Suppression of Dendrite Formation and Corrosion on Zinc Anode of

Novel zinc anodes are synthesized via electroplating with organic additives .... The separator used in the batteries was absorptive glass mat (AGM, NS...
0 downloads 0 Views 3MB Size
Research Article www.acsami.org

Suppression of Dendrite Formation and Corrosion on Zinc Anode of Secondary Aqueous Batteries Kyung E. K. Sun, Tuan K. A. Hoang, The Nam Long Doan, Yan Yu, Xiao Zhu, Ye Tian, and P. Chen* Department of Chemical Engineering and Waterloo Institute for Nanotechnology, University of Waterloo, 200 University Avenue West, Waterloo, Ontario N2L3G1, Canada

ACS Appl. Mater. Interfaces 2017.9:9681-9687. Downloaded from pubs.acs.org by LANCASTER UNIV on 01/14/19. For personal use only.

S Supporting Information *

ABSTRACT: Novel zinc anodes are synthesized via electroplating with organic additives in the plating solution. The selected organic additives are cetyltrimethylammonium bromide (CTAB), sodium dodecyl sulfate (SDS), polyethylene-glycol (PEG-8000), and thiourea (TU). The synthesized zinc anode materials, namely, Zn-CTAB, Zn-SDS, Zn-PEG, and Zn-TU, are characterized by powder X-ray diffraction and scanning electron microscopy. The results show that each additive produces distinctively different crystallographic orientation and surface texture. The surface electrochemical activity is characterized by linear polarization when the zinc is in contact with the battery’s electrolyte. Tafel fitting on the linear polarization data reveals that the synthetic zinc materials using organic additives all exhibit 6−30 times lower corrosion currents. When using Zn-SDS as the anode in the rechargeable hybrid aqueous battery, the float current decreases as much as 2.5 times. The batteries with ZnSDS, Zn-PEG, and Zn-TU anodes display the capacity retention of 79%, 76%, and 80% after 1000 cycles of charge−discharge at 4C rate, whereas only 67% obtained from the batteries using the anode prepared from commercial zinc foil. Among these electroplated anodes, Zn-SDS is the most suitable for aqueous batteries thanks to its low corrosion rate, low dendrite formation, low float current, and high capacity retention after 1000 cycles. KEYWORDS: rechargeable battery, aqueous electrolyte, electroplating, additive, corrosion inhibitor, dendrite

1. INTRODUCTION Zinc and its alloys have been exploited as electrodes of rechargeable aqueous batteries.1−10 Zinc electrodes have a very high specific capacity of 820 mAh·g−1 and are less expensive than lithium.11,12 Furthermore, Zn metal is able to function in aqueous environment and this allows zinc batteries to use aqueous based electrolytes, which are less expensive and safer than the nonaqueous based electrolytes.13 Zinc electrodes exhibit good reversibility, high overpotential for hydrogen evolution in acidic environment. Despite the important advantages, the corrosion of zinc in acidic media and the dendrite formation are crucial issues which must be overcome.14,15 The rechargeable hybrid aqueous battery (ReHAB), introduced in 2012, employs a lithium intercalation cathode, zinc foil anode, and an aqueous electrolyte.16−19 The ReHAB exhibits an energy density of ∼50−80 Wh·kg−1, which doubles that of commercial lead-acid battery (∼30−37 Wh·kg−1).20,21 Furthermore, the ReHAB system is lead-free and it is constructed with environmentally benign materials; hence there is no flammable hazard. Despite these advantages, the anode of the ReHAB suffers from corrosion, dendrite formation, and other side reactions. Thus, it is necessary to formulate new zinc containing anodes which exhibit higher resistance to such side reactions, thus, the calendar life of the ReHAB can be improved. © 2017 American Chemical Society

The dendrite formation is a challenging and troublesome process occurring on the zinc surface. According to the literature regarding rechargeable alkaline Zn batteries, the newly created Zn like to deposit on the dislocated places on the electrode surface, and this is a linear-diffusion control process, which changes to spherical-diffusion controlled.22 The dendrites are normally in the needle-shape and growing up continuously on the surface of the Zn surface when the batteries are working.23−25 Eventually the dendrites will reach the cathode and the battery will have a sudden short circuit or there may be a sudden drop in capacity to near zero. During the dendrite growth, the surface of the anode increases and this causes an increase in the surface area. Therefore, the corrosion and other surface dependent reactions are augmented, lead to a faster decrease of battery performance. Thus, inhibition of dendrite growth is a crucial task. Previous studies proved that dendrites can be destroyed by slight overcharging of aqueous batteries so that the oxygen gas can be generated and react with the dendrites.26 However, this is a temporary method because the oxygen gas is generated as the expense of water consumption from the aqueous electrolyte. Furthermore, the oxidized products created from the reaction of Zn and O2 are Received: December 23, 2016 Accepted: February 27, 2017 Published: February 27, 2017 9681

DOI: 10.1021/acsami.6b16560 ACS Appl. Mater. Interfaces 2017, 9, 9681−9687

Research Article

ACS Applied Materials & Interfaces generally insulating, and the Zn surface is slowly passivated. It is an important mandate that an effective dendrite suppression method must be discovered. The literature regarding dendrite elimination from Zn deposits is abundant, but only for alkaline Zn batteries. There is little knowledge about dendrite suppression for secondary Zn batteries using neutral or mild acidic media as electrolytes. The successful usage of dendrite suppression additives directly in the electrolytes of rechargeable zinc batteries is limited because the additives must not only limit the dendrite but also improve or preserve the battery performance. In this regard, the effects of Pb2+, sodium lauryl sulfate, and Triton X-100 additives on the suppression of dendrite on the Zn anode of the Znpolyaniline batteries were reported. These compounds were used as additive in the electrolytes and the scanning electron microscopy (SEM) images of the postrun Zn anodes revealed that the Zn surface is more flat if appropriate additives were used.27 Tetra-alkyl ammonium hydroxides were used as the additive in the electrolyte of the Zn−Ni batteries. The electrochemistry at the Zn-electrolyte interface was characterized by a time-current profile (similar to chronoamperometry), and the results suggested that each of the tetraalkyl ammonium hydroxide had an optimal concentration at a predetermined charge/discharge current of the Zn−Ni battery. Even though the surfaces of Zn electrodes were flat at most of the additive concentrations (suggesting that dendrite formation was eliminated), the battery just worked effectively at an optimal concentration of each additive.28 In a recent study, we exploited the use of electroplating to synthesize new zinc anodes.29 The plating condition has been optimized and several inorganic additives have been utilized. We find that the surface texture and the surface crystal structure can be tuned so that the synthesized zinc exhibits higher dendrite resistance and higher corrosion tolerance−up to an 8fold decrease in corrosion current density, as determined from a Tafel fit on the linear polarization data. The batteries assembled from such electroplated zinc anodes all out-perform the ones using commercial zinc foil as the anode. In this study, we investigate the roles of organic additives in the electroplating process. The synthesized zinc samples are used as the anodes in the rechargeable hybrid aqueous battery. The possibility of dendrite formation is evaluated. We find out that this is not only closely related to surface morphology, as reported in literature, but also associated with surface crystallographic orientation. We believe that this is the first time the cohesive effect of the surface properties of a Zn anode toward battery performance is evaluated and explained at fundamental level. Furthermore, the corrosion rate of the prepared zinc electrodes when in contact with the electrolyte are 30 times lower than that of the commercial zinc foil. The batteries assembled from such zinc electrodes exhibit smaller float charge currents, lower open-circuit voltage drops, and higher capacity retention after 1000 cycles than those with the commercial zinc foil.

A parallel cell was used to deposit zinc ions with 100 mL of the electroplating solution by placing the anode, in this case, zinc foil (Rotometals, thickness 0.2 mm) and the cathode, brass foil, 5 cm apart facing each other. The current density of 65 mA·cm−2 was applied for 10 min by using a BK Precision instrument. After the synthesis, the samples were dried overnight in air. The brass foil was pretreated with nitric acid (HNO3) before it was used as the substrate to deposit zinc ions on. The foil was dipped into 25 vol % HNO3 solution for 5 s, and washed with deionized (DI) water. After the treatment, it was dried in air. 2.2. Preparation of Separators, Cathodes, Electrolytes, and Batteries. The cathode for the battery was synthesized by mixing 86 wt % of LiMn2O4 (MTI Co.), 7 wt % KS-6 graphite (Timcal), and 7 wt % polyvinylidene fluoride (PVDF, Kynar, HSV900) in n-methyl-2pyrrolidone (NMP, Sigma-Aldrich Co., 99.5% purity) for 2 min with Planetary Centrifugal Mixer (AR-100, ThinkUSA). The slurry was cast on a graphite foil (Alfa Aesar, 99.8%) and placed in 60 °C vacuum chamber for 3 h. The dried slurry was cut in a circle with 12 mm diameter. The battery electrolyte was prepared by dissolving 1 M zinc sulfate heptahydrate (Alfa Aesar Co., 98%) and 2 M lithium sulfate (SigmaAldrich, 98%) in DI water. The pH of the solution was adjusted to 4.00 ± 0.05 using sulfuric acid or lithium hydroxide solution. The separator used in the batteries was absorptive glass mat (AGM, NSG Corporation). Two types of batteries were employed: coin cell and two-electrode SwagelokTH-type cells. Coin cell types were used for cyclability test, and SwagelokTH-type for the float current test. 2.3. Instrumentation. D8 Discover Powder X-ray diffractometer (XRD, Brüker Co., CuKα 1.5406 Å, 40 kV, and 40A) was used to identify the crystallinity of the samples. The samples were scanned (exsitu) over 2 theta range of 10−90° at the rate of 0.003 deg min−1 with LynxEye detector. The XRD results were analyzed with the Brüker XRD search and match program EVATH. Field emission scanning electron microscopy (FE-SEM, Carl Zeiss Ultra Plus Field Emission SME, Zeiss Co.), operated at 10 kV, was used to observe the morphology of the samples. VMP3 potentiostat/galvanostat (Bio-Logic Science Instrument Co.) was used to analyze the corrosion of the synthesized samples and the commercialized zinc (Rotometals, thickness 0.2 mm) using a threeelectrode cell system. The working (WE), counter (CE), and reference (REF) electrodes were zinc, platinum, and Hg/Hg2SO4, respectively. Linear polarization technique was applied to the system by scanning between −0.25 and 0.25 V vs Ewe from its open circuit voltage (OCV) at the rate of 0.166 mV/s. The surface area of the tested electrode was ∼1 cm2. It was used to convert corrosion current density (icorr) to corrosion current (Icorr). 2.4. Electrochemical Measurements. Coin cell type was used to test the cycle life of the batteries with the synthesized and the commercialized anodes. The galvanostatic charge−discharge cycling of the batteries was carried out between 1.4 and 2.1 V vs Zn2+/Zn, in constant current−constant voltage (CC−CV) mode at room temperature at 4 C rate (1 C = 120 mA·g−1) with NEWARE battery tester (NEWARE Battery Test System, Neware Co. Ltd., China). The current calculation was referred to the mass loading of LiMn2O4 in the cathode. The Zn anode was used in excess. The current cutoff during constant voltage charging was set at 10% of the charging current in the constant current charging step at 4 C rate. The two-electrode SwagelokTH-type cells were used to perform the float current test with the same battery tester at room temperature. The potential of the batteries was maintained at 2.1 V for 7 days, and the necessary current to maintain the potential was recorded.

2. EXPERIMENTAL SECTION

3. RESULTS AND DISCUSSION 3.1. Effect of Organic Additives. The XRD results of the synthesized zinc with and without the organic additives and the commercialized zinc (labeled as the “Commercial”) are shown in Figure 1. According to the XRD results of the synthesized and the commercialized zinc, the zinc crystals grow in various orientations. On the XRD pattern of the commercialized zinc

2.1. Preparation of Anodes. Zinc sulfate (0.6 M; Alfa Aesar, 98%), 0.1 M ammonium sulfate (Sigma-Aldrich, 99%), and 100 ppm of organic additive were dissolved in DI water. The selected additives were: cetyltrimethylammonium bromide (CTAB, Sigma-Aldrich, 98%), sodium dodecyl sulfate (SDS, Sigma-Aldrich, 99%), polyethylene glycol 8000 (PEG, Sigma-Aldrich, MW 8000, 99%), and thiourea (Alfa Aesar, 99%). 9682

DOI: 10.1021/acsami.6b16560 ACS Appl. Mater. Interfaces 2017, 9, 9681−9687

Research Article

ACS Applied Materials & Interfaces

preferential orientation of Zn-PEG is the same as that of Zn− No, PEG significantly reduced the peak representative in (101) planes, which is the second dominant peak of Zn−No. According to Mackinnon and Fenn, the preferred orientation of (103) represents zinc deposition with lower current efficiency compared with (101) orientation.35 This phenomenon is observed when using polyethylene glycol (MW = 2000) additive.37 This may suggest that the (103) plane is more inert than the (101). However, the crystal growth on the (103) plane is nearly horizontal with the electrode surface while the (101) supports the growth with ∼70° to the surface of the electrode, thus dendrite formation on the (103) is much less likely.36 Zn-CTAB shows very low (002) and (103) peak intensities while the (100), (110) present. The crystal growth on the (100) and (110) are perpendicular to the sample surface, so the dendrite formation on the Zn-CTAB is very likely. The Zn-SDS does not possess (100) and (110) planes but the (101) facet is high, thus, the dendrite formation on this material is still likely, but it will not be as significant as the Zn-CTAB. According to XRD results, the possibility of dendrite formation on the ZnSDS and the Zn−No is much lower than the commercial Zn and the Zn-CTAB. The XRD of Zn-TU indicates that some of the planes are oriented in (102) and (103), but their intensities are much lower than (101): in other words, thiourea enhances the zinc growth only in one orientation. The deposition of Zn on (101) has high current efficiency, so it can be expected that the ZnTU will exhibit less polarization when in battery, but the dendrite formation is still possible on this kind of anode. On the XRD pattern of the Zn-PEG, (002) and (103) planes are dominant. Thus, dendrite formation on this material is less likely because (002) and (103) support basal morphology growth of crystals. The morphology of the commercialized zinc is shown in Figure 2. The surface of the commercialized zinc is flat and smooth with defects (holes), which may be produced during its manufacturing process. Figure 2 also shows the SEM images of the electroplated zinc with the organic additives. The morphology of Zn-CTAB is porous needle-like crystals with uniform size distributed evenly. This kind of structure is obtained due to a strong blocking effect of CTAB (e.g., CTAB surfactant molecules are absorbed onto the active sites, which affect the kinetics of the electron transfer) that increases the competition between the nucleation and crystal growth.30 On the other hand, the morphologies of the Zn-SDS and Zn-PEG are regular and uniformly distributed. The zinc deposits are growing perpendicular to the substrate and in various directions. This kind of structure is obtained when the active site and nucleation rate are reduced as additives are being adsorbed onto the surface of a substrate.30,38 As for Zn-TU, the deposit is also uniform and regular, but is not compact as ZnSDS and Zn-PEG. Some of the areas are empty possibility due to the hydrogen gas presented on the surface, blocking the zinc deposition. On the SEM of the Zn−No, the surface contains irregular layer structure, but some of the features (e.g., at the bottom right place) are growing perpendicular to the surface and this may affect the separator of the battery when in contact. The batteries have been charged to full capacity at 2.1 V under constant current charging mode, then they are floatcharged constantly at 2.1 V for 7 days. The recorded currents are called the float charge currents, which can be used to evaluate the currents or the energies required to keep the batteries at 100% state-of-charge. Higher currents mean more

Figure 1. XRD results of the zinc anode electroplated with and without organic additives and commercialized zinc.

foil, the highest intensity peak is found at 42°, indicating the zinc growth is mostly in (101) orientation. The other significant planes are (102), (103), (100), (002), (110), (112), (200), and (201). These peaks are identical with the literature.30,31 When depositing the zinc without any additives, labeled as Zn−No in Figure 1, the growth orientation of this material is mostly directed in (103). The peak locations of the electroplated zinc with the organic additives, shown in Figure 1, are identical to the commercialized zinc foil and literature.30,31 The peak intensities of all patterns are normalized, taking the intensity of the (101) peak as the reference. The XRD patterns of the Zn-CTAB and brass foil are provided (Figure S1 and S2) for further judgment. From Figure S2, the intensity ratio between XRD peaks at 42.29° and 72.24° is 5.23, while on Figure S1, this ratio is about 1.8. This means the peak at 72.2° is mainly contributed by Zn (110). This is reasonable because this Zn (110) peak intensity is varies among patterns. If it is solely from brass, it must have an identical intensity. The intensities of the Zn peaks, however, are not the same, indicating different preferred directions of crystal growth−each of the organic additive produces a unique crystalline surface structure. The variation of the surface structure is due to the ability of the additives to alter the surface energy. Thus, the crystal growth is favored in a new lowest surface energy.30 Furthermore, the experimental procedure may be another factor contributing to the variations in XRD peak intensity.30 The XRD results indicate that the usage of CTAB, SDS, and thiourea during the zinc deposition (Zn-CTAB, Zn-SDS, ZnTU) produces a strong orientation at (101), which is also seen in literature.32−34 The dominant (101) peak is also found in the XRD of the commercialized zinc foil. The changes in the preferential growth from (103) to (101) (Zn−No to Znorganic additive) suggests that CTAB, SDS, and thiourea modify the electroplated anode, rendering the likely crystallographic surface to the commercial zinc foil. Also, Mackinnon and Fenn explains that having the preferential growth of zinc in (101) implies that high current efficiency is achieved when zinc ions are deposited.35 This finding is in accordance with D.J. Mackinnon et al. provided data of high zinc deposition efficiency obtained with organic additives.36 On the XRD pattern of the zinc electroplated with PEG (ZnPEG), zinc crystalline is mostly oriented in (103). Although the 9683

DOI: 10.1021/acsami.6b16560 ACS Appl. Mater. Interfaces 2017, 9, 9681−9687

Research Article

ACS Applied Materials & Interfaces

current from the battery with Zn-TU anode is about the same as that with the commercialized zinc foil. Moreover, the float current of the battery using Zn−No is similar to that with ZnCTAB and Zn-TU. Hence, the modifications of the surface of zinc with CTAB and thiourea are not as effective as with SDS and PEG in this regard. The linear polarization curves of the synthesized zinc with and without the organic additives and commercialized zinc are presented in Figure 4. The corrosion potential and the

Figure 2. SEM images of synthesized anode with and without organic additives and commercialized zinc (magnification 5k). CTAB, SDS, PEG, thiourea: Electroplated Zn using CTAB, SDS, PEG, or Thiourea containing electrolyte, respectively. Commercial: Commercial zinc foil material. No additive: Electroplated zinc without using any additive in the electroplating bath.

Figure 4. Linear polarization curves of the synthesized zinc with and without organic additives and commercialized zinc foil when in contact with the battery’s electrolyte.

energies are required for the purpose. This method has been used extensively in the literature to evaluate batteries functioning at nearly full state-of-charge.39−44 Figure 3 Table 1. Corrosion Potential and Current for the Synthesized Zinc with and without Organic Additives and Commercialized Zinc Foil corrosion potentiala (mV) no additive CTAB SDS PEG thiourea commercial

4 2 2 11 10 0

± ± ± ± ± ±

5 2 5 8 8 2

corrosion current (μA·cm−2) 1136 220 75 163 43 1422

± ± ± ± ± ±

82 7 14 32 7 110

a

The corrosion potential of the commercial Zn is chosen as the standard.

respective corrosion currents, tabulated in Table 1, are obtained via applying Tafel fit on the data. As shown in the table, both Zn−No and the commercialized zinc exhibit high corrosion currents, which may be due to the outcome of the crystalline structure. The XRD data shows that both samples contain two highest peaks at (101) and (103) planes. Hence, these two planes may be susceptible to high corrosion rates. The Zn-SDS and Zn-PEG also have high peaks at (101) and (103), but they have low reactive (002) facets on the surface, which may be the reason why the Zn-SDS and Zn-PEG have low corrosion rates. On the basis of the corrosion data, the corrosion current is increased in the order of Zn-SDS, Zn-PEG, and Zn-CTAB. Thus, the corrosion rates follow the same increasing trend as the float current results (Zn-CTAB > Zn-PEG > Zn-SDS). The XRD peaks reflecting (100) and (110) crystallographic facets also decrease in the same order. Among the three samples, the

Figure 3. Float current of batteries with zinc deposited with/without organic additives and commercialized zinc foil.

represents the plot of float charge current of the batteries using the synthesized zinc anodes with and without the use of organic additives and the commercialized zinc foil. The lowest and the highest float currents are obtained with Zn-SDS and Zn-TU, respectively. Hence, the side reactions (e.g., hydrogen evolution) are effectively slowed down with Zn-SDS anode, but not with Zn-TU. The next lowest float current is observed from the battery with Zn-PEG anode. The battery with Zn-CTAB produces higher float current than that with Zn-PEG due to the higher surface area of ZnCTAB anode; but it is lower than that with Zn-TU. The float 9684

DOI: 10.1021/acsami.6b16560 ACS Appl. Mater. Interfaces 2017, 9, 9681−9687

Research Article

ACS Applied Materials & Interfaces

(002) zinc has more resistance against dendrite formation. When dendrites grow, the surface of zinc anode is increased, and thus, the corrosion effect will be more severe. This may explain the good battery performance of the Zn-SDS and ZnPEG materials due to low possibility of dendrite formation and low corrosion currents. In general, hydrogen evolution on these anodes is decreased as well thanks to the upper shifts of equilibrium corrosion potentials. This is because the corrosion potential is where the Zn2+ deposition and the Zn dissolution are balance and the H2 evolution potential is about 0.4−0.5 V more negative than the Zn2+/Zn potential at the battery working condition.16 An upper move of such potential would allow a father distance from H2 evolution zone. Therefore, except for Zn-CTAB, all the electroplated samples are able to solve the problems associated with the commercialized zinc foil by reducing the float current and the corrosion rate. Among the organic additives, SDS is the most preferred as it reduces the side reactions (float current, dendrite formation, and corrosion) and the capacity of the battery is well maintained until the 1000th cycle.

highest corrosion is observed from Zn-CTAB. This result is expected due to the increased surface area, shown in the SEM images, because corrosion rate observed on a surface is proportional to its surface area. As for Zn-TU anode, the smallest corrosion current and highest float current are obtained. In other words, the Zn-TU anode cannot solve the problem of hydrogen evolution when the battery is in the charged state (in the float current experiment), but it is still able to reduce the corrosion rate on the anode when the batteries are in idle states (e.g., storage of batteries at different state of charges or during open-circuit voltage monitoring). K. Song et al. determined that thiourea acts as an inhibitor of hydrogen evolution but yet promotes hydrogen absorption on the Zn deposits..37 Hence, the lower corrosion rate on the Zn-TU is reasonable because it is the counter reaction of the hydrogen evolution, it may be further explained by the lower amount of zinc deposited compared to that in other samples, and the higher float current with the higher amount of adsorbed hydrogen in the deposits (desorbing out and forming hydrogen gas inside the battery). After the above experiments, the coin batteries were assembled with the synthesized zinc, and their performance is summarized in Figure 5. Voltage profiles of all batteries are

4. CONCLUSION The anodes synthesized with organic additives mitigate some major problems associated with the current zinc foil in the battery. We have proved that the crystallographic properties and morphology of the Zn surface can be adjusted by using different additives in the electroplating process. When Zn has strong (100) and (110) surface planes, dendrite formation is likely. When Zn surface is dominated by (002) and (103) facets, dendrite formation is less likely. Furthermore, denser surface may support less corrosion and lower float currents of the Zn batteries. When these anodes are used in the coin batteries, 79%, 76%, and 80% of the capacity retained after 1000 cycles with Zn-SDS, Zn-PEG, and Zn-TU, respectively. These values are much higher than that of the coin battery assembled with the commercialized zinc foil (67%).



ASSOCIATED CONTENT

S Supporting Information *

Figure 5. Cyclability of the batteries with zinc anode with organic additives and commercialized zinc foil. Estimated errors: ± 2.5%.

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.6b16560.

plotted in Figure S3. The battery assembled from commercial Zn foil exhibits 113 mA·h·g−1, while others are from 94−103 mA·h·g−1. Howerver, all the batteries, except for the ones with Zn-CTAB, display drastically improved cyclability. The average capacity retentions of 79, 76, 80, and 67% are obtained at the 1000th cycle from the batteries with Zn-SDS, Zn-PEG, Zn-TU, and commercialized zinc foil, respectively. As for the batteries with Zn-CTAB, a sudden capacity drop occurs after running about 350 cycles, which may be the result of the formation of extended dendrites on the anode. During this capacity dropping period, the Coulombic efficiency fluctuates significantly. Otherwise this efficiency is always near 100%. At the 1000th cycle, only 12% of original capacity remains. The high surface area of Zn-CTAB possibly increases the chance of the dendrites to grow and leads to the failure of the batteries. Furthermore, the XRD peaks responsible for (100) and (110) planes are very high in the pattern of Zn-CTAB. According to Sawada, dendrites are most preferred to growth from this particular crystallographic surface.45 Zinc metal with exposed (100) and (110) planes are prone to dendrite formation while the exposed



XRD patterns of Zn-CTAB and brass foil and the voltage profiles of all batteries at 1st and 1000th cycles (PDF)

AUTHOR INFORMATION

Corresponding Author

*Tel.: +1-519-888-4567 ext. 35586. Fax: +1-519-888-4347. Email: [email protected]. ORCID

Tuan K. A. Hoang: 0000-0002-5632-1976 Author Contributions

K.E.K.S. conducted most of the experiment work under the supervision of T.N.L.D. and T.K.A.H. K.E.K.S. drafted the manuscript, which was improved by T.K.A.H. P.C. is the PI and leads the research. Y.Y., X.Z., Y.T. were helping K.E.K.S. in instrumentation. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest. 9685

DOI: 10.1021/acsami.6b16560 ACS Appl. Mater. Interfaces 2017, 9, 9681−9687

Research Article

ACS Applied Materials & Interfaces



(16) Yan, J.; Wang, J.; Liu, H.; Bakenov, Z.; Gosselink, D.; Chen, P. Rechargeable Hybrid Aqueous Batteries. J. Power Sources 2012, 216, 222−226. (17) Han, Z.; Askhatova, D.; Doan, T. N. L.; Hoang, T. K. A.; Chen, P. Experimental and Mathematical Studies on Cycle Life of Rechargeable Hybrid Aqueous Batteries. J. Power Sources 2015, 279, 238−245. (18) Lu, C. Y.; Hoang, T. K. A.; Doan, T. N. L.; Zhao, H. B.; Pan, R.; Yang, L.; Guan, W. S.; Chen, P. Rechargeable Hybrid Aqueous Batteries Using Silica Nanoparticle Doped Aqueous Electrolytes. Appl. Energy 2016, 170, 58−64. (19) Lu, C. Y.; Hoang, T. K. A.; Doan, T. N. L.; Acton, M.; Zhao, H. B.; Guan, W. S.; Chen, P. Influence of Different Silica Gelling Agents on the Performance of Aqueous Gel Electrolytes. J. Ind. Eng. Chem. 2016, 42, 101−106. (20) Rydh, C. J. Environmental Assessment of Vanadium Redox and Lead-Acid Batteries for Stationary Energy Storage. J. Power Sources 1999, 80, 21−29. (21) Hadjipaschalis, I.; Poullikkas, A.; Efthimiou, V. Overview of Current and Future Energy Storage Technologies for Electric Power Applications. Renewable Sustainable Energy Rev. 2009, 13, 1513−1522. (22) Diggle, J. W.; Despic, A. R.; Bockris, J. O’M. The Mechanism of the Dendritic Electrocrystallization of Zinc. J. Electrochem. Soc. 1969, 116, 1503−1514. (23) Banik, S. J.; Akolkar, R. Suppressing Dendrite Growth during Zinc Electrodeposition by PEG-200 Additive. J. Electrochem. Soc. 2013, 160, D519−D523. (24) Yuan, Y. F.; Tu, J. P.; Wu, H. M.; Wang, S. F.; Zhang, W. K.; Huang, H. Effects of Stannous Ions on the Electrochemical Performance of the Alkaline Zinc Electrode. J. Appl. Electrochem. 2007, 37, 249−253. (25) Banik, S. J.; Akolkar, R. Suppressing Dendritic Growth during Alkaline Zinc Electrodeposition using Polyethylenimine Additive. Electrochim. Acta 2015, 179, 475−481. (26) McLarnon, F. R.; Cairns, E. J. The Secondary Alkaline Zinc Electrode. J. Electrochem. Soc. 1991, 138, 645−664. (27) Kan, J. Q.; Xue, H. G.; Mu, S. L. Effect of inhibitors on Zndendrite formation for zinc-polyaniline secondary battery. J. Power Sources 1998, 74, 113−116. (28) Lan, C. J.; Lee, C. Y.; Chin, T. S. Tetra-alkyl ammonium hydroxides as inhibitors of Zn dendrite in Zn-based secondary batteries. Electrochim. Acta 2007, 52, 5407−5416. (29) Sun, K. E. K. Synthesis of Novel Zinc Anode via Electroplating for Rechargeable Hybrid Aqueous Batteries. Master Thesis, University of Waterloo, Canada, 2016. (30) Gomes, A.; da Silva Pereira, M. I. Pulsed Electrodeposition of Zn in the Presence of Surfactants. Electrochim. Acta 2006, 51, 1342− 1350. (31) Nayana, K. O.; Venkatesha, T. V. Bright zinc Electrodeposition and Study of Influence of Synergistic Interaction of Additives on Coating Properties. J. Ind. Eng. Chem. 2015, 26, 107−115. (32) Tripathy, B. C.; Das, S. C.; Singh, P.; Hefter, G. T. Zinc Electrowinning from Acidic Sulphate Solutions. Part III: Effects of Auaternary Ammonium Bromides. J. Appl. Electrochem. 1999, 29, 1229−1235. (33) Ghavami, R. K.; Rafiei, Z.; Tabatabaei, M. Effects of Cationic CTAB and Anionic SDBS Surfactants on the Performance of Zn− MnO2 Alkaline Batteries. J. Power Sources 2007, 164, 934−946. (34) Li, M. C.; Jiang, L. L.; Zhang, W. Q.; Qian, Y. H.; Luo, S. Z.; Shen, J. N. Electrodeposition of Nanocrystalline Zinc from Acidic Sulfate Solutions Containing Thiourea and Benzalacetone as Additives. J. Solid State Electrochem. 2007, 11, 549−553. (35) Mackinnon, D. J.; Fenn, P. L. The Effect of Tin on Zinc Electrowinning from Industrial Acid Sulphate Electrolyte. J. Appl. Electrochem. 1984, 14, 701−707. (36) Mackinnon, D. J.; Brannen, J. M.; Fenn, H. P. Characterization of Impurity Effects in Zinc Electrowinning from Industrial Acid Sulphate Electrolyte. J. Appl. Electrochem. 1987, 17, 1129−1143.

ACKNOWLEDGMENTS This research is financially supported by Positec Canada Ltd., Natural Sciences and Engineering Research Council of Canada (NSERC), Canadian Foundation for Innovation (CFI), Mitacs (IT04444, IT06144, IT06460), and the University of Waterloo.



ABBREVIATIONS CTAB, cetyltrimethylammonium bromide SDS, sodium dodecyl sulfate PEG, polyethylene glycol 8000 TU, thiourea Zn-CTAB, zinc anode electroplated with CTAB Zn-SDS, zinc anode electroplated with SDS Zn-PEG, zinc anode electroplated with PEG Zn-TU, zinc anode electroplated with TU



REFERENCES

(1) Deyab, M. A. Application of Nonionic Surfactant as a Corrosion Inhibitor for Zinc in Alkaline Battery Solution. J. Power Sources 2015, 292, 66−71. (2) Parker, J. F.; Chervin, C. N.; Nelson, E. S.; Rolison, D. R.; Long, J. W. Wiring Zinc in Three Dimensions Re-writes Battery PerformanceDendrite-free Cycling. Energy Environ. Sci. 2014, 7, 1117−1124. (3) Winsberg, J.; Janoschka, T.; Morgenstern, S.; Hagemann, T.; Muench, S.; Hauffman, G.; Gohy, J.-F.; Hager, M. D.; Schubert, U. S. Poly(TEMPO)/Zinc Hybrid-Flow Battery: A Novel, “Green,” High Voltage, and Safe Energy Storage System. Adv. Mater. 2016, 28, 2238− 2243. (4) Yan, C.; Wang, X.; Cui, M.; Wang, J.; Kang, W.; Foo, C. Y.; Lee, P. S. Batteries: Stretchable Silver-Zinc Batteries Based on Embedded Nanowire Elastic Conductors. Adv. Energy Mater. 2014, 4, 1301396. (5) Braam, K. T.; Volkman, S. K.; Subramanian, V. Characterization and Optimization of a Printed, Primary Silver−Zinc Battery. J. Power Sources 2012, 199, 367−372. (6) Zhang, B.; Liu, Y.; Wu, X.; Yang, Y.; Chang, Z.; Wen, Z.; Wu, Y. P. An Aqueous Rechargeable Battery Based on Zinc Anode and Na0.95MnO2. Chem. Commun. 2014, 50, 1209−1211. (7) Xu, C. J.; Li, B. H.; Du, H.; Kang, F. Energetic Zinc Ion Chemistry: The Rechargeable Zinc Ion Battery. Angew. Chem., Int. Ed. 2012, 51, 933−935. (8) Xu, Y.; Zhang, Y.; Guo, Z.; Ren, J.; Wang, Y.; Peng, H. Flexible, Stretchable, and Rechargeable Fiber-Shaped Zinc−Air Battery Based on Cross-Stacked Carbon Nanotube Sheets. Angew. Chem., Int. Ed. 2015, 54, 15390−15394. (9) Lu, X.; Li, G.; Kim, J. Y.; Lemmon, J. P.; Sprenkle, V. L.; Yang, Z. A Novel Low-cost Sodium−Zinc Chloride Battery. Energy Environ. Sci. 2013, 6, 1837−1843. (10) Park, J.; Park, M.; Nam, G.; Lee, J. − S.; Cho, J. All-Solid-State Cable-Type Flexible Zinc−Air Battery. Adv. Mater. 2015, 27, 1396− 1401. (11) Yu, X. W.; Licht, S. High Capacity Alkaline Super-Iron Boride Battery. Electrochim. Acta 2007, 52, 8138−8143. (12) Lee, J.-S.; Tai Kim, S.; Cao, R.; Choi, N. − S.; Liu, M.; Lee, K. T.; Cho, J. Metal-Air Batteries: Metal−Air Batteries with High Energy Density: Li−Air versus Zn−Air. Adv. Energy Mater. 2011, 1, 34−50. (13) Trócoli, R.; La Mantia, F. An Aqueous Zinc-Ion Battery Based on Copper Hexacyanoferrate. ChemSusChem 2015, 8, 481−485. (14) Hoang, T. K. A.; Doan, T. N. L.; Sun, K. E. K.; Chen, P. Corrosion Chemistry and Protection of Zinc & Zinc Alloys by Polymer-containing Materials for Potential Use in Rechargeable Aqueous Batteries. RSC Adv. 2015, 5, 41677−41691. (15) Nakata, A.; Murayama, H.; Fukuda, K.; Yamane, T.; Arai, H.; Hirai, T.; Uchimoto, Y.; Yamaki, J. − I.; Ogumi, Z. Transformation of Leaf-like Zinc Dendrite in Oxidation and Reduction Cycle. Electrochim. Acta 2015, 166, 82−87. 9686

DOI: 10.1021/acsami.6b16560 ACS Appl. Mater. Interfaces 2017, 9, 9681−9687

Research Article

ACS Applied Materials & Interfaces (37) Song, K.-D.; Kim, K.-B.; Han, S.-H.; Lee, H. K. Effect of Additives on Hydrogen Evolution and Absorption during Zn Electrodeposition Investigated by EQCM. Electrochem. Solid-State Lett. 2004, 7, C20. (38) Trejo, G.; Ruiz, H.; Ortega Borges, R.; Meas, Y. Influence of Polyethoxylated Additives on Zinc Electrodeposition from Acidic Solutions. J. Appl. Electrochem. 2001, 31, 685−692. (39) Dufo-López, R.; Lujano-Rojas, J. M.; Bernal-Agustín, J. L. Comparison of different lead-acid battery lifetime prediction models for use in simulation of stand-alone photovoltaic systems. Appl. Energy 2014, 115, 242−253. (40) Niu, J. J.; Conway, B. E.; Pell, W. G. Comparative studies of selfdischarge by potential decay and float-current measurements at C double-layer capacitor and battery electrodes. J. Power Sources 2004, 135, 332−343. (41) Fernandez, M.; Ruddell, A. J.; Vast, N.; Esteban, J.; Estela, F. Development of a VRLA battery with improved separators, and a charge controller, for low cost photovoltaic and wind powered installations. J. Power Sources 2001, 95, 135−140. (42) Ruetschi, P. Aging mechanisms and service life of lead-acid batteries. J. Power Sources 2004, 127, 33−44. (43) Bullock, K. R. Carbon reactions and effects on valve-regulated lead-acid (VRLA) battery cycle life in high-rate, partial state-of-charge cycling. J. Power Sources 2010, 195, 4513−4519. (44) Zhao, H. B.; Hu, C. J.; Cheng, H. W.; Fang, J. H.; Xie, Y. P.; Fang, W. Y.; Doan, T. N. L.; Hoang, T. K. A.; Xu, J. Q.; Chen, P. Novel Rechargeable M3V2(PO4)3//Zinc (M = Li, Na) Hybrid Aqueous Batteries with Excellent Cycling Performance. Sci. Rep. 2016, 6, 25809. (45) Sawada, Y. Transition of Growth Form from Dendrite to Aggregate. Phys. A 1986, 140, 134−141.

9687

DOI: 10.1021/acsami.6b16560 ACS Appl. Mater. Interfaces 2017, 9, 9681−9687