You are on page 1of 6

Electrochimica Acta 146 (2014) 792–797

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

On the relation between onset of bubble nucleation and gas


supersaturation concentration
M.E. Tawfik, F.J. Diez ∗
Rutgers, The State University of New Jersey, Piscataway, NJ, 08854 USA

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

Article history: The time taken for gas bubbles to start forming on an electrode operating in a stagnant aqueous solution
Received 4 June 2014 is driven by the dissolved gas mass-transfer near the electrode. This is the result of two processes that
Received in revised form 28 August 2014 compete with one another. One adds dissolved gas molecules from the chemical reactions at the elec-
Accepted 30 August 2014
trode surface and the other diffuses these molecules towards the bulk. Using this mechanism, a model
Available online 21 September 2014
is proposed that can predict the dissolved gas supersaturation concentration up to the onset of bubble
nucleation at the electrodes (heterogeneous nucleation). Experimental measurement of the bubble onset
Keywords:
nucleation time is incorporated within this model to calculate the critical supersaturation concentration
Supersaturation concentration
Onset bubble nucleation
for dissolved hydrogen and oxygen gas. The results show a strong link between the applied current den-
Diffusion sity and the supersaturation concentration at the electrode surface. A relation was obtained that predict
Hydrogen concentration the onset of bubble nucleation showing an excellent agreement with the measurement.
Electrolysis © 2014 Elsevier Ltd. All rights reserved.

1. Introduction The dissolved gas concentration is governed by the rate at which


the electroactive species are brought to the electrode surface, by the
During electrochemical reaction processes such as electrolysis process of desorption from the liquid phase to the gaseous phase
of water, electrons flow from the anode to the cathode where they and by the diffusion of the dissolved gas towards the bulk [8–10].
are consumed by hydrogen ions to form hydrogen gas. The dis- At a gas-evolving electrode the gas concentration is dependent on
solved gas builds up until a critical supersaturation concentration the mass transfer due to single-phase free convection from the
is reached at the electrode surface which marks the onset of bub- density gradients and also it is dependent on the two-phase free
ble nucleation and undergoes a phase change [1,2]. To balance the convection induced by rising bubbles [11,12]. Many studies have
electrical charge, hydroxide ions move through the electrolyte solu- considered the different mechanisms that affect the mass transfer
tion to the anode, where they give away electrons. Recent work at gas-evolving electrodes and the efficiency when bubbles adhere
has aimed at preventing or controlling gas bubble generation. For to the electrode surfaces [13–16]. Present studies model the bub-
example, in electro-osmotic pumps, the flow rate is restricted due ble production using empirical correlations with current density
to the formation of bubbles at the membrane [3]. In electrophoretic [17,18] without solving for the concentration distribution. These
deposition the bubbles generated at the electrodes may yield dam- models implicitly assume that the nucleation and growth of bub-
aged and poor quality coatings [4]. One approach at controlling bles is very fast. Other models assume that the bubbles nucleation
gas generation has been performing the electrochemical process and growth is governed by a linear rate law which depends on the
under modulated electric fields such as pulsed direct current and change of the dissolved gas supersaturation concentration [19,1].
alternating current [5–7]. A better understanding of the relation During electrolysis, the mass transfer of the electroactive species
between the dissolved gas supersaturation concentration and the (radicals) is controlled by their migration, diffusion and convec-
bubble nucleation time at the electrode surface during the electro- tion and it can be modeled by the Nernst-Planck equation [20]
chemical reaction could be used to suppress or eliminate the gas [9,10,21]. These radicals convert to dissolved gas by losing or
generation. gaining electrons at the electrodes. At the short initial stage of
the bubble nucleation in a stagnant electrolyte (the focus of the
present work), the single-phase free convection mass-transfer can
be neglected, and the dissolved gas concentration at the electrode
∗ Corresponding author. is governed by the diffusion equation (Fick’s second law) [22,20].
E-mail address: diez@rutgers.edu (F.J. Diez). This simplification required a convectionless liquid. The dissolved

http://dx.doi.org/10.1016/j.electacta.2014.08.147
0013-4686/© 2014 Elsevier Ltd. All rights reserved.
M.E. Tawfik, F.J. Diez / Electrochimica Acta 146 (2014) 792–797 793

gradient as shown in Fig. 1. The temporal evolution of the dissolved


gas concentration can be written in terms of Fick’s second law as,
2
∂Ci (x, t) ∂ Ci (x, t)
= Di , (2)
∂t ∂x2
where Di is the diffusion coefficient of species i in the working elec-
trolyte. At the electrode boundary, the dissolved gas flux, Ji (x, t), is
proportional to the concentration gradient
∂Ci (x, t)
Ji (x, t) = −Di . (3)
∂x
The analytical solution to Eq. 2 provides the evolution of the con-
centration, Ci (x, t) sketched in Fig. 1 but it can not predict the onset
of bubble nucleation t* . The goal is then to combine the solution
Fig. 1. Graphical illustration of CH+ and CH2 mass transfer near the electrode.
to Eq. 2 with experimental measurements of the bubble formation
time t* to calculate Ci∗ . To solve Eq. 2, the boundary and initial con-
ditions are needed. We can assume that, before the electric field is
gas concentration increases with time up to the critical supersatu- applied, the species concentration, Ci , is equal to the bulk satura-
ration concentration when the onset of bubble nucleation occurs. tion concentration of the dissolved gas, Ci∞ . Considering that the
Since the bubbles do not nucleate infinitely fast, some time elapses electrolysis cell is large compared to the diffusion thickness, the
until the appearance of the gas phase [23]. This differs substantially concentration far from the electrode reaches the bulk concentra-
from what happens in continuous electrolysis where nucleation tion, Ci∞ . This can be written as,
time is very small or possibly inexistent. While bubble production
and growth has received much attention, there is a lack of infor- Ci (x, 0) = Ci∞
(4)
mation on the prediction of onset of bubble formation. Only the lim Ci (x, t) = Ci∞ .
work by Dinkelacker [24] reported the bubble formation time as a x→∞

function of current density. The present work relates the onset gas When a constant current, I, is applied across the electrolysis cell,
formation time with the critical supersaturation concentration near the resultant flux across the electrode surface, Js , is constant and
the electrode in a stagnant electrolyte. A supersaturation model is given by
proposed that uses Fick’s law and results from experimental mea-  
surements of the bubble formation time. I ∂Ci (x, t)
−Js = −Ji (0, t) = = Di
zFA ∂x x=0
2. Theoretical model We can now solve Eq. 2 to obtain the concentrations of oxidized
and reduced species at the electrode (Sand [22] and Karaoglanoff
2.1. Chemistry of water electrolysis [26]) giving
 1/2    
Js Di t x2 x
The present study models the increase in the dissolved gas con- Ci (x, t) = Ci∞ + 2 exp − − x erfc (5)
Di  4Di t 2(Di t)
1/2
centration Ci , of species i from basic principles to solve for the
critical supersaturation concentration Ci∗ . The number of dissolved This equation can be written in terms of the non-dimensional
gas moles generated at the electrode, n, is related to the current parameter  as
drawn by Faraday’s law,  
Js x 1
It Ci (x, t) = Ci∞ + √ exp(−2 ) − erfc() (6)
n= (1) Di 
Fz
where x is nondimensionalized
√ by a diffusion
√ thickness ı, with a
where I is the constant current drawn by the electrolyte, z is the
general scaling of ı∼ Dt, so that  = x/ 4Dt. Further details about
valency number of ions of the substance (electrons transferred per
the choice of ı will be given next. We assume that before the onset
ion), and F is the Faraday’s constant.
of bubble formation, this diffusion equation (Eq. 6) dominates the
In our experiments we have chosen to work with simple, low
mass transfer. Once bubbles are formed, other mass transfer mech-
ionic concentration aqueous buffers and inert platinum electrodes
anisms such as single-phase and two-phase convection and gas
so that the electrodes half-reactions (acid) is,
adsorption can dominate. These have been widely considered in
Cathode (Reduction) : 2H+ (l) + 2e− −→ H2 (l) the study of the gas-evolving electrodes [8–12,27,28], but will not
be considered here since our analysis of the problem stops at the
Anode (Oxidation) : 2H2 O(l) −→ O2 (l) + 4H+ (l) + 4e− onset. This diffusion equation is plotted in Fig. 2, showing how the
concentration of the dissolved gas, Ci (x, t), varies as we move away
The dissolved H2 and O2 gas, formed at the electrodes, build from the electrode as well its variation at the electrode Ci (0, t). As
∗ and C ∗ respec-
up until it reaches the critical supersaturation CH
2 O2 time increases, the concentration at the electrode surface increases
tively. Beyond this threshold and in the presence of a nucleation and also the the boundary layer thickness from the surface where
site, crack or cavity in the electrode surface, bubbles are initiated the dissolved gas concentration CH2 is larger than the bulk con-
[25]. centration. This region of influence corresponds  to a diffuse layer
thickness ı which can be shown to scale as ∼ Di t.
2.2. Concentration gradient of dissolved gas The results in Fig. 2, show that the concentration of dissolved
gas at the electrode increases with time. In such an environment
The dissolved gas generated at the electrodes diffuses away rich in excess dissolved gas molecules, the conditions are being set
towards the bulk. This produces a higher concentration near the for bubble nucleation. The process is not well understood with the
electrode than at the bulk solution resulting in a concentration traditional view being that the onset of nucleation is the result of
794 M.E. Tawfik, F.J. Diez / Electrochimica Acta 146 (2014) 792–797

Fig. 2. Concentration profile of H2 (l) before onset of bubble nucleation showing a) Concentration of the dissolved gas as a function of distance from the electrode, and b)
Concentration evolution with time at the surface of the electrode.

this environment reaching some energy state where instantaneous


nucleation occur. Nevertheless, some macroscopic-scale proper-
ties can be extracted from this process. First, higher fluxes at the
electrode surface, Js , will increase the dissolved gas concentration
due to increase in the reaction rate. This will increase interactions,
which suggest that the time it takes to reach t* will be faster with an
increase in the applied current. Experimentally we will show this
dependence with I/A. Second, for the proposed diffusion-controlled
model, a diffuse layer, that is a function of time, grows from the
electrode surface. Physically, the diffuse layer represents the region
where the majority of the supersaturated dissolve gas being gener-
ated at the electrode is being contained and which increases with
time. It is expected that the onset of nucleation will be dependent
on this layer properties as will be shown from the experiments.
There are multiple way
√ of defining the diffusion thickness ı, with
a general scaling of ∼ Dt. Here, it is defined as the distance from
the electrode surface where the mass-flux ratio reaches J/Js ∼ 0.01.
Fig. 3. Average Concentration (
CH2 − CH∞ ) evolution with time for different current
The mass-flux J(x, t) can be calculated by differentiating Eq. 5 and 2
densities. Circles represent onset of bubble nucleation measurements discussed in
substituting it into Eq. 3. The solution can then be written in terms section 4.
of the non- dimensional parameter  as

J(x, t) = Js erfc() (7)


√ layer. To find this relation, first we will evaluate the nucleation
where  = x/ 4Dt. This shows that the mass-transfer and the time experimentally.
mass-flux ratio J/Js depends only on the non- dimensional param-
eter . For the ratio of J/Js ∼ 0.01, it can be shown from Eq. 7 that
 ∼ 1.82. This is the distance from the electrode where the mass-flux 3. Experimental setup
is nearly zero and can be written as,
 The electrochemical measurements are conducted at room tem-
ı  3.64 Di t. (8)
perature (20 ◦ C) in a 10 ml electrochemical cell. The primary
The diffuse layer contains most of the dissolved hydrogen CH2 solutions used are deionized ultra-filtered water and 1 mM borate
that is being generated at the electrode. This can be estimated by buffer made with sodium tetraborate(≥99% pure) and boric acid
calculating the average concentration CH2 within the diffuse layer (≥99% pure) from Sigma Aldrich. Platinum wires (99.9 % pure)
as, 0.5 mm in diameter from Sigma Aldrich are used as the cell elec-

ı
trodes. A Trek model 5/80 amplifier with a Tektronix AFG 3021B
1 function generator is used to supply power to the electrochemical
Ci (t) = Ci (x, t) dx
ı 0
cell. A microscope is used to measure the bubble formation time
(9)
0.274 Js √ by taking a video of the working electrode while applying a con-
  t + Ci∞ . stant current across the electrochemical cell. The bubble nucleation
DH2
time depends on the critical supersaturation CH∗ value of hydrogen
2
As the diffuse layer grows with time, the average concentration dissolved gas. Experimentally, this is evaluated by measuring the

CH2 within the diffuse layer also grows as shown in Fig. 3. These are time it takes for the first bubbles to form from the time the volt-
the results for four different constant current densities, I/A, applied. age is applied. The time is measured until the first bubble reaches
The measurements show that for high currents the concentration 50 m in diameter. This type of technique has been applied in the
increases faster with time as expected. Although this plot resolves past and there can be small variations in the results depending on
the behavior of the concentration in the diffuse layer, additional the method used to detect the formation of the first bubbles [23].
information is needed to determine the onset of bubble nuclea- There could be a systematic bias on the time measured by different
tion. The hypothesis is that the onset time, t* , should be related methods but it would not affect the relation between the nuclea-
to the concentration of the gas molecules in the growing diffuse tion time and supersaturation concentration. In that relation the
M.E. Tawfik, F.J. Diez / Electrochimica Acta 146 (2014) 792–797 795

Fig. 4. Onset of bubbles nucleation as a function of current density for a) H2 bubbles, and b) O2 bubbles.

Fig. 5. Average critical supersaturation concentration as a function of a) onset time and b) current density, for both H2 and O2 .

slope would remain the same but the constant could show small
variations.

4. Results

The time t* at which the first hydrogen bubble is observed


forming on the surface of the cathode is measured for different
current densities and compared with published data [24] in Fig. 4a.
Similar measurements are performed at the anode, to resolve the
time needed for the onset of oxygen bubbles for different applied
currents as shown in Fig. 4b. The small variations between the
present work and that by Dinkelacker [24] in Fig. 4a is attributed to
different measurement methods used and different definitions of
the critical supersaturation concentration (Ci∗ ) discussed in the lit-
erature. In Dinkelacker [24] measurements are for Nickel electrodes
in 5 molar KOH solution run at 20 ◦ C, while the present experi- Fig. 6. Diffusion layer thickness as a function of current density at t* .

ment uses Platinum electrodes in 1 mM Borate buffer. Shibata [29]


defines Ci∗ at the electrode surface, while Glas [30] defines it at
between the calculated ∗
CH and CH2 (0, t ∗ ) and applied current
the bubble surface. A comparison between these two definitions 2
has been reported by Vogt [31] which discusses the importance of density. The thickness of the diffusion layer, ı, can also be obtained
each definition. A comparison between the results in Fig. 4a and 4b at the nucleation time from Eq. 8 for the five different current den-
shows that for an applied current the onset of H2 bubbles is two to sities tested and it is plotted in Fig. 6. The results show that at the
three times faster than for O2 bubbles. onset time t* the diffusion thickness decreases as the current den-
Once the nucleation time is experimentally measured, the sity is increased. The results are compared
√ to Nefedov [27] which
concentration at the electrode surface Ci (0, t* ) and the average con- proposed a diffusion thickness, ı = Dt based on an experimen-
centration within the diffuse layer,
Ci∗ can be obtained from Eqs. 5 tal analysis of the bubble growth. Although the work in [27] deals
with bubble growth rate and how they affect mass transfer, we can
and 9 respectively. ∗ and
CH CO∗ are plotted in Fig. 5a showing a log- compare their relation at the onset of bubble formation with Eq. 8
2 2
arithmic decay for increasing t* . This suggests that the nucleation as shown in Fig. 6. The trend is the same for both curves, since both
time does not occur at constant average concentration within the equations only differ by a constant due to the manner in which they
diffuse layer but rather
Ci∗ decreases for increasing t* when the cur- are obtained. One (Eq. 8) is obtained from the diffusion equation,
rent applied decreases. Similarly, Fig. 5b captures the dependence while the other (Eq. 9) relates to bubble growth. In summary, this
796 M.E. Tawfik, F.J. Diez / Electrochimica Acta 146 (2014) 792–797

for an aqueous buffer. Both the onset of H2 and O2 bubble forma-


tion was measured for a range of applied currents. At the calculated
time, t* , the dissolved gas concentration at the electrode surface and
the average concentration within the diffuse layer was predicted
using the diffusion model. The results show that the nucleation
time does not occur at constant Ci∗ , but rather it depends on the cur-
rent applied, with Ci∗ decreasing as I/A is decreased. Similarly, the
results also show that the diffusion thickness decreases at the onset
time t* , as the current density is increased. In summary, the nuclea-
tion time t* is a function of the applied current density, and there
is a strong link between the applied current density and the con-
centration in the diffuse layer. An empirical relation was obtained
that demonstrated this dependency showing an excellent agree-
ment with the measurements. These were used to obtain a relation
between t* and the applied current density that predicts the onset
of bubble nucleation in inert electrodes with only the additional
knowledge of the diffusivity of the dissolved gas. This, in turn, can be
Fig. 7. The onset time dependency on the concentration at the electrode surface used with the diffusion model to obtain the critical supersaturation
and the current density.
concentration.

thin diffuse layer surrounding the electrode describes where the


excess CH2 is found before bubble nucleation, and according to [27] Acknowledgment
it plays a role during bubble growth.
The results have shown that the nucleation time t* is a function The authors gratefully acknowledge the financial support pro-
of the applied current flux. They have further shown that it is also a vided for this study by the Office of Naval Research (ONR), Grant No.
function of the concentrations at the electrode surface C(0, t* ) and N00014-11-1-0019, with Dr. Thomas F. Swean of the Ocean Engi-
the average concentration within the diffuse layer, C ∗ . It should be neering and Marine Systems Program serving as Program Manager.
* ∗
noted that both C(0, t ) and C can be obtained from the diffusion
equation and are simply related by a constant. The measurements References
were performed for two different reactions (i.e.: H2 and O2 gas gen-
eration) with different diffusivities, D, (DH2 = 4.5 × 10−9 m2 /s−1 , [1] S. Van Damme, P. Maciel, H. Van Parys, J. Deconinck, A. Hubin, H. Deconinck,
Bubble nucleation algorithm for the simulation of gas evolving electrodes, Elec-
DO2 = 2.4 × 10−9 m2 /s−1 [32]) with the results showing similar trochemistry Communications 12 (5) (2010) 664–667.
scaling. This further supports that the nucleation time is diffusion- [2] S. Jones, G. Evans, K. Galvin, Bubble nucleation from gas cavities - a review,
controlled. An empirical relation can be obtained that includes the Advances in Colloid and Interface Science 80 (1) (1999) 27–50.
[3] D. Piwowar, M. Tawfik, F. Diez, High field asymmetric waveform for ultra-
dependency of t* on C(0, t* ) and I/A which can be written as enhanced electroosmotic pumping of porous anodic alumina membranes,
Microfluidics and Nanofluidics 15 (6) (2013) 1–12.
C(0, t ∗ )a (I/A) = constant.
b
(10) [4] M. Ammam, Electrophoretic deposition under modulated electric fields: a
review, RSC Advances 2 (20) (2012) 7633–7646.
By fitting the experimental results for both H2 and O2 with Eq. [5] P. Selvaganapathy, Y. shun Leung Ki, P. Renaud, C. Mastrangelo, Bubble-free
10, a value of a = 1 and b = −0.5 are obtained which support the electrokinetic pumping, Journal of Microelectromechanical Systems 11 (5)
dependencies observed from the measurements. This relation, mul- (2002) 448–453.
[6] A. Nold, R. Clasen, Bubble-free electrophoretic shaping from aqueous suspen-
tiplied by t* , is plotted in Fig. 7, and the data fitted by 0.85 t* . By sion with micro point-electrode, Journal of the European Ceramic Society 30
substituting into Eq. 10 the concentration at the electrode surface (14) (2010) 2971–2975.
obtained from the diffusion Eq. 5 as [7] B. Neirinck, J. Fransaer, O.V.d. Biest, J. Vleugels, Aqueous electrophoretic
deposition in asymmetric AC electric fields (AC-EPD), Electrochemistry Com-
2(I/A) √ munications 11 (1) (2009) 57–60.
C(0, t) − C ∞ = √ t, (11) [8] E.O. Vilar, E.B. Cavalcanti, I.L. Albuquerque, A Mass Transfer Study with elec-
zF D trolytic gas production, Advanced Topics in Mass Transfer 9 (2011), 978-953.
we can solve for t* . This provides a relation between t* and the [9] H. Vogt, The role of single-phase free convection in mass transfer at gas evolving
electrodes-I. Theoretical, Electrochimica acta 38 (10) (1993) 1421–1426.
applied current density, I/A, as [10] H. Vogt, The role of single-phase free convection in mass transfer at gas evolv-
 I −1 ing electrodes-II. Experimental verification, Electrochimica acta 38 (10) (1993)
t ∗ = 0.19z 2 F 2 D . (12) 1427–1431.
A [11] L. Janssen, E. Barendrecht, The effect of electrolytic gas evolution on mass trans-
fer at electrodes, Electrochimica Acta 24 (6) (1979) 693–699.
This relation can be used to predict the onset of bubble nuclea- [12] G.M. Whitney, C.W. Tobias, Mass-transfer effects of bubble streams rising near
tion. It is plotted in Fig. 4 showing an excellent agreement with the vertical electrodes, AIChE Journal 34 (12) (1988) 1981–1995.
t* measurements. [13] L. Janssen, Mass transfer at gas evolving electrodes, Electrochimica Acta 23 (2)
(1978) 81–86.
[14] H. Vogt, The rate of gas evolution of electrodes I. An estimate of the efficiency of
5. Conclusion gas evolution from the supersaturation of electrolyte adjacent to a gas-evolving
electrode, Electrochimica Acta 29 (2) (1984) 167–173.
[15] H. Vogt, On the gas-evolution efficiency of electrodes I - Theoretical, Elec-
The mass-transfer mechanism that governs the dissolved gas trochimica Acta 56 (3) (2011) 1409–1416.
concentration at an electrode before the onset of bubble nucleation [16] H. Vogt, On the gas-evolution efficiency of electrodes. II-Numerical analysis,
is discussed. This is a diffusion-control mechanism where the dis- Electrochimica Acta 56 (5) (2011) 2404–2410.
[17] H. Vogt, R. Balzer, The bubble coverage of gas-evolving electrodes in stagnant
solved gas concentration distribution due to electrolysis is obtained electrolytes, Electrochimica Acta 50 (10) (2005) 2073–2079.
by solving Fick’s second law. The closed form solution is evaluated [18] R. Wuthrich, C. Comninellis, H. Bleuler, Bubble evolution on vertical elec-
to obtain the concentration at the electrode surface as well as the trodes under extreme current densities, Electrochimica Acta 50 (25-26) (2005)
5242–5246.
average concentration within the diffuse layer. [19] P. Maciel, T. Nierhaus, S.V. Damme, H.V. Parys, J. Deconinck, A. Hubin, New
The onset of bubble nucleation, t* , at which bubbles are first model for gas evolving electrodes based on supersaturation, Electrochemistry
observed forming on the surface of inert electrodes was measured Communications 11 (4) (2009) 875–877.
M.E. Tawfik, F.J. Diez / Electrochimica Acta 146 (2014) 792–797 797

[20] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamentals and Applica- [26] V.Z. Karaoglanoff, Über Oxydations-und Reduktionsvorgänge bei der Elek-
tions, 2 edn., John Wiley & Sons, 2001. trolyse von Eisensalzlösungen, Zeitschrift für Elektrochemie und angewandte
[21] H. Vogt, Mechanisms of mass transfer of dissolved gas from a gas-evolving physikalische Chemie 12 (1) (1906) 5–16.
electrode and their effect on mass transfer coefficient and concentration over- [27] V.G. Nefedov, Mass transfer to a gas-generating electrode, Russian journal of
potential, Journal of Applied Electrochemistry 19 (1989) 713–719. electrochemistry 34 (1) (1998) 16–24.
[22] H.J. Sand III, On the concentration at the electrodes in a solution, with special [28] V. Nefedov, O. Artyushenko, E. Kashevarova, Mass transfer to horizontal
reference to the liberation of hydrogen by electrolysis of a mixture of cop- gas-generating electrodes, Russian Journal of Electrochemistry 42 (6) (2006)
per sulphate and sulphuric acid, Philosophical Magazine Series 6 1 (1) (1901) 638–642.
45–79. [29] S. Shibata, The concentration of molecular hydrogen on the platinum cathode,
[23] D. Kashchiev, A. Firoozabadi, Kinetics of the initial stage of isothermal gas phase Bulletin of the Chemical Society of Japan 36 (1) (1963) 53–57.
formation, The Journal of chemical physics 98 (1993) 4690. [30] J. Glas, J. Westwater, Measurements of the growth of electrolytic bubbles, Inter-
[24] M. Dinkelacker, Zur Gasentwicklung und Blasenbildung an Elektronen, Thesis national Journal of Heat and Mass Transfer 7 (12) (1964) 1427–1443.
Universitat Stuttgart (1989). [31] H. Vogt, On the supersaturation of gas in the concentration boundary layer of
[25] B.C. Donose, F. Harnisch, E. Taran, Electrochemically produced hydrogen bub- gas evolving electrodes, Electrochimica Acta 25 (5) (1980) 527–531.
ble probes for gas evolution kinetics and force spectroscopy, Electrochemistry [32] D.R. Lide, CRC Handbook Chemistry and Physics, 85th Edition, CRC Press, 85
Communications 24 (2012) 21–24. edition edn., 2004.

You might also like