepipedic glass cell 5.04
(width a ¼ 70 mm, thickness b ¼ 5 mm, height c ¼ 50 mm) that
was exposed to positive pressure steps of carbon dioxide, from
0.2 bar to 0.8 bar in amplitude. The cell was illuminated by a laser 5.03

P (bar)
sheet (wavelength 514 nm), parallel to (a, c)-sides to excite the
fluorescent dye. 5.02
This type of experiment is limited to pure water as the initial
state of the liquid, and to volumes much inferior to a liter, but
5.01
has enough generality as it clearly shows the onset and evolution
of chemi-convection. Main features should be similar starting from
Henry equilibrium states with finite concentrations of CO2 and lar- 5.00 (a)
ger volumes. Fig. 5 shows a typical Rayleigh–Taylor instability that
70 80 90 100 110 120 130
has evolved into characteristic fingers, called ‘‘plumes’’. The dark
5.030 P0
features are CO2 -rich zones, while clear zones correspond to about P∞
pure water. The dark horizontal band on top of the photos is the

P (bar)
diffuse layer, which is strongly unstable, as we explained. Plumes ΔPs
5.025
are the source of general convection inside the cell, and convection
acts back on them, producing complex patterns, as the figure
illustrate.
(b)

5. Effects of shaking and opening 20.3

After the response to a small perturbation in pressure, we now T (◦ C)


want to investigate the effect of vigorously shaking the bottle con- 20.2
taining the carbonated solution. A popular belief is that the pres- (c)
sure increases when the bottle is shaken, and that this is the 70 80 90 100 110 120 130
reason for gushing of the liquid when the bottle is opened. We t (s)
show in this section that shaking in fact produces a small but oppo-
site effect, in the form of a transient pressure decrease. We start Fig. 6. Typical responses to shaking recorded with water in the aluminum tank
(model system). (a) Raw pressure signal; (b) filtered pressure signal; and (c)
with experimental observations and move to qualitative interpre-
temperature signal.
tations afterwards.

5.1. Experimental observations stands for ‘‘shaking’’), followed by a monotonous raise up of P over
about a minute, up to a plateau value denoted as P1 .
We tested the effect of shaking with champagne, sparkling Different experiments of the same type, with different contain-
white wine (‘‘Blanc Foussy’’), and simply water. The signals shown ers and liquids, revealed that P1 in general differs from P 0 , but the
in Fig. 6 were obtained with the metal container (Fig. 1) and car- difference may be of positive or negative sign. What seems to be a
bonated water, which was initially at equilibrium (as far as it could non reproducibility in the sign of the difference between the final
be, as we explained in Section 2) at about 5 bar of CO2 pressure. and initial pressures, is very likely due to the fact that the solution
Shaking was operated by hand for about 20 s, and the container is never completely at thermodynamic equilibrium. As the diffu-
was put back in the temperature controlled bath (’20 °C). sion time s is of the order of months, it is practically impossible
The graph in Fig. 6a features an oscillation regime followed by a to start from a real equilibrium (Henry) state. Shaking just helps
smooth evolution of the pressure. Oscillations are located within accelerating CO2 exchange between the gas and liquid phases. As
the period of active shaking of the container, between 70 and a result the solution after shaking is closer to Henry’s equilibrium.
90 s along the time axis. These oscillations are not of primary inter- As a corollary of the latter statement, repeating shaking only a
est as they are most probably a signature of the fluid dynamic pres- short while (less than one hour, say) after the first perturbation
sure which raises and drops in the course of shaking. Interestingly, produces about no change in pressure – i.e P 1 ¼ P 0 .
the oscillations can be eliminated by low-frequency filtering of the The graph in Fig. 6c shows a transient temperature decrease, in
signal, resulting in the graph shown in Fig. 6b. The graph reveals an parallel to the pressure transient. The experimental records then
overall decrease of the pressure, DP s ’ 10 mbar (the subscript s suggest that P and T transients have a common cause.

Fig. 5. Images of plumes and convective cells consecutive to positive pressure steps of (a) 0.2, (b) 0.4 and (c) 0.8 bar at t = 10, 7 and 4 s respectively. The horizontal dimension
of each image is 6 cm. Horizontal streaks are optical parasites due to imperfections in the windows of the cubic chamber. See text for details of experimental conditions and
method.
48 A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53

Shaking obviously generates bubbles. In the case of champagne,


P0 , VG,0 P0 , nG,1 , VG,1 P1 , VG,1
it was possible to observe bubbles inside the bottle, through the
nG,0
optical window that was fitted to the bottle neck (see Fig. 1b). VB
We thus could make videos of the bubbles a few centimeters below
the liquid–gas interface, within the liquid bulk. Since the device
only compensates for defects on the exterior interface of the bottle,
the optical quality is not optimal, but bubbles were well discern-
ible and their sizes could be measured from video images within
5 lm in uncertainty.
Within a few seconds after shaking, videos show huge amounts
of bubbles that very rapidly flow up across the images towards the
bottle top. Initially the motion is too fast with too many bubbles to
provide well resolved images (at 25 frames/s), but very large bub-
bles (larger than 100 lm in diameter) can be discerned. The
motion, together with the population of bubbles, definitely
decreases within about a minute, and largest bubbles disappear. N bubbles
After about 2 min, remaining bubbles tend to roughly the same
size, which we estimate 60 ± 5 lm in diameter. It takes about (a) (b) (c)
220 s for bubbles to completely disappear from video records.
Fig. 7. Sketch of the bubble collapse scenario (no buoyancy). (a) Before shaking. (b)
After shaking: N bubbles have been engulfed in the bulk. (c) After collapse. In (c),
V G;1 is the volume of the gas phase (white region + gray region); the gray region is
5.2. Dynamics of the bubble assembly the volume gain due to bubbles collapse (V B ).

The presence of bubbles in the champagne bottle suggests a


the pressure in the container after the bubble collapse (P 1 ). The
strong correlation between the pressure transient and bubble
amplitude of the pressure variation is:
dynamics. We now come to our proposition to explain the
observed transients. We start with a most simplified model, that VB
DPs ¼ P1  P0 ¼ 2P 0 ; ð17Þ
hopefully conveys the key ideas about the evolution of the bubble V G;0
population and its consequence on the pressure. This model is
Eq. (17) indicates that bubble collapse has a very sensitive
based on Epstein–Plesset theory for bubble dissolution [8]. In a sec-
impact on the pressure. Suppose e.g. V G;0 ¼ 100 mL, and
ond part we propose a less restrictive version of the model, that
P0 ¼ 6 bar. A very small amount of gas, e.g. V B ¼ 0:05 cm3 initially
takes into account the polydispersity of bubble sizes. This part ends
in the bubble phase is sufficient to produce 50 mbar as the ampli-
with a numerical resolution of the model.
tude DP s of the pressure drop.
The above reasoning was built on the assumption that all bub-
5.2.1. One-bubble-size model bles would collapse and contribute to the pressure decrease. We
Let us consider a single bubble in the liquid bulk. For simplicity, ignored the effects of buoyancy, essentially the rising up of bubbles
we first ignore the rising up motion of the bubble due to buoyancy due to the Archimedes force. Bubbles that reach the upper inter-
and suppose that it stays in bulk water. In the absence of capillary face before collapsing hardly contribute to the pressure decrease.
tension (and buoyancy), a bubble would be at Henry’s equilibrium Clearly very small bubbles collapse in very short times, losing the
and would stay stable. The effect of the interfacial tension (r) is to whole volume of their gas into water. These bubbles efficiently
increase the bubble internal pressure. As a consequence, the bub- contribute to the pressure drop as estimated by Eq. (17) and the
ble ‘‘sees’’ the surrounding liquid as under-saturated. Surface ten- effect lasts for a long time, because the dissolved gas can return
sion forces the gas content of the bubble to dissolve in the liquid. to the top gas phase only by advection/diffusion through the inter-
The initial state is unstable and the bubble has to decrease in size. face. Conversely, large bubbles survive much longer, leaving them
The process creates an excess of carbon dioxide concentration enough time to reach the top interface and return their gas content
around the bubble. This extra concentration progressively gets directly into the top phase.
diluted and, at long times, relaxes by diffusion across the macro- Let be sc the collapse time of a bubble of initial radius a0 ; sr the
scopic water/gas interface, bringing the system back to Henry’s time it takes for the same bubble to move up over the length hL
equilibrium. If we regard the top gas phase as a giant bubble (if (height of liquid), and aH 0 the limit size for which both times are
convection is neglected), the transfer of gas from the collapsing about equal, explicitly:
small bubbles to the macroscopic gas phase may be viewed as a
case of Ostwald ripening [26]. The important consequence of this sr ðaH0 Þ ¼ sc ðaH0 Þ ¼ sH : ð18Þ
process is that the gas contained in the bubbles is temporarily Bubbles of radius a0 < aH0 will collapse before reaching the surface
‘‘eaten’’ by the liquid. and will contribute the pressure drop for long times (t  sH ). Large
In the following, we consider that the water can accommodate bubbles a0 > aH0 will briefly and slightly contribute the pressure
the molecules from the bubbles with about no volume change (the drop since they will have hardly recessed when they reach the
approximation may be justified quantitatively). We adopt the fol- surface.
lowing notations, see Fig. 7 for illustration: P0 ; V G;0 ; nG;0 are the We may give estimates of the above characteristic time and size
initial pressure, volume of gas phase and number of CO2 moles in using Epstein–Plesset (EP) theory of bubble dissolution [8]. In the
gas phase; see Fig. 7a. We suppose that shaking instantaneously case of a bubble surrounded by a water/dissolved gas solution at
generates N bubbles, all of them of size (radius) a (see Fig. 7b). Henry’s equilibrium with a gas top phase (coefficient f ¼ 1 in EP
The bubble phase (of volume V B ) and the top gas phase (of volume original paper), the theory gives the following expression for the
V G;1 ) contain nB and nG;1 moles of gas, respectively. recession time (Eq. (41) in the original paper):
We want to estimate the decrease in pressure after the bubbles
have collapsed. Combining equations for volume conservation, 1 a20  a0 
sc ¼ 1þ

; ð19Þ
V G;1 ¼ V B þ V G;0 , and mass conservation, nG;1 ¼ nG;0  nB , we obtain 3d D a
A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53 49

where d ¼ c=qG is the ratio between the concentration of dissolved choose equal to 1 mm. Initial radii and altitude are denoted as an0
gas in bulk and the concentration in gas in the top phase (qG ). In and zn0 respectively (n being the index of the bubble).
, which has the dimension of a length, is given by:
Eq. (19), a To represent the time evolution of the bubble radius, we use
Eq. (40) of original EP article, which is written as Eq. (B3) in Appen-
2rM dix B. The latter equation only holds approximately in our case of a

a : ð20Þ
qG RT near to Henry equilibrium system, but is sufficient in the context of
our crude model, see the discussion in Appendix B and Fig. 11.
Eq. (19) indicates that the bubble lifetime considerably increases
Given the value an of the radius and the altitude zn at time ti (i
 is very small (we
with a0 . In practice, the characteristic length a
being the time index), we deduce the altitude of the bubble at time
estimate a   100 nm). So, except for extremely small bubbles, we
tiþ1 according to:
may approximate Eq. (19) as:
2qL n 2
a30 zn ðtiþ1 Þ ¼ zn ðt i Þ þ a ðt i Þ Dt: ð25Þ
sc ’ 
; ð21Þ 9g
3dDa
The bubble rise time may be estimated directly from the Stokes The previous equation results from the balance between buoyancy
drag coefficient of a sphere of initial size a0 : and viscous drag [27]. We also implicitly ignored hydrodynamic
couplings between bubbles. This assumption is probably an over-
9ghL
sr ’ ; ð22Þ simplification at times shortly after shaking, but hopefully tends
2qga20 to correct in the later stage of the process when only few bubbles
are left.
where the mass density difference between the liquid and gas
The evolution in time of the size distribution Dt ðaÞ is presented
phases has been approximated to the liquid mass density q; g is
on Fig. 8. As might be anticipated, the distribution quickly narrows;
the gravity acceleration and g is the viscosity of the liquid.
after 30 s the size amax at which Dt ðaÞ is extremum (most repre-
Combining the latter expression with Eqs. (18) and (22), we R
sented size) and the mean value hai ¼ Dt ðaÞada are almost the
obtain:
same, as shown in the insert of Fig. 8. The right side of the distribu-
 3=5 tion, corresponding to large radii, recedes under the action of buoy-
9ghL
sH ’ ð3dDaÞ2=5 ; ð23Þ ancy. The left side, corresponding to small bubbles, recedes due the
2qg
 1=5 collapse process. The value of amax quickly drops down to about
ghL
27dDa 50 lm, which later slowly decays down to 20 lm ðt ’ 200 sÞ, and
aH
0 ¼ : ð24Þ
2qg finally falls to zero within a few seconds ðt ’ 215 sÞ. Note that
the maximum of the distribution eventually goes to zero together
Using q ¼ 103 kg m3 and the numerical values listed in Table 2
with the total number N of bubbles, as shown in Fig. 9. The [50 lm,
to feed Eqs. (23) and (24), we find aH 0 ¼ 33 lm and s ¼ 42 s. As
H
20 lm] interval includes the radius aH 0 provided by the simple
reported in paragraph Section 5.1 A, microscope observations per-
model (33 lm) and is in line with the size observed in the experi-
formed near the interface (z  1 cm) through the glass of the
ment (2a = 60 ± 5 lm). The main improvement of the model, com-
champagne bottle showed that the characteristic diameter of the
pared to the ‘‘one-bubble-size’’ model, resides in our estimate of
bubbles tends towards a limit value of 60 ± 5 lm and we observed
how the bubble population evolves in time. We find that it takes
that the last visible bubble would vanish at t  220 s after shaking.
about 30 s for the distribution to focus around the ultimate bubble
The simple model presented above predicts that bubbles smal-
size, and that no bubbles are left after 215 s (extinction time).
ler than 33 lm in radius should get dissolved, and then should not
The relaxation time of the bubble population can be estimated
be observed, while bubbles larger than 33 lm should survive long
from the linear part of the log-lin plot of Fig. 9. We find
enough to reach the interface. The model then successfully repro-
duces one main feature of the observations, namely the average
size of the bubbles near the end of the bubble population lifetime. D t (a)
However the estimated time sH beyond which no bubble can 4
survive in the bulk (because they already collapsed or reached 3.0×10
-4
1×10
the free flat surface) seems too small compared to what observa-
tions indicate, namely 200 s, approximately. 2.5×10
4
80 s -5
8×10
a, amax (m)

-5
4 6×10
5.2.2. Improved model 2.0×10
A weakness of the above model stems from the fact that only a -5
4×10
constant radius is envisaged in Eq. (22). We now propose an 1.5×10
4
-5
improved version of the model where the EP collapse dynamics 2×10

is explicitly taken into account in the bubble buoyant motion. 4 0


Instead of a single bubble, we reason on a size-polydisperse distri- 1.0×10 0.0 1 2 2 2
5.0×10 1.0×10 1.5×10 2.0×10
bution. We consider an ensemble D0 of N ¼ 106 bubbles uniformly t (s)
3
10 s
distributed in the bulk (z 20; hL ½) and whose radii are also uni- 5.0×10 5s
1s 0s
formly distributed between 0 and a maximum size, which we
0.0
-4 -4 -4 -4
0 1×10 2×10 3×10 4×10
Table 2
Numerical values of the parameters used to feed Eqs. (23) and (24). a (m)
3
D ¼ 1:8  109 m2 s1 (Ref. [16]) c ¼ 7:5 kg m Fig. 8. Evolution in time of the size distribution Dt ðaÞ. The initial flat distribution
qG ¼ 8:31 kg m3 d ¼ 1:17 quickly evolves to a narrower distribution whose maximum lies between 50 and
 ¼ 0:13  106 m
a g ¼ 103 Pa s 20 lm (see text for further details). In the insert are plotted the maximum and the
hL ¼ 0:1 m mean values of the bubble size. We see that both values are almost equal once
g ¼ 9:81 m s2
buoyancy has ‘‘filtered’’ the large bubbles.
50 A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53

10
6 to 60 s and 325 s repectively, consistently with the sH / g3=5 scal-
ing provided by the simple model.
5 The conclusion is then that differences in surface tension and
10
τN ∼ 40 s viscosity of champagne, compared to values for pure water, do
change predictions of the model, but not so strongly, in so far as
4
10 the model cannot do better than providing estimates of character-
istic times for bubbles dynamics. The differences may not be signif-
N (t)

10
3
icant in view of the limitations of the model. As an important
simplification, we neglected hydrodynamic interactions between
2 the bubbles and the large scale convection due to ascending bub-
10
bles. This type of convection is evident in observations, but only
1
in the first seconds after shaking; it is hopefully negligible in the
10 later evolution of the system, when large bubbles have disap-
peared. Another possibly important point, in the case of cham-
0
10 0 pagne, is the effect of the liquid–gas interface rigidification,
50 100 150 200
which expectedly plays a role in slowing down the collapse of bub-
t (s)
bles reaching the surface [29]. This mechanism should be taken
Fig. 9. Evolution in time of the total number of bubbles. The characteristic time into account in a more elaborate version the model.
sN  40 s extracted from the linear part of the log-lin curve can be seen as the decay
time of the number of bubbles left in the bulk.

6. Conclusion
sN ’ 40 s, not far from the characteristic time s found with the H
A carbonated water solution inside a closed system, the main
‘‘one-bubble-size’’ model. In fact, neglecting the variation of the
object of our study, is the formal equivalent of a sparkling beverage
drag with the radius, as we first did, amounts to make no distinc-
inside a closed bottle. We have investigated how such a system
tion between the extinction time and the decay time sN . Finally,
evolves between a given equilibrium configuration, ruled by
the numerical model provides a satisfying value of the ‘‘extinction’’
Henry’s law, to another equilibrium. As causes of perturbation to
time (215 s), close to what is observed experimentally (220 s).
drive such changes, we investigated the effects of a small variation
of the gas content of the system and of mechanical shaking.
We showed that the system might respond to a pressure change
5.2.3. Discussion essentially through molecular diffusion, in conditions where con-
In the above analysis, we used values of r and g corresponding vection might be reduced to a minimum. We could elaborate an
to pure water. Values for champagne are definitely different, due to experimental procedure whereby these conditions were satisfied.
the presence of active molecules (amphiphilic molecules, proteins, We offered an exact resolution of the diffusion problem in this sit-
alcohol, etc.) [28]. The static surface tension of a ‘‘pure’’ hydroalco- uation, and found that the pressure response would follow a
holic solution at 12.5% of alcohol in volume is about 48 mN m1, square root law in time, within a few hours after the perturbation.
whereas it is about 46 mN m1 for champagne. Ethanol is mainly This prediction was shown to be in agreement with experimental
responsible for this drop (compared to water for which data recorded in the case of gentle decompression
r  70 mN m1 ), the other molecules being responsible for the (DP  0:1 bar). Conversely to conditions for molecular diffusion
small extra drop of 1 or 2 mN m1. In the ‘‘one-bubble-size’’ model, to be the dominant mechanism, we observed that compression of
the surface tension dependence is contained in the quantity a  (cf. the system in general leads to chemi-convection. Severe decom-
Eqs. (23) and (24)), which is proportional to r. Consequently, sH pression also drives convection due to transient cooling of the
and aH are proportional to r2=5 and r1=5 , respectively. Thus the gas phase.
dependence on r is weak, and then the bubbles dynamics should The main outcome of the experiments on shaking is that the
not differ much from that with pure water. pressure inside the bottle does not change ‘‘much’’, in so far as
Viscosity also varies according to both the chemical composi- the system initially was close to Henry equilibrium. This observa-
tion of the beverage and the temperature. Viscosity of champagne tion then rules out the common belief that shaking a bottle of
at 20° is about 1.6 mPa s (but goes up to 2.5 mPa s at 4 °C). The vis- champagne increases the internal pressure. Significant changes of
cosity has an influence on the rising time of the bubble up to the the pressure do occur if the system initially was far from equilib-
flat surface and on the shrinking time due to the local undersatu- rium, meaning that stirring the liquid just helps in quickly estab-
ration. This twofold influence is reflected in the expression for lishing the equilibrium pressure. The latter statement can be
the time sH , which is proportional to g3=5 . Note that changing the illustrated in everyday-life with a cola conditioned in a plastic bot-
viscosity from 1 to 2.5 mPa s increases sH by less than a factor 2. tle. The envelope of the bottle initially is rigid due to the high pres-
Using the numerical model, we studied the sensitivity of the sure inside. If the bottle is opened, for instance to serve a glass of
bubbles population dynamics to changes in g and r. We tested the cola, the pressure in the gas phase is temporarily lowered to
two combinations: r ¼ 40 mN m1 with g ¼ 103 Pa s, and the outside room pressure, and the envelope becomes flaccid. It
r ¼ 70 mN m1 with g ¼ 2  103 Pa s. remains so even after the bottle has been closed back. At that step,
We found that decreasing the surface tension has a small but shaking has the immediate effect of restituting the rigidity of the
visible influence. The characteristic time sN (which is the equiva- envelope.
lent of sH for the improved model) is shifted from 40 to 50 s, and Looking at the pressure response more closely, we discovered
the extinction time is shifted from 220 s to 275 s. Both changes that shaking systematically produces a small transient pressure
are consistent with the sH / r2=5 scaling provided by the simple drop, having a few tens of seconds lifetime. We observed that
model. the latter effect was concomitant to the generation of bubbles.
Increasing the viscosity from 1 to 2 mPa s has a more sensitive We could give an interpretation to the pressure based on
influence. However, the general shape of the size distribution is not Epstein–Plesset theory of bubble dissolution and collapse. The sim-
changed. The decay time sN and the extinction time are raised up ple analytical and numerical models presented in Section 5.2 were
A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53 51

satisfactory in reproducing the experimentally observed evolution where HðtÞ is the Heavyside function. The solution to Eq. (9) com-
of the bubble population, in size and lifetime. However, deriving a pleted by the conditions (8), (A1) and (A2) admit the classical
quantitative relationship between the bubble distribution and the solution
small pressure drop (amplitude and profile) turned out difficult   
z
since the initial conditions just after shaking are unknown. DcH ðz; tÞ ¼ HðtÞDP 1  erf pffiffiffiffiffiffi ; ðA3Þ
We end this Section with a few remarks about effervescence. 2 Dt
The latter phenomenon is typically observed when a bottle (of where erf stands for the error function. In our case, however, the
champagne, say) has been shaken and immediately opened, result- pressure is not constant in time. Thus, we have to generalize result
ing in foaming and gushing out of the bottle neck. Stated more for- (A3) to the case of a non perfect (still unknown) function DPðtÞ. We
mally, this situation amounts to gathering both types of the above begin by writing the pressure as a sum of impulsions, namely
mentioned perturbations, namely mechanically induced bubbling Z 1
and application of a large pressure drop DPo (the subscript ‘‘o’’ DPðtÞ ¼ DPðuÞdðt  uÞdu; ðA4Þ
stands for opening), about 6 bar in amplitude in the case of cham- 0
pagne. In this situation, bubbles lie in a highly super-saturated
since for t < 0; DPðtÞ ¼ 0. Integrating by part once leads to
liquid, and almost all of them quickly grow in size and start rising
Z 1
up, generating a kind of explosive foam. As well-known, bubble dDP
DPðtÞ ¼ ½DPðuÞHðt  uÞ1
0 þ Hðt  uÞdu; ðA5Þ
growth only concerns those bubbles which are larger than a lower 0 du
limit size, called ‘‘critical radius’’ and given by ac ¼ 2r=DP o , see e.g.
which can finally be rewritten as
[30,31].
If opening occurs just after shaking, the immense majority of Z t

bubbles are super-critical (a > ac  0:2 lm), and then efferves- DPðtÞ ¼ DP0 ðuÞHðt  uÞdu; ðA6Þ
0
cence starts fiercely. There is no pressure increase prior to opening,
and the pressure in the bottle neck keeps close to zero once the where P0 ðuÞ ¼ dDP=du. Since equations are linear, Dcðz; tÞ may sim-
bottle is opened. Gushing only lasts for a few seconds, but can be ply be derived by adding the responses to a sum of steps of infini-
prolonged by continuous shaking of the bottle. As the concentra- tesimal amplitudes P 0 ðuÞdu, which yields
tion of carbon dioxide in the liquid quickly decreases, so does the Z !
t
corresponding Henry pressure PH [32]. Conversely, the critical z
Dcðz; tÞ ¼ kH DP0 ðuÞf pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi du; ðA7Þ
radius, ac  2r=P H , divergently increases. Thus, gushing lasts as 0 2 Dðt  uÞ
long as the generated bubbles are larger than the critical size.
As a final remark, we stress again that characteristics (shapes where f ðxÞ ¼ 1  erfðxÞ. Integrating (A7) from z ¼ 0 to infinity, one
and sizes) of bubbles generated by shaking are not known a priori. gets the number of moles of dissolved gas in excess (compared to
Because predictions about the evolution of the system are sensitive the initial quantity) in the liquid phase
Z 1
to what is supposed as the initial size distribution, this problem S
deserves a dedicated study. Exploiting the Faraday instability DnL ðtÞ ¼ Dcðz; tÞdz
M 0
[33] may be a route to producing approximately controlled bubble  1=2 Z
D SkH t 0
sizes by shaking. Other methods to generate bubbles, based on ¼ DP ðuÞqðt  uÞdu; ðA8Þ
ultrasound induced cavitation [34] and shocks [35] are interesting p M 0
alternatives to study effervescence in a controlled framework. where qðxÞ ¼ x1=2 .
The conservation Eq. (11) is valid for t > 0. Its generalized form,
Acknowledgments which is valid at any time, can be written as:

This work was supported by Region Champagne-Ardenne in the dDnG dDnL


¼ þ DndðtÞ: ðA9Þ
frame of VINEAL project. We thank E. Laurichesse, the instrumen- dt dt
tation and mechanics groups of CRPP for their continuous help The differential pressure profile can be written as the sum of a
with the experiments, A. De Wit and F. Nallet for illuminating heavyside step DP HðtÞ and a continuous function DPgðtÞ whose
discussions. qualitative profiles are presented in Fig. 10. Doing this, Eq. (A8)
can be rewritten as:
Appendix A. Diffusion-based model of relaxation  1=2
D SkH
DnL ðtÞ ¼ ½qðtÞ þ g 0  qðtÞ: ðA10Þ
To make the calculation tractable, the volume of liquid is sup- p M
posed to be much larger than the volume of the vapor phase, so
Taking the Laplace transform of (A10) and using the initial con-
that we assume V L =V G ! 1. A direct consequence of the previous
dition gð0Þ ¼ 0 leads to
simplification is that the pressure in the gas phase will relax to its
 1=2
initial value – i.e. P1 ¼ P0 . On the other hand, the vanishing flux D SkH
d
D nL ðsÞ ¼ ^ðsÞ½1 þ sg^ðsÞ
sq
condition in system (8)–(11) can be replaced by a condition of van- p M
ishing differential concentration at infinity:  1=2
D SkH cPðsÞ;
Dc ! 0 for z ! 1; ðA1Þ ¼ ^ðsÞ D
sq ðA11Þ
p M
Now, consider first the basic case of a perfect pressure step – i.e. ^ ðsÞ ¼ 1=s. Finally, considering that DnL ð0Þ ¼ 0, the Laplace
since H
a situation in which the differential pressure DP ¼ P  P0 is kept at
transform of the time derivative of DnL takes the form
a constant value DP for t > 0þ (for example, by letting the intake
 1=2
tap open, in contact with a regulation manometer) D SkH 2
d
D n0L ðsÞ ¼ s q cPðsÞ;
^ðsÞ D ðA12Þ
DPðtÞ ¼ DPHðtÞ; ðA2Þ p M

Using again the decomposition of DPðtÞ into a singular and a contin-


uous part, and supposing that DnG (resp. Dn) is linked to DP (resp.
52 A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53

ΔP, ΔnG
-5
5×10

-5
4×10
t
Θ 40
-5

a (m)
3×10 30

δa/a (%)
20
-5
2×10 10
t 0
g
-5 -10
t 1×10
-20 -6 -5 -4 -3
10 10 10 10
a0 (m)
0
0 20 40 60 80 100 120 140
ΔP  , ΔnG t (s)

Fig. 11. Radius of a bubble of initial radius a0 ¼ 50 lm as a function of time. The


t relative error dsc =sc between the approximate and exact solutions to Eq. (B1)
evaluated at the time of collapse is shown in insert.

urated solution – i.e. a solution in which the dissolved gas is at


Henry’s equilibrium with the vapor phase. In this case and in
Fig. 10. Shapes of the functions involved in the resolution of Eqs. (A8) and (A9). See absence of surface tension, a bubble of any size would be stable
text for definitions. The Dirac distribution is represented by a vertical arrow. and would stay in the bulk endlessly (if buoyancy is neglected).
However the surface tension slightly increases the bubble inner
pressure (by an amount equal to Laplace pressure), making the sur-
DP) by the ideal gas law, one gets for the Laplace transform of Eq. rounding solution slightly under-saturated: the contained gas then
(A9): empties in the bulk.
RT d0 We consider a bubble of initial radius a0 , in a saturated carbon
cPðsÞ ¼ 
sD DnL ðsÞ þ DP: ðA13Þ
VG dioxide/water solution, that is to say a solution for which EP
parameter f is equal to 1. Using the same notations as in paragraph
^ðsÞ ¼ p1=2 =ð2s3=2 Þ, introducing (A12) in (A13) yields
Given that q Section 5.2.1 B1, the equation which rules the time evolution of the
radius a can be written (Eq. (34) of EP original paper) as:
cPðsÞ ¼ DPs1=2 ½a þ s1=2 ;
D ðA14Þ " #
da b=ðaqG Þ 1 1
where ¼ Dd þ ; ðB1Þ
dt 1 þ 2b=ð3aqG Þ a ðpDtÞ1=2
SkH D1=2 RT
a¼ : ðA15Þ where b ¼ 2Mr=ðRTÞ. This equation can be put in a dimensionless
2MV G 1=2
form by taking  ¼ a=a0 and x ¼ ð2DdtÞ =a0 , and one gets
Once taken the inverse Laplace transform, one finally obtains
d n= hx i
2
DPðtÞ ¼ DPea t ½1  erfðat 1=2 Þ: ðA16Þ ¼ þ 2b ; ðB2Þ
dx 1 þ 2n=ð3Þ 
Expanding the previous function in successive power of t 1=2 leads to  and b ¼ ½d=ð2pÞ1=2 . An approximate solution corre-
where n ¼ a0 =a
DPðtÞ p sponding to Eq. (B2) where the constant b has been neglected, can
¼ 1  ðt=sÞ1=2 þ ðt=sÞ þ O½ðt=sÞ3=2  ðA17Þ be written in the following polynomial form:
DP 2
At early times, the first two terms could be sufficient to properly 1  2 þ nð1  2 Þ ¼ ð3n=2Þx2 : ðB3Þ
render the pressure relaxation process, provided the third one is Once solved the previous polynomial equation and going back
negligible, that is to say ðt=sÞ1=2 > ð10p=2Þðt=sÞ – if we consider that to dimensioned variables, we obtain the radius of the bubble as
the second term must be an order of magnitude greater than the an explicit function of time. The latter is quite cumbersome and
third one. This leads to the limit beyond which the expansion the will not be given here. We stress that Eq. (B3) is valid for small
O½ðt=sÞ3=2  term has to be taken into account: b, since this parameter has been neglected in Eq. (B2). Whereas b
is small in the case of over-saturated or under-saturated solution,
4 102
t6 s: ðA18Þ it is not the case for the saturated solution at Henry’s equilibrium.
p2 Thus, we can expect some discrepancy between the approximate
The previous results have been established in the case of an infinite solution given by (B3) and the exact solution to the differential
volume of liquid. We guess that Eq. (A16) still holds for early times Eq. (B1). The relative difference in collapse time between the two
in the case of a finite volume of liquid since, considering the concen- solutions as a function of the initial radius lies between 0% and
tration profiles (at early time), the vanishing flux boundary condi- 20% in absolute value, as can be seen in Fig. 11.
tion has nearly no effect on the diffusion process.
References
Appendix B. Simplified Epstein–Plesset solution
[1] S.M. Benson, D.R. Cole, Elements 4 (5) (2008) 325–331.
[2] B. Duteurtre, Le Champagne: de la tradition à la science, Ed. Lavoisier, Paris,
We recall in this appendix some theoretical results presented at 2010.
the end of EP paper [8], concerning the collapse of a bubble in a sat- [3] G. Liger-Belair, G. Polidori, V. Zéninari, Anal. Chim. Acta 732 (2012) 1–15.
A. Vreme et al. / Journal of Colloid and Interface Science 439 (2015) 42–53 53

[4] G. Liger-Belair, Uncorked: The Science of Champagne, Princeton University [22] F. Charru, Instabilites Hydrodynamiques, CNRS Editions, Paris, 2007.
Press, New Jersey, 2013. [23] E. Guyon, J.-P. Hulin, L. Petit, Physical Hydrodynamics, Ed. Oxford, Oxford,
[5] K.M. Valant, Hétéronucléation de bulles dans des liquides sursaturés en CO2. 2001.
Ph.D. Thesis, Thèse de l’Université de Marne-la-vallée, 2005. [24] T.J. Kneafsey, K. Pruess, Transp. Porous Med. 82 (2010) 123–139.
[6] M. Vignes-Adler, Angew. Chem. Int. Ed. 52 (2013) 187–190. [25] L. Lemaigre, M.A. Budroni, L.A. Riolfo, P. Grosfils, A. De Wit, Phys. Fluids 25
[7] A. Okhotsimskii, M. Hozawa, Chem. Eng. Sci. 53 (14) (1998) 2547–2573. (2013). 014103–18.
[8] P.S. Epstein, M.S. Plesset, J. Chem. Phys. 18 (1950) 1505–1509. [26] P. Taylor, Adv. Colloid Interface Sci. 75 (1998) 107–163.
[9] F. Nadal, P. Meunier, B. Pouligny, E. Laurichesse, J. Fluid Mech. 719 (2013) 203– [27] Note that we used a viscous drag coefficient of the form 6pga instead of
229. 4pga, which can seem irrelevant in the case of a bubble. However, in real
[10] A.D. Britt, W.B. Moniz, IEEE J. Quant. Electron. 8 (12) (1972) 913–914. sparkling beverages, bubbles mostly behave as solid spheres due the presence
[11] C. Arcoumanis, J.J. McGuirk, J.M.L.M. Palma, Exp. Fluids 10 (1990) 177–180. of amphiphilic components at the liquid/gas interface (see C. Ybert, J.M. Di
[12] R. Sjoback, J. Nygren, M. Kubista, Spectrochim. Acta Part A 51 (1995) L7–L21. Meglio, Eur. Phys. J. B 4 (1998) 313–319).
[13] G.G. Agabaliantz, Bull. OIV 36 (1963) 703–714. [28] J. Senée, B. Robillard, M. Vignes-Adler, Food Hydrocolloids 13 (1999) 15–26.
[14] G. Liger-Belair, M. Bourget, C. Cilindre, H. Pron, G. Polidori, J. Food Eng. 116 (14) [29] K. Abou-Saleh, V. Aguié-Béghin, L. Foulon, M. Valade, R. Douillard, Colloids
(2013) 78–85. Surf. A 344 (2009) 86–96.
[15] A. Tamimi, Edward B. Rinkerand, O.C. Sandall, J. Chem. Eng. Data 39 (1994) [30] C.A. Ward, A. Balakrishna, F.C. Hooper, J. Fluids Eng. 92 (1970) 695–704.
330–332. [31] C.A. Ward, P. Tikuisis, R.D. Venter, J. Appl. Phys. 53 (1982) 6076–6084.
[16] M.J.W. Franck, J.A.M. Kuipers, W.P.M. van Swaaij, J. Chem. Eng. Data 41 (2) [32] We call Henry pressure the saturation pressure corresponding to the actual
(1996) 297–302. concentration in the bulk, which, after opening, does not match the external
[17] W. Lu, Huirong Guo, I.M. Chou, R.C. Burruss, L. Li, Geochim. Cosmochim. Acta partial pressure of carbon dioxide.
115 (2013) 183–204. [33] C.L. Goodridge, W. Tao Shi, D.P. Lathrop, Phys. Rev. Lett. 76 (11) (1996) 1824–
[18] S.P. Cadogan, G.C. Maitland, J.P. Martin Trusler, J. Chem. Eng. Data 59 (2014) 1827.
519–525. [34] K.K. Sahu, Y. Hazama, K.N. Ishihara, J. Colloid Interface Sci. 302 (2006) 356–
[19] J.J. Carroll, A.E. Mather, J. Sol. Chem. 21 (1992) 607–621. 362.
[20] L.W. Diamond, N.N. Akinfiev, Fluid Phase Equilib. 208 (1-2) (2003) 265–290. [35] J. Rodriguez-Rodriguez, A. Casado, D. Fuster, Why Does a Beer Bottle foam Up
[21] A. Hebach, A. Oberhof, N. Dhamen, J. Chem Eng. Data 49 (2004) 950–953. After a Sudden Impact on its Mouth?, 2013 (arXiv (arxiv.org/abs/1310.3747)).

S-ar putea să vă placă și