You are on page 1of 9

Journal of The Electrochemical

Society

You may also like


- Visualization of Mossy Zinc Electroplating
Effects of Deposition Conditions on the Structure Evolution Via Operando
Nanotomography
Morphology of Zinc Deposits from Alkaline Zincate Fan Wang, Mingyuan Ge, Andrew Hitt et
al.

Solutions - Rota-Hull Cell Study on Pulse Charging of


Zinc/Air Redox Flow Batteries
Christian Zelger, Jennifer Laumen and
To cite this article: R. Y. Wang et al 2006 J. Electrochem. Soc. 153 C357 Bernhard Gollas

- The Dynamic Interfacial Understanding of


Zinc Electrodeposition in Ammoniacal
Media through Synchrotron Radiation
Techniques
View the article online for updates and enhancements. Yuexian Song, Jiugang Hu, Jia Tang et al.

This content was downloaded from IP address 24.130.24.166 on 21/03/2023 at 16:21


Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲 C357
0013-4651/2006/153共5兲/C357/8/$20.00 © The Electrochemical Society

Effects of Deposition Conditions on the Morphology of Zinc


Deposits from Alkaline Zincate Solutions
R. Y. Wang,a,*,z D. W. Kirk,a,** and G. X. Zhangb
a
Department of Chemical Engineering and Applied Chemistry, University of Toronto, Toronto, Ontario,
Canada, M5S 3E5
b
Teck Cominco Metals, Limited, Mississauga, Ontario, Canada, L5K 1B4

The morphology of zinc deposits from alkaline zincate solutions was systematically investigated over a wide range of conditions.
Zinc deposit morphology was categorized as heavy spongy, dendritic, boulder, layerlike, and filamentous mossy. The formation of
dendritic, layerlike, and mossy deposits is under single-factor control, and these morphologies are defined as type I. The formation
of heavy spongy and boulder deposits is under multiple-factor control, and they are defined as type II. Within type I, dendrite
initiation and growth are both under diffusion control, while layer and mossy formation is under activation control. Based on the
effects of deposition conditions, morphology diagrams were constructed. These diagrams demonstrate that the control factors are
not independent, and no universal critical overpotential for a specific deposit morphology exists.
© 2006 The Electrochemical Society. 关DOI: 10.1149/1.2186037兴 All rights reserved.

Manuscript submitted September 30, 2005; revised manuscript received January 17, 2006.
Available electronically March 31, 2006.

Electrodeposition of zinc is important in electrowinning, electro- heavy spongy deposits20 due to their similar appearance, i.e., black
galvanizing, and zinc rechargeable batteries.1 The morphology of powder. Jaksic21,22 studied the inhibition of mossy deposits and
zinc deposits and its control play a critical role in the quality and found that copper catalyzed hydrogen reduction and triggered
performance of the final products. Early morphology studies of zinc spongy deposits; molybdenum, lead, and mercury suppressed or
deposition were driven by the application of zinc electrowinning, eliminated mossy formation. Jaksic also claimed that mossy forma-
where deposits were obtained from acidic chloride or sulfate solu- tion was an autocatalytic, mass-transfer-controlled process. Cachet23
tions. Later, interest in developing rechargeable zinc batteries as a attributed mossy formation to anodically dissolved zinc.
high-energy power source has stimulated sustained research on McBreen and Gannon24 examined 2–4% w/v 共weight/volume兲
deposition from alkaline zincate solutions. The cycle life of re- addition of HgO, Ti2O3, PbO, CdO, In2O3, and Ga2O3
chargeable zinc batteries has been found to be limited by the mor- to pasted ZnO electrodes. These metals 共except for gallium兲 were
phological changes of the zinc electrode related to shape change, believed to deposit before zinc and act as a substrate. Those with
dendritic formation, and loss of porosity. interatomic distances close to zinc yielded fine deposits. Other
Naybour2 reported three types of deposits, dendritic, layer-like additives that have been found to influence zinc morphology
mixed with boulder, and mossy. Diggle and co-workers3 found that include tin,25,26 antimony,18 lead,17,25-27 tetraethyl ammonium,27
below an overpotential of 85 mV, a spongy deposit formed, and perfluorosurfactants,28 polyoxylethylene,29 citrate,30 and fluorinated
above 140 mV, a mixed spongy and dendritic deposit formed. A surfactants.23
critical overpotential for dendritic initiation was determined to be In spite of the numerous investigations on the morphology of
75–85 mV. Diggle concluded that dendritic initiation was under dif- zinc deposits from alkaline solutions, there is still a lack of system-
fusion control, while the growth stage was under activation control. atic characterization and understanding. It is the objective of this
Mansfeld and Gilman observed that at 200 mV, dendrites developed study to investigate the complete range of morphological variations
from the electrode surface directly,4 rather than from the tips of with respect to the common deposition conditions. The results will
protrusions or pyramids.3 Moshtev and Zlatilova5 found the average be presented in three papers. The aim of this paper is to characterize
radius of curvature at the protrusion tips increased proportionally and categorize the morphological features of zinc deposits. Attempt
with time. The effects of variables, including overpotential, electro- is made to link the controlling mechanism to the deposit morphol-
lyte composition, and substrate orientation on the dendritic precur- ogy change. Understanding of the mechanisms involved in the for-
sors were investigated. Moshtev stated that the overpotential was the mation of various morphologies will be dealt in a following paper.
critical variable rather than the current density 共cd兲 for deposit mor- Experimental
phology. Efforts have been made toward preventing dendritic forma-
tion by applying various charging methods,6-8 lowering zincate The electrolyte was 7 mol L−1 共M兲 KOH solution saturated with
concentration,9-14 and applying convection.15,16 ZnO 共0.7 M兲. It was prepared by dissolving analytical reagent 共AR兲-
Other deposit structures have also been studied to a more limited grade KOH pellets and ZnO powder in triple-distilled water. The
extent. Bockris and co-workers17 attributed layerlike growth of zinc solution was stirred for 4 h accompanied by nitrogen bubbling to
at low overpotential to the bunching of monatomic steps into mac- remove oxygen. A gelled electrolyte was prepared by mixing
rosteps due to the higher surface energy of the edge plane than the 7% w/v potassium polyacrylate with 93% w/v 0.7 M zincate solu-
surface plane. At higher overpotential, nucleation occurred more tion. The mixture was stirred at room temperature for 70 h, followed
generally, giving rise to a uniform growth of boulder deposits. Based by a few minutes vacuum to remove air bubbles trapped in the gel.
on cyclic voltammogram 共CV兲 experiments, Lamping and Polycrystalline zinc sheet was obtained from Alfar Aesar 共99.9%兲.
O’Keefe18 proposed that zinc followed a lateral growth mode at low Zinc single crystals 共slowly cast, bulk zinc containing large grains,
overpotential, but at high overpotential a vertical growth mode courtesy of Teck Cominco兲 were obtained by cleavage of the zinc
dominated. sample at liquid nitrogen temperatures. In order to obtain the desired
The formation of mossy deposits has been reported.2,6,17,19 Nev- surface area, foil electrodes were coated with epoxy 90HS 共Amer-
ertheless, before the development of the scanning electron micro- coat Canada兲, and single crystals were mounted in epoxy resin
scope 共SEM兲, there was difficulty in distinguishing mossy from 共LECO兲.
In the study of the effects of deposition conditions, a rectangular
cell was modified to obtain a 267-mL Hull cell31 共Fig. 1兲. The rect-
* Electrochemical Society Student Member. angular cell was divided by a clear separator to keep the trapezoidal
** Electrochemical Society Active Member. shape of the Hull cell. For in situ microscopy, the electrolyte was
z
E-mail: rainey@chem-eng.utoronto.ca filled in both sides of the separator to avoid reflection from the
C358 Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲

of deposition. Filamentous mossy deposits have the appearance of


tangled whiskers with a typical diameter of 50–200 nm and lengths
that can exceed 5 ␮m.
Effect of surface preparation.— The initial deposition is greatly
affected by electrode surface conditions, such as surface roughness,
contamination, and oxidation condition. These conditions are also
critical to experimental reproducibility in electrochemistry. Deliber-
ately changing surface conditions can provide valuable information
for revealing deposition mechanisms. Prior to deposition, zinc elec-
Figure 1. The modified Hull cell: 共1兲 separator, 共2兲 auxiliary cathode, 共3兲 trodes are usually degreased, mechanically ground, and acid-etched.
cathode, 共4兲 anode, and 共5兲 microscope with CCD camera. The effect of electrode preparation was demonstrated by studying
the effect of mechanical grinding and acid-etch.
Effect of grit size of Si carbide paper.— After grinding with
conventional angled Hull cell. Another modification was that an 1200-grit Si carbide paper, half of the electrode surface was ground
auxiliary cathode was placed in front of the cathode and was pulled with 180 grit. As a result, coarse and smooth surfaces were gener-
up to expose one-quarter more of the cathode at each time interval. ated on the same electrode. As illustrated in Fig. 3a, at low cd,
Thus, deposits on the cathode were divided into four time regions. mossy deposits tended to grow along grooves caused by grinding. At
The Hull cell was calibrated at each experimental condition. For high cd 共Fig. 3b兲, dendrites developed on both smooth and coarse
calibration, a cathode was sectioned into ten insulated pieces, and surfaces. The dendrites were relatively uniform in terms of number
the current passing through each piece was determined. The current and size over the entire electrode, indicating much less sensitivity to
density 共cd兲 vs location plot was then used to interpret site-specific surface roughness.
cd. After deposition, boundaries dividing adjacent deposit mor-
phologies on a Hull cell cathode were identified using an optical Effect of acid-etch.— After grinding with 1200-grit Si carbide
microscope 共Olympus SZ11兲, and the apparent cd at boundaries was paper, half of the electrode was etched with 25% HNO3 for 10 s. At
determined from the cd calibration plot. low cd, electrodeposition produced large numbers of mossy deposits
The reference electrode was Hg/HgO. A Solartron 1280B unit on the unetched surface compared to few mossy deposits on the
was used for current or potential control. In potential control, current etched surface 共Fig. 4a兲. At higher cd, the morphology was boulder
interrupt was used for IR compensation. Hitachi S4500 and S5200 on both surfaces. Fewer but larger boulder deposits were found on
SEM were used for scanning electron microscopy. A Philips the unetched electrode surface than on the etched electrode surface
PW3710 diffractmeter was used for crystallographic determination. 共Fig. 4b兲.
A Coulter SA3100 was used to determine deposit Brunauer-Emmett- The effect of etching suggests that during mossy initiation, mi-
Teller 共BET兲 surface area, using N2 as absorbent. crozinc particles generated in mechanical grinding provide consid-
erable numbers of heterogeneous nucleation sites. The acid-etch dis-
Results solves these sites and reduces the opportunity for mossy deposits to
nucleate. At high cd, nucleation of boulder and dendritic deposits is
Categorization of deposit morphology.— Zinc deposits were re- possible regardless of availability of nucleation centers. This result
ported as dendritic, layerlike, and mossy, developing at high, mod- is consistent with the grit size results. The abrasion sites in deep
erate, and low cd, respectively.2 In this study, based on deposit crys- grooves produced in mechanical grinding cannot totally be removed
tallinity and microstructure, zinc deposits were divided into five in washing or acid etching, and thus act as nucleation centers in
categories: heavy spongy, dendritic, boulder, layerlike, and filamen- mossy initiation. The above effects suggest mossy initiation is de-
tous mossy, as summarized in Table I. termined by the nucleation process and thus is highly site-selective,
As illustrated in Fig. 2, heavy spongy deposits are large boulder while dendritic and boulder formation is not selective.
agglomerates and highly branched dendrites, developing at very
high cd and after long deposition time. Dendrites are treelike and Effect of substrates.— On substrates other than zinc, the crystal-
can be hexagonal if deposition is slow. Boulders are typically hex- lographic misfit between the substrate and zinc deposit affects de-
agonal shape in discrete assemblies. Layer-by-layer growth on indi- posit morphology. As demonstrated in Fig. 5, at low cd, layerlike
vidual boulder deposits can be seen from the image. Layerlike de- deposits were not obtained on copper, nickel, or iron substrates.
posits are epitaxial growth and are typically found at the beginning Instead, boulder deposits followed by mossy deposits were found.

Table I. Categorization of Zn deposits.

Heavy spongy Dendritic Boulder Layerlike Mossy


Appearance 共unaided Black powder Metallic Gray metal Shiny metal Black powder
eye兲 crystal
Microstructure Agglomerate Fern, leaf, Granular Ridge, layerlike Filament,
共under microscope兲 large boulder hexagon boulder whisker
Adherence Nonadherent Nonadherent Adherent Adherent Nonadherent
Porosity Dispersed Dispersed Compact Compact Highly porous
nonporous nonporous
Crystallinitya Anisotropically Isotropically Anisotropically Epitaxial Isotropically
oriented oriented oriented oriented oriented
Growth cd Highest Very high Moderate Low Lowest
Nucleation site Nonselective Nonselective Selective Nonselective Highly
selectivity selective
Category Type II Type I Type II Type I Type I
a
Anisotropically oriented: deposits grow three-dimensionally; bulk and individual deposits are polycrystalline. Isotropically oriented: deposits grow
two-dimensionally; individual deposits are single crystals. 共To be discussed in a following paper.兲
Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲 C359

Figure 2. SEM images of five distinct Zn


deposit morphologies

Copper and nickel crystals are face-centered cubic 共fcc兲 struc- was more selective. Only on grooves produced in mechanical grind-
ture, for which the close-packed plane is 共111兲 with an interatomic ing did zinc crystals preferentially initiate, and crystals were few in
distance of 0.2556 nm for copper and 0.2492 nm for nickel.32 At number but large in size.
low cd, zinc deposits tend to grow at 具0001典. The
共0001兲Zn /共111兲Cu,or Ni fitting should be the best crystal face match. Effect of impurities.— Lead, tin, and a number of other additives
have been found to be suppressors of dendrites.21 Their effects on
Based on an interatomic distance of 0.2665 nm for the zinc basal
other morphologies, such as mossy, have not been examined in de-
plane 共0001兲, copper has a closer interatomic distance, and Lamping
tail. Some common impurities present in commercial zinc have been
reported fine deposits were obtained.33 Because of larger misfit of
investigated. Images of the deposits from deliberate addition of
共0001兲Zn /共111兲Ni, zinc forms islands 共boulders兲 on nickel in accor-
these impurities to the zincate solutions are illustrated in Fig. 6.
dance with the Volmer–Weber model with negative misfit.34 The addition of indium, arsenic, lead, and bismuth caused re-
Iron has a body-centered cubic 共bcc兲 structure and the 共110兲 duced cd at constant potential 共Fig. 7兲, and reduced the roughness of
plane is the close-packed plane with an interatomic distance of deposits correspondingly 共Fig. 6兲.
0.2482 nm, which is smaller than that of the zinc basal plane. For Addition of arsenic resulted in a similar cd–time curve as that of
iron, 共0001兲Zn /共110兲Fe fitting at 具112̄0典Zn /具111典Fe direction was pro- no additive. However, after a steady-state growth period of layerlike
posed to have the least misfit.24,35,36 Similar to the growth on nickel, or boulder deposits, there was a rapid increment of cd, suggesting
zinc formed islands on top of iron in accordance with the Volmer– the rapid increase of effective surface area. At this time, the growth
Weber mode with negative misfit.34 Moreover, the exchange cd of of mossy zinc was observed. Thus, addition of arsenic facilitates
hydrogen reduction on iron is ten times higher than that on zinc.37 mossy initiation.
Much higher hydrogen evolution was observed on iron. Nucleation In the presence of indium, bismuth, and lead additives, the cd
C360 Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲

Figure 3. Effect of grit size: 共a兲 10 mA cm−2, 5 min; 共b兲 40 mA cm−2,


5 min. Figure 4. Effect of acid etch on Zn deposit morphology: 共a兲 13 mA cm−2,
5 min; 共b兲 30 mA cm−2, 20 min.

remained constant for long plating times 共Fig. 7兲, indicating that the
mossy formation was inhibited 共Fig. 6兲. Addition of indium did not
influence cd dramatically but did cause smooth large grain deposits. concentration of zincate and KOH, temperature, and stirring on de-
Therefore, bismuth, indium, and lead are suppressors of den- posit morphology were investigated to construct morphology dia-
drites 共in agreement with previous literature兲 as well as inhibitors of grams. These diagrams not only graphically illustrate different mor-
mossy deposits. Arsenic is unusual in that it is an accelerator for phologies developing at different conditions, but also serve as a
mossy formation but has little effect on dendritic growth. quantitative reference for morphological control.
As demonstrated in Fig. 9, the change of the three most distinct
Effect of gelled electrolyte.— In a gelled electrolyte, the solution structures with time and cd, as well as with other deposition condi-
viscosity is so high that diffusion control should dominate. Figure 8a tions, is plotted. Heavy spongy is not included because the cd at
shows deposits obtained at a cd of 5 mA cm−2. Bulk deposits were which it develops is out of the cd range studied. Layerlike and
layerlike islands growing laterally, which is typical of a low-cd boulder deposits are reported as a compact deposit.
growth mode.18 On the other hand, dendrites grew on top of the
layerlike islands vertically, which is typical of a diffusion-controlled
growth mode.18 In nongelled systems, only one of these two growth Effect of zincate concentration.— From 0.7 M zincate solution
modes dominates, but in the gelled electrolyte, both can be seen. In 共in the middle of Fig. 9兲, dendrites develop at a cd higher than
addition, dendrites initiated from the edges of the islands and grew 32 mA cm−2. Mossy deposits form at a cd lower than 15 mA cm−2
laterally, which is similar to thin layer growth.38,39 Dendrites possi- and at a higher cd when time increases.
bly followed a horizontal diffusion direction because the vertical At a higher zincate concentration of 0.85 M, both boundaries
diffusion path was blocked by hydrogen bubbles trapped in the move to a higher cd, particularly for the dendritic-compact bound-
gelled electrolyte. ary. At a lower zincate concentration of 0.35 M, both boundaries
As shown in Fig. 8b, at 2 mA cm−2 mossy deposits developed on move to a lower cd, leaving a much smaller room for the growth of
layerlike deposits. This low cd causes very selective nucleation site compact deposits.
growth. For instance, whiskers were observed to develop preferen-
tially on the grain at the upper left corner in Fig. 8b. This grain
appears to be the basal plane. Effect of temperature.— At a higher temperature of 50°C, den-
drites develop at a much higher cd of 63 mA cm−2, and mossy de-
Morphology diagram.— In addition to the qualitative character- posits develop onto compact deposits after about 30 min. At a lower
ization of the effects of surface preparation, substrate, additive, and temperature of 0°C, a cd higher than 15 mA cm−2 induces dendritic
gelled electrolyte, the quantitative effects of cd, deposition time, formation.
Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲 C361

Discussion
The cd at dendritic-compact boundaries 共Fig. 9兲 was found to
increase linearly with increased zincate concentration, and the den-
dritic growth was always accompanied with visible hydrogen evo-
lution. The cd at the boundary 共measured apparent cd兲 is therefore
expected to be the limiting diffusion cd. This is in agreement with
the conclusion from Diggle.3 Substituting D = 1.21
⫻ 10 cm S 共estimated from CV experiments兲 into nFD/␦,
−5 2 −1

which is the slope determined according to the equation,


ilim = nFDC0 /␦, gives diffusion layer thickness ␦ = 0.046 cm. This
number is in good agreement with the value of 0.05 cm for unstirred
solution.40 It follows from morphology diagrams that dendritic ini-
tiation is favored at high cd, low zincate concentration, high KOH
concentration, unstirred solution, or highly viscous gelled electro-
lyte. All these effects confirm dendritic initiation is under diffusion
control, which is in agreement with other works.3,17,41 Vise versa,
compact and mossy growth cannot be under diffusion control.
A disagreement arises about what is the controlling mechanism
of dendritic growth. Bockris and other workers have concluded den-
dritic growth is under activation control,3,17,39 but this is not sup-
ported by our experimental results. First, long after dendritic initia-
tion, as shown in Fig. 9, deposition conditions are still consistent
with diffusion control. If deposition conditions are deliberately
changed to those favoring activation control, e.g., low cd, dendrites
should then still grow perpendicularly to the surface if dendritic
growth is under activation control. However, Fig. 10 demonstrates
that dendrites did not continue to grow vertically. Instead, nonden-
dritic, lateral, layer-by-layer growth dominated. Most importantly,
mossy deposits were found growing on the pre-existing dendrites.
These two deposit morphologies are not found with dendritic
growth. Thus, dendritic growth did not continue under activation
control conditions.
Secondly, dendrites initiate and grow perpendicularly to the sur-
face at a thermodynamically unfavorable crystal direction, 具112̄0典,
while lateral growth of activation control is in the more favorable
具0001典 direction.18 Generally, activation control conditions do not
provide sufficient energy for dendritic growth.
Finally, dendrites always grow following the diffusion path
which is perpendicular to the substrate surface, strongly suggesting a
growth mechanism of diffusion control.
From the above arguments it is concluded that dendritic growth
is not under activation control but diffusion control. The detailed
interpretation of this growth model will be included in a following
paper.
As noted earlier, the zinc deposit morphology depends critically
on deposition conditions. Accordingly, the growth mechanism must
be linked to certain deposition conditions at which the distinct mor-
phology develops. For layerlike deposits growing at low cd, the
deposits are oriented crystals. Since the deposition is likely under
single-factor control 共activation control兲, deposits grow epitaxially
to minimize activation energy. For higher cd, the concentration gra-
Figure 5. Effect of substrate on Zn deposit morphology 共10 mA cm−2, dient becomes significant and a mixed control of diffusion and ac-
5 min兲 共a兲 on Cu, 共b兲 on Ni, and 共c兲 on Fe.
tivation is found. Under this condition, there is competition between
the growth direction of the diffusion field and lateral layer growth.
Deposits are boulder and anisotropically oriented. After further in-
crease of cd to the point when the concentration gradient is domi-
Effect of stirring.— Stirring significantly increases the cd at the nant, pure diffusion control takes place, and deposits grow in the
dendritic-compact boundary several times, which is consistent with diffusion direction perpendicular to electrode surface. This single
other works.15,16 The effect of stirring on mossy formation, however, controlling factor and the hexagonal close-packed 共hcp兲 crystal
is negligible. structure of zinc form the dominant dendritic shape of zinc. Hydro-
gen is evolved but is not rapid due to the low hydrogen exchange cd.
Effect of KOH concentration.— At a higher KOH concentration At higher cd, the stirring effect of hydrogen evolution takes con-
of 10 M 共0.7 M zincate兲, dendrites develop at a lowered cd of trol. The controlling factor is diffusion plus convection, which
22 mA cm−2. Usually, dendrites develop first from the bottom of causes the local concentration gradient to oscillate. Deposits again
electrodes where replenishment of ions is difficult and hence a become anisotropically oriented, yielding the heavy spongy depos-
higher pH value is expected. 共Reduction of 1 M zincate ions re- its.
leases 4 M OH− ions, which increases pH value locally.兲 The cd at At extremely low cd, nucleation overpotential plays a decisive
the compact-mossy boundary also drops a few milliamperes. role. Deposit initiation becomes highly selective on specific nucle-
C362 Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲

Figure 6. SEM images of deposits in the presence of additives: 共a兲 100 ppm Bi, 50 mA cm−2; 共b兲 100 ppm Bi, 10 mA cm−2; 共c兲 100 ppm Pb, 50 mA cm−2; 共d兲
100 ppm Pb, 10 mA cm−2; 共e兲 100 ppm In, 50 mA cm−2; 共f兲 100 ppm In, 10 mA cm−2; 共g兲 100 ppm As, 50 mA cm−2; and 共h兲 100 ppm As, 10 mA cm−2, 0.7 M
zincate, 7 M KOH, 25°C, unstirred.
Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲 C363

Figure 7. Plots of cd–time for Zn deposition at 40 mV.

ation sites. The deposition process is under mixed charge-transfer


and nucleation control, which disagrees with a simple mass-transfer
mechanism suggested by Jaksic.21 This leads to the anisotropic
mossy growth morphology. A detailed analysis will be presented in
a following paper.
The sequential change of crystallinity and other properties of
deposits reflects the change of the controlling mechanism. There-
fore, in addition to categorizing deposit structures according to their
physical appearance, it is revealing to categorize the deposit based

Figure 9. Morphology diagrams of Zn deposits.

on formation mechanism. As shown in Table I, dendritic and layer-


like deposits are attributed to type I, which form under a single-
factor control 共spongy deposits have similar formation mechanism
as layerlike兲; heavy spongy and boulder deposits are attributed to
type II, which develop under mixed control factors. This connection
of the controlling mechanism and morphology will improve our un-
derstanding of the morphology development with the change of
deposition conditions.

Figure 8. Effect of gelled electrolyte: 共a兲 5 mA cm−2, 0.5 h; 共b兲 2 mA cm−2, Figure 10. Mossy growth after dendritic initiation at 50 mA cm−2, 5 min
1 h. followed by 10 mA cm−2, 10 min.
C364 Journal of The Electrochemical Society, 153 共5兲 C357-C364 共2006兲

Conclusions 9. R. F. Thornton and E. J. Carlson, J. Electrochem. Soc., 127, 1448 共1980兲.


10. J. T. Nichols, F. R. McLarnon, and E. J. Cairns, Chem. Eng. Commun., 37, 355
A diverse range of zinc deposits can be generated from alkaline 共1985兲.
zincate solutions, which are grouped into five categories: heavy 11. E. G. Gagnon, J. Electrochem. Soc., 133, 1989 共1986兲.
spongy, dendritic, boulder, layerlike, and filamentous mossy. 12. E. G. Gagnon and Y. Wang, J. Electrochem. Soc., 134, 2091 共1987兲.
13. E. G. Gagnon, J. Electrochem. Soc., 138, 3173 共1991兲.
Of type I deposits, dendritic initiation and growth are under dif- 14. T. C. Adler, F. R. McLarnon, and E. J. Cairns, J. Electrochem. Soc., 140, 290
fusion control while layer initiation is under activation control. 共1993兲.
Morphology diagrams were developed to graphically and quan- 15. R. D. Naybour, J. Electrochem. Soc., 116, 520 共1969兲.
titatively delineate deposit morphology change from one type to 16. C. Chen, J. Electrochem. Soc., 137, 212C 共1990兲.
another under various deposition conditions. They also display the 17. J. O’M. Bockris, Z. Nagy, and D. Drazic, J. Electrochem. Soc., 120, 30 共1973兲.
18. B. Lamping and T. O’Keefe, Metall. Trans. B, 7B, 551 共1976兲.
variables that are correlated and which can be decisive on the final 19. C. Cachet, B. Saidani, and R. Wiart, J. Electrochem. Soc., 138, 678 共1991兲.
structure of deposits. 20. M. M. Jaksic, J. Electroanal. Chem. Interfacial Electrochem., 24, 193 共1985兲.
A universal overpotential or cd that controls the initiation of 21. M. M. Jaksic, Surf. Coat. Technol., 28, 113 共1986兲.
dendrites alone does not exist. Instead, any conditions that can drive 22. M. M. Jaksic, Surf. Coat. Technol., 29, 95 共1986兲.
the reaction into diffusion control can trigger dendritic initiation. 23. C. Cachet, Z. Chami, and R. Wiart, Electrochim. Acta, 32, 465 共1987兲.
24. J. McBreen and E. Gannon, Electrochim. Acta, 26, 1439 共1981兲.
Low temperature, low zincate, or high KOH concentration, no stir- 25. F. Mansfeld and S. Gilman, J. Electrochem. Soc., 117, 588 共1970兲.
ring, and high viscosity can all dramatically reduce the minimum 26. F. Mansfeld and S. Gilman, J. Electrochem. Soc., 117, 1328 共1970兲.
dendritic initiation overpotential from 70–90 mV to 20–30 mV. 27. J. Diggle and A. Damjanovic, J. Electrochem. Soc., 117, 65 共1970兲.
28. J. Zhu, Y. Zhou, and C. Gao, J. Power Sources, 72, 231 共1998兲.
Acknowledgments 29. Y. Ein-Eli and M. Auinat, J. Electrochem. Soc., 150, A1606 共2003兲.
30. R. Renuka, S. Ramamurthy, and K. Muralidharan, J. Power Sources, 76, 197
Financial support from Teck Cominco Metals, Limited, and Ma- 共1998兲.
terials & Manufacturing Ontario is gratefully acknowledged. 31. R. O. Hull, Monthly Review of the American Electroplaters’ Society., 26, 753
共1939兲.
Natural Sciences and Engineering Research Council of Canada 32. C. S. Barrett, Structure of Metals, 3rd ed., p. 626, McGraw-Hill, New York 共1980兲.
(NSERC) assisted in meeting the publication costs of this article. 33. B. A. Lamping and T. J. O’Keefe, Metall. Trans. B, 6B, 551 共1976兲.
34. E. Budevski, G. Staikov, and W. J. Lorenz, Electrochemical Phase Formation and
References Growth, p. 6, VCH Publishers, New York 共1996兲.
1. G. X. Zhang, Corrosion and Electrochemistry of Zinc, p. 36, Plenum Publishers, 35. H. Ohtsubo, T. Matsumoto, and Y. Ohmori, ISIJ Int., 34, 1002 共1994兲.
New York 共1996兲. 36. A. Fan, W. Tian, and M. Kurosaki, J. Uni. Sci. Tech. Beijing, 10, 35 共2003兲.
2. R. D. Naybour, Electrochim. Acta, 13, 763 共1968兲. 37. T. S. Lee, J. Electrochem. Soc., 118, 1278 共1971兲.
3. J. Diggle, A. Despic, and J. O’M. Bockris, J. Electrochem. Soc., 116, 1503 共1969兲. 38. D. P. Barkey, D. Watt, Z. Liu, and S. Raber, J. Electrochem. Soc., 141, 1206
4. F. Mansfeld and S. Gilman, J. Electrochem. Soc., 117, 1521 共1970兲. 共1994兲.
5. R. Moshtev and P. Zlatilova, J. Appl. Electrochem., 8, 213 共1978兲. 39. A. Kuhn and F. Argoul, J. Electroanal. Chem., 371, 93 共1994兲.
6. S. Arouete, K. F. Blurton, and H. G. Oswin, J. Electrochem. Soc., 116, 166 共1969兲. 40. A. J. Bard, Electrochemical Methods: Fundamentals and Applications, John Wiley,
7. D. T. Chin, R. Sethi, and J. McBreen, J. Electrochem. Soc., 129, 2677 共1982兲. New York 共2001兲.
8. J. McBreen, E. Gannon, D.-T. Chin, and R. Sethi, J. Electrochem. Soc., 130, 1641 41. K. Popov, N. Kristajic, and M. Cekerevac, in Modern Aspects of Electrochemistry,
共1983兲. Vol. 30, p. 261, Plenum Press, New York 共1996兲.

You might also like