Você está na página 1de 8

Journal of Chemical Technology and Biotechnology

J Chem Technol Biotechnol 79:927934 (online: 2004) DOI: 10.1002/jctb.1082

Recycling of nickelmetal hydride batteries. II: Electrochemical deposition of cobalt and nickel
N Tzanetakis and K Scott
School of Chemical Engineering and Advanced Materials, University of Newcastle upon Tyne, Newcastle upon Tyne, NE1 7RU, UK

Abstract: A combination of hydrometallurgical and electrochemical processes has been developed for the separation and recovery of nickel and cobalt from cylindrical nickelmetal hydride rechargeable batteries. Leaching tests revealed that a 4 mol dm3 hydrochloric acid solution at 95 C was suitable to dissolve all metals from the battery after 3 h dissolution. The rare earths were separated from the leaching solution by solvent extraction with 25% bis(2-ethylhexyl)phosphoric acid (D2EHPA) in kerosene. The nickel and cobalt present in the aqueous phase were subjected to electrowinning. Galvanostatic tests on simulated aqueous solutions investigated the effect of current density, pH, and temperature with regard to current efciency and deposit composition and morphology. The results indicated that achieving an NiCo composition with desirable properties was possible by varying the applied current density. Preferential cobalt deposition was observed at low current densities. Galvanostatic tests using solutions obtained from treatment of batteries revealed that the aqueous chloride phase, obtained from the extraction, was suitable for recovery of nickel and cobalt through simultaneous electrodeposition. Scanning electron micrography and X-ray diffraction analysis gave detailed information of the morphology and the crystallographic orientation of the obtained deposits. 2004 Society of Chemical Industry

Keywords: nickelmetal hydride batteries; rare earth elements; liquid extraction; recycling; electrodeposition; nickel

1 INTRODUCTION The worldwide market of rechargeable batteries is growing at a record pace due to increased consumer demand for portable devices. The nickelmetal hydride (NiMH) battery is one of the most popular rechargeable batteries for use in portable devices, power tools, phones, camcorders and portable computers.1 3 The high energy density and long cycle life of this battery make it a leading technology for the power source for electric vehicles.4 7 Commercial production of NiMH began in 1993 with 100 million cells. By the year 1999 more than 900 million cells, corresponding to 20 000 tonnes of batteries were manufactured world-wide.8,9 In the year 2000, the one billion mark was passed, representing 4% of the world-wide rechargeable battery production, with a market of 720 million. In the year 2010, NiMH batteries are expected to comprise 29% of the total battery market, with a world-wide battery market of 9 billion.10 Since the use of NiMH batteries is large, the expected amount of scrap at the end of their life is

very large, representing an important source of environmental pollution if spent batteries are thrown away. Recycling of batteries is necessary both from an environmental point of view and from the fact that NiMH cells have a valuable metal content of nickel, cobalt, and rare earth elements. Various investigations for recovering valuable metals from NiMH rechargeable batteries have been reported,11 18 most of which use a combination of mechanical and hydrometallurgical processes including dissolution in inorganic acids,11 solvent extraction, and precipitation.14,15 There are many cell types of nickelmetal hydride cells available on the market (ie cylindrical, prismatic, and button cell). Since battery composition and materials differ between manufacturers, the work reported here describes what can be expected from a hydrometallurgical and electrochemical treatment of NiMH batteries. The present study evaluates the recycling of NiMH rechargeable batteries using hydrometallurgical processes with electrowinning for the recovery of nickel and cobalt. An aim of this work is to recover a nickelcobalt alloy, from spent batteries,

Correspondence to: K Scott, School of Chemical Engineering and Advanced Materials, University of Newcastle upon Tyne, Newcastle upon Tyne, NE1 7RU, UK E-mail: k.scott@ncl.ac.uk Contract/grant sponsor: EPSRC/HEFCE; contract/grant number: JIF4NESCEQ (Received 2 February 2004; revised version received 25 March 2004; accepted 15 April 2004) Published online 28 July 2004 927

2004 Society of Chemical Industry. J Chem Technol Biotechnol 02682575/2004/$30.00

N Tzanetakis, K Scott

with good properties, suitable for various technological applications. Electrodeposition of nickelcobalt alloys has been widely studied due to their excellent properties, such as hardness and strength, and various applications in the electronic, mechanical, magnetic, and aerospace industries.19 23 A study of the inuence of deposition parameters was carried out to assess the effect of these parameters both on the electrochemical response and energy consumption, and morphology and composition of the obtained deposit. Optimum operating conditions for alloy electrodeposition were experimentally established.

Saturated Calomel Electrode Working Electrode 10 cm Catholyte Outlet

Cation-Exchange Membrane 2 cm Counter Electrode

Anolyte Outlet

22 cm

Anolyte Inlet Catholyte Inlet Flow Distribution 3 cm

2 EXPERIMENTAL 2.1 Dissolution tests Cylindrical AA-size NiMH batteries, manufactured in Japan (Toshiba) were used in this work. The batteries, which weighed an average of 23 g, were cut in half longitudinally and after removal of the external stainless steel case, were leached in aqueous HC1 solutions. Hydrochloric acid was employed as a leaching agent due to the feasibility of extracting nickel ions from chloride solutions. After dissolution, the leachate and insoluble materials, such as the polypropylene separator and the iron grid, were separated by ltration. 2.2 Solvent extraction tests Solvent extraction tests were conducted on the chloride solutions obtained from leaching. The extractions were performed by mechanical shaking of 500 cm3 conical asks each containing 100 cm3 of the aqueous and organic solutions. A solution of 25% bis-(2-ethylhexyl)phosphoric acid (D2EHPA) in kerosene was chosen as a solvent for the feasibility of selective complexation and extraction of rare earths, at low pH, but not of nickel and cobalt.24 27 The pH of the aqueous and organic solution was adjusted to the desired value by drop-wise addition of concentrated sodium hydroxide solution. After shaking, the two phases were allowed to disengage completely and then separated into rafnate and organic phases. 2.3 Electrodeposition tests Electrochemical experiments were performed in a rectangular acrylic ow cell with removable electrodes. A platinum oxide-based coated titanium anode and a stainless steel cathode, each of 35 cm2 area, were used with a saturated calomel electrode as a reference. The cell was divided into two compartments by a NaonR 117 membrane. Figure 1 shows the design characteristics of the ow cell used for the tests. Prior to each experiment the stainless steel cathode was abraded with silicon carbide paper (P600), rinsed with acetone and then distilled water. Electrochemical experiments were accomplished with a potentiostat/galvanostat (Sycopel Scientic) controlled with
928

Figure 1. Design characteristics of the two-compartment acrylic electrodeposition cell.

Sycopel Scientic electrochemistry software. During the test, cell voltages were continuously recorded with a multimeter. The simulated solutions were prepared from analytical grade metal sulfate salts (Aldrich). Boric acid, concentration of 0.15 g dm3 , was used as a buffer and 1 mol dm3 Na2 SO4 as an anolyte. Each experiment was carried out using a fresh solution which was deoxygenated with nitrogen for 30 min prior to electrolysis. The solutions were circulated from the two 1 dm3 glass reservoirs to the cell by a peristaltic pump at a ow rate of 0.12 dm3 min1 . Throughout the tests the solutions in the reservoirs were stirred with magnetic stirrers. Samples were taken from the catholyte every hour and the amount of deposited metal was calculated by difference from the initial value. Small amounts of metal ions migrated into the anolyte after long term electrolysis and these amounts were allowed for in determining the current efciencies and deposit compositions. Depending on the charge passed, up to approximately 300 and 35 ppm of nickel and cobalt respectively, diffused into the anolyte. The electrochemical equivalent of the deposit, ie the charge used for total metal deposition (practical charge), was determined from its composition. The value was then compared with the electric charge passed through the system (theoretical charge) to determine the current efciency. The metal ion content of various elements was determined by atomic absorption analysis, using a Unicam 929 AA spectrometer. The morphology and the crystal structure of the deposits were examined using a Hitachi S-2400 scanning electron microscope and a Philips X Pert Pro diffractometer respectively.

3 RESULTS AND DISCUSSIONS 3.1 Leaching tests The Parameters investigated during the dissolution tests were acid concentration, dissolution time, and
J Chem Technol Biotechnol 79:927934 (online: 2004)

Recycling of nickelmetal hydride batteries. II

temperature. Detailed description of this work is given in another communication.28 It is important to mention that dissolution of all metals was found to depend signicantly on leaching temperature. Nearly complete dissolution occurred for cobalt, nickel, rare earths, zinc and manganese after 3 h using 3 mol dm3 HC1 at a temperature of 95 C. Low iron dissolution of the electrode grids was desirable for minimal contamination of Ni and Co during recovery. However some extraction of Fe was unavoidable due to the need for high nickel recovery. After ltration of the leaching solution the resulting green chloride solution had a pH close to zero and a composition of (in g dm3 ): 55.7 Ni, 5.7 Co, 2.2 Mn, 0.99 Zn, 2.8 Fe, 1.1 Al, 6.4 La, 6.2 Ce and 2.5 Nd. 3.2 Solvent extraction experiments Solvent extraction using 25% D2EHPA (0.75 mol dm3 ) as an extractant, demonstrated the feasibility of recovering the rare earths, while leaving Ni, Co and Mn in the aqueous phase to later undergo electrowinning. Extraction tests over a wide range of aqueous pH (03) indicated that, at a pH close to 2.5, all the rare earths and most of the iron could be extracted into the organic phase and hence separated from nickel and cobalt.28 Extraction of manganese, aluminium and zinc was relatively low over the whole range of acidity possibly as a result of preferential extraction of rare earths and iron.14 3.3 Electrochemical deposition This is a preliminary study to illustrate the feasibility of using electrodeposition for the recovery of nickel and cobalt from the leach solutions from spent NiMH batteries and initially investigation of the voltammetric characteristics of the systems was carried out. Studies using simulated solutions were performed to evaluate the effect of current density, pH, and temperature on the composition and morphology of the electrodeposits. The results are then compared with those obtained using solutions derived from the hydrometallurgical treatment of batteries. 3.3.1 Voltammetric studies Figure 2 shows typical linear sweep voltammograms (LSVs) obtained for the nickel, cobalt, and simulated battery solutions containing Ni, Co and Mn, and for the solution obtained from hydrometallurgical treatment of the batteries. The curves for the simulated solutions revealed clear nucleation processes followed by hydrogen evolution at more negative potentials. The currentpotential curves showed only a small difference in the nucleation potential for nickel and cobalt. Electrochemical reduction of nickel and cobalt began at a potential of approximately 850 mV vs SCE, while hydrogen reduction occurred at potentials more negative than 1100 mV. The limiting current densities for nickel, cobalt, and simulated battery
J Chem Technol Biotechnol 79:927934 (online: 2004)

-30 -25 -20 J/mA cm-2 -15 -10 -5 0 -500 -700 -900 -1100 E/mV vs. SCE -1300 -1500 Nickel Cobalt Simulated solution Battery solution

Figure 2. Linear sweep voltammograms of 0.09 mol dm3 NiCl2 + 1 mol dm3 NaCl, 0.009 mol dm3 CoCl2 + 1 mol dm3 NaCl, simulated battery solution containing 0.09 mol dm3 NiCl2 , 0.009 mol dm3 CoCl2 , 0.004 mol dm3 MnCl2 + 1 mol dm3 NaCl and actual battery solution. All solutions contained 0.25 mol dm3 H3 BO3 , pH 3, and scan rate 20 mV s1 .

solutions were approximately 13, 1.5, and 17 mA cm2 respectively. For the potential range of approximately, 970 to 1050 mV, the different systems operated under mass transport control. The LSV of the simulated battery solution shifts to a potential more positive than that observed in the pure nickel, which indicates that the formation of a nickelcobalt alloy from the simulated solutions was thermodynamically favourable. The voltammogram for the solution, from leaching of a NiMH battery revealed that deposition in this case occurred less readily than from simple chloride solutions and without a clear nucleation overpotential. The average difference in nucleation potential between the simulated and the battery solution was around 200 mV. The low limiting current and the shift in the reduction potential towards more negative values in the case of the battery solution was due to the complexity of this solution and the presence of impurities from the previous leaching and extraction steps. Figure 3 compares cyclic voltammograms from a solution of only nickel ions, and from simulated, and actual battery solutions. In all cases the voltammograms exhibited a single reduction wave and suggested the presence of a nucleation process which was less pronounced in the case of the actual battery solution. The tests with the simulated solution showed a greater deposition rate, ie larger current densities, than that of the battery solution. The voltammetric response associated with the deposition from the battery solution showed a lower current intensity and a shift of the hydrogen evolution reaction, towards more negative potentials. It is worth re-stating at this point the impurities that might be present in the case of the actual battery solution due to the previous hydrometallurgical treatment. The solution from the solvent extraction cycle contained, in addition to nickel, cobalt, and
929

N Tzanetakis, K Scott
10 5 0 J/mA cm-2 -5 -10 -15 -20 -25 -1400 -1200 -1000 -800 -600 -400 E/mV vs. SCE Nickel Simulated solution Battery solution -200 0 200

Figure 3. Cyclic voltammograms of 0.09 mol dm3 NiCl2 + 1 mol dm3 NaCl, simulated battery solution containing 0.09 mol dm3 NiCl2 , 0.009 mol dm3 CoCl2 , 0.004 mol dm3 MnCl2 + 1 mol dm3 NaCl and actual battery solution. All solutions contained 0.25 mol dm3 H3 BO3 , pH 3. The potential was rst scanned cathodically from the rest potential with a scan rate of 20 mV s1 . Arrows indicate the potential scan direction.

manganese, small quantities of iron (0.36 g dm3 ), aluminium, zinc (0.66 g dm3 ), and possibly traces of rare earth elements (0.10.2 g dm3 ) and the organic solvent. Due to these impurities, a much higher overpotential was necessary to begin the nucleation process in the actual battery solution. Despite the fact that both the simulated and the actual battery solutions contained nickel and cobalt as the major constituents, the presence in the latter of the above metals was responsible for the different electrochemical behaviour. The anodic processes revealed a clear oxidation wave in the case of the nickel ion and the simulated solutions, and a smaller one for the battery solution. It can be observed that the simulated solution showed similar characteristics to the nickel curve as nickel was the major component of that solution. In the case of nickel ion and the simulated solutions, the voltammetric charge of the cathodic peak was approximately equal to that of its corresponding anodic peak, which indicated good electrochemical reversibility for these systems. Overall from the voltammetric study it is expected that, in the case of the solution from hydrometallurgical treatment of the NiMH batteries, the deposition process will be inhibited. This inhibition was caused by substrates, other than the metal ion of interest, present either at the surface of the electrode or in the diffusion layer, which hindered the cathodic process and partially covered the surface of the cathode. They also affected the overpotential and the texture of the surface. 3.3.2 Simulated solutions Galvanostatic electrodeposition tests were conducted with solutions containing nickel, cobalt, and manganese ions, the predominant elements expected in
930

the aqueous phase after extraction. The composition of the solution (g dm3 ): 5.57 Ni, 0.57 Co, and 0. 22 Mn, was similar to that obtained after diluting the leaching solution by a factor of 10. Electrowinning experiments from dilute solutions should generally be operated at the maximum useful current density at which efcient deposition takes place. Figure 4 shows the typical effect of current density on current efciency and composition of nickel and cobalt in the deposit. For these experiments the charge passed (3250 C) was identical. The thickness of the deposit obtained was approximately 1.7 mm. The data conrmed co-deposition of nickel and cobalt, while in all cases the deposit contained no more than 2% of manganese (Table 1). A greater proportion of cobalt was deposited at low current densities although the amount was approximately one tenth that of nickel. Higher cobalt content potentially results in high strength and hardness of the obtained alloy.19 Overpotentials of 8201100 mV were observed at the beginning of the deposition process at current densities of 10100 A m2 (Table 1). The overpotentials decreased by approximately 50100 mV during the runs. It is also apparent that the cobalt content in the deposit decreased when the applied current density was increased. The current efciencies decreased at higher current densities due to metal ion depletion, which resulted
120 100 Current efficiency/% 80 60 60 40 40 20 0 0 10 20 30 40 50 60 70 80 Current density/A m-2
Figure 4. Effect of current density on alloy current efciency and composition. Experiments were carried out in baths containing 0.09 mol dm3 NiCl2 , 0.009 mol dm3 CoCl2 , 0.004 mol dm3 MnCl2 , 1 mol dm3 NaCl and 0.25 mol dm3 H3 BO3 at 22 C and at pH 3. Charge passed, 3250 C.

Ni

100 80

Co

20 0 90 100 110

Table 1. Effect of current density on the composition of the deposit, the cell voltage and the energy consumption for tests using simulated battery solutions at pH 3

Initial Cell Energy Current Ni Co Mn Overpotential Voltage Consumption (A m2 ) (%) (%) (%) (mV) (V) (kWh mol1 ) 10 25 50 75 100 70 83 87 87 88 30 15 12 11 10 0 2 1 2 2 820 850 900 1050 1100 2.1 2.6 2.9 3.15 3.44 0.109 0.133 0.15 0.248 0.296

J Chem Technol Biotechnol 79:927934 (online: 2004)

Amount in deposit/%

Recycling of nickelmetal hydride batteries. II

in operating conditions above the limiting current and greater hydrogen evolution. When experiments were performed at current densities higher than 100 A m2 the deposits included green nickel hydroxides which would detract from the properties of the deposit. The nickel hydroxide was possibly formed due to operation above the limiting current at which the local pH at the electrode increased, due to localised formation of OH ions. Table 1 shows the cell voltages and energy consumptions obtained at the termination of the electrodeposition experiments. Higher current densities resulted in signicantly greater cell voltages. During all electrodeposition tests the cell voltage decreased slightly with time and hence so did the energy consumption. Experiments carried out to optimise the operating pH indicated that at a pH between 2 and 4 higher metal removal at greater current efciencies and better quality deposits were obtained (Fig 5). The composition of the deposit was similar regardless of the solution pH. However at low pH, reduction of hydrogen ions was high which led to low current efciencies. Since cogeneration of hydrogen with nickelcobalt electrodeposition would tend to increase pH locally at the electrode surface, formation and inclusion of metal hydroxides into the deposit is likely. Thus maintaining the pH in the appropriate range during these experiments was necessary. Precipitation of metals and occlusion of hydroxides in the growing deposit was observed during the runs at pH values above 4.5. The formation of hydroxides inhibited the deposition process and induced the formation of cracked non-homogeneous deposits. During all runs an increase in the catholyte pH was observed, which was adjusted to the appropriate value by addition of hydrochloric acid solution. Figure 6 shows scanning electron microscopic (SEM) images of the surface morphology of the nal deposits obtained under various electrodeposition conditions. A constant alloy composition throughout

the thickness of the deposit was formed. In all cases deposits adhered well to the stainless steel substrate. It is clear that the morphology of the structures varied signicantly with the experimental conditions used. The grain size became gradually rened on decreasing the current density and by ensuring that the nucleation took place slowly. The higher cobalt content at low current densities increased the crystalline nature of the alloy and gave a compact, homogeneous, dense, and ne-grained structure. At high current densities, the alloy structure was characterised by large, randomly distributed grains without visible porosity. Black inclusions were distinguished in the micrographs while the colour of the alloy changed from silver to black. The deposit obtained at higher temperatures was globular, with variable grain sizes, without a uniform distribution over the surface. Deposition tests performed at 40 and 60 C did not give a signicant improvement in current efciency. The structure of the alloy and the size distribution of the crystallites became more homogeneous as the pH increased and a metallic deposit was obtained. The deposit at high pH exhibited a silver foil appearance. Low pH (<1), on the other hand, provoked inhibition of the growth process while deposition of hydrogenrich alloy possibly occurred. In this case a brittle deposit, that partially lifted from the electrode, probably due to hydrogen evolution, was obtained. 3.3.3 Battery solutions Batch galvanostatic electrodeposition experiments were performed with solutions obtained after the dissolution and solvent extraction processing of NiMH batteries, for 10 h or until a specic electric charge had passed. Table 2 gives the deposit compositions obtained for tests performed at three different current densities. The deposits had a dendritic appearance, especially at the higher current density, possibly due to the presence of impurities in the bath. Figure 7 shows the variations in nickel and cobalt catholyte concentrations versus charge passed for two current densities. The rate of deposition for both metals increased at higher current densities. After 10 h electrolysis, approximately 82% and 68% of nickel and cobalt respectively were deposited at a current density of 50 A m2 . As the concentration of metal ions fell during the tests, the current efciency and consequently the metal removal rate also fell. For instance for the run at 50 A m2 , the current efciency was initially close to 100% and fell to 65% at the end of the experiment. Under these conditions, the competing electrochemical reaction of hydrogen evolution became more dominant, thus using a higher percentage of the applied current. In general relatively high current densities are benecial, from an industrial point of view, to achieve fast metal removal. However tests at higher current densities of 200500 A m2 resulted in poor metal recovery and low current efciencies (1020%)
931

120 100 Current efficiency/% 80 60 40 20 0 0 1 2

Ni

100 Amount in deposit/% 80 60 40 Co 20 0 3 4 5

pH
Figure 5. Effect of pH on alloy current efciency and composition. Experiments were carried out in baths containing 0.09 mol dm3 NiCl2 , 0.009 mol dm3 CoCl2 , 0.004 mol dm3 MnCl2 , 1 mol dm3 NaCl and 0.25 mol dm3 H3 BO3 at 22 C and at a current density = 50 A m2 . Charge passed, 6120 C.

J Chem Technol Biotechnol 79:927934 (online: 2004)

N Tzanetakis, K Scott

(a)

(b)

(c)

(d)

Figure 6. Scanning electron micrographs of deposits obtained under different experimental conditions. (a) CD = 10 A m2 , pH = 3, T = 25 C, (b) CD = 100 A m2 , pH = 3, T = 25 C, (c) CD = 50 A m2 , pH = 3, T = 40 C, (d) CD = 50 A m2 , pH = 1.5, T = 25 C. Table 2. Effect of current density on the composition of the deposit and the current efciency for tests using solutions obtained from treatment of batteries at pH 3 for all runs

6000 5000 Ni concentration/ppm 4000 3000 2000 1000 0 0

1250 1050 850 650 450 250 500 1000 1500 2000 2500 3000 3500 4000 Charge passed/Coulombs
, Ni, 20 A m-2 , Ni, 50 A m-2 , Co, 20 A m-2 , Co, 50 A m-2

Current density (A m2 ) 10 50 100

(%)

Ni (%)

Co (%)

Mn (%)

Al (%)

Fe (%)

Zn (%)

Rare earths (%) 0 0.8 1.2

100 83 14.1 0 82 76.2 12 7 39 82 11.2 3

1 1.2 0.7 4 0 0 0.7 0.5 1.4

= current efciency.

because of re-dissolution of metals, due to their poor adhesion to the substrate caused by excessive hydrogen evolution. At such higher current densities the surface of the electrode was covered with metal oxides or hydroxides which appeared as black or green deposits. However runs at low current densities were more successful since deposits adhered well to the electrode and contained lower amounts of impurities (Table 2). In this case though, a high electrode area is necessary for efcient metal recovery. Hence a high surface area (three dimensional) electrode is recommended to maximise deposition rate and the space time yield. Figure 8 shows the X-ray diffraction patterns of deposits obtained at three current densities.
932

Figure 7. Variation in nickel and cobalt concentrations in the catholyte with charge passed. Experiments were performed at two current densities at pH 3.

The crystal structure of the deposits was affected by variations in current densities as were the morphologies shown by the SEM photographs (see Fig 6). The deposits obtained from battery solutions showed amorphous structures with irregular morphology and thus the SEMs are not shown. The deposit obtained at 10 A m2 had a face-centred cubic (fcc) structure with sharp reections from nickel
J Chem Technol Biotechnol 79:927934 (online: 2004)

Co concentration/ppm

Recycling of nickelmetal hydride batteries. II


(111) (200) Intensity (Arbitrary units) (220) (311) 10 Am-2

50 A m-2

100 Am-2 0 20 40 60 2-Theta (Degrees)


Figure 8. X-ray diffraction patterns of deposits obtained from solutions obtained from battery treatment, at three current densities, 10, 50 and 100 A m2 . All runs performed at pH 3.

80

100

120

and cobalt crystalline grains, with very strong (111) growth orientation (2 = 44 ) with signicant but weaker (200), (220) and (311) reections at 2 values of 52 , 76 and 93 respectively. This crystallographic orientation is similar to that observed from pure state nickel crystallites.29 Moreover the sharp reections throughout the spectrum indicated the presence of large-size crystallites. Single peaks were observed for nickel and cobalt due to their indistinguishable X-ray scattering powers. Similarly the spectrum obtained at a current density of 50 A m2 exhibited the same (111), (200), (220) and (311) oriented planes although at a lower intensity when compared with those at 10 A m2 . The two spectrams clearly proved the crystalline structure of the deposit at low current density. At higher current densities an increase in the fcc lattice parameter and an absence of clear diffraction peaks was observed, which indicated the presence of inclusions in the deposit. The particle size of the deposit obtained, especially at higher current densities, was irregular due to the formation of dendrites. In addition to some nickel crystallites formed during electrodeposition, nickel oxideshydroxides were also embedded in the deposit. The absence of diffraction peaks is also partly attributed to the presence of other metal oxides in the deposit. This speculation is supported by the amorphous nature of these deposits. The presence of metal impurities and the adsorption of oxides into the electrode surface inhibited, to a great extent, cobalt and nickel reduction by lowering the limiting current. Future work will focus on purication of the battery solution which will enable efcient operations at higher current densities. Iron and aluminium removal using precipitation steps prior to the aforementioned deposition tests and anodic deposition of manganese as an oxide will be investigated.

spent rechargeable nickelmetal hydride batteries. The work revealed that appropriate leaching can be performed with 4 mol dm3 HC1 at a temperature of 95 C and a leaching time of 3 h or less. Rare earth elements were separated from nickel and cobalt by solvent extraction using D2EHPA as an extractant at a pH close to 2.5. The feasibility of recovering nickel and cobalt as an alloy by electrowinning from the rafnate was demonstrated using simulated and real battery solutions. Galvanostatic tests on simulated solutions showed that composition and surface morphology of the alloy is largely inuenced by modifying the deposition parameters (mainly pH and current density). The crystalline nature of the alloy increased as the cobalt content of the deposit increased. Electrodeposition tests using solutions obtained after hydrometallurgical treatment of NiMH batteries showed that recovery of nickel and cobalt was feasible. Operations at low current densities gave a deposit with a crystalline structure at high current efciencies, while higher current densities gave amorphous materials with signicant reduction in current efciencies. Overall the data reported can be used for the development of a process which will allow efcient electrochemical recovery of Ni and Co from spent NiMH batteries.

ACKNOWLEDGEMENTS Cumberland Electrochem (Bootle) are thanked for providing the electrode materials for the project. The work was performed in research facilities provided through an EPSRC/HEFCE Joint Infrastructure Fund award, no JIF4NESCEQ. REFERENCES
1 Bulletin of British Battery Manufacturers Association, Dec (1996). 2 Wiaux JP, in 49th ISE meeting, Sept, Kokura, Japan, abstract K-11-18-02, p 817 (1998). 3 Ruetschi P, Meli F and Desilvestro J, Nickelmetal hydride batteries. The preferred batteries of the future. J Power Sources 57:8591 (1995). 4 Soria ML, Chacon J, Hernandez JC, Moreno D and Ojeda A, Nickel metal hydride batteries for high power applications. J Power Sources 96:6875 (2001). 5 Zhan F, Jiang LJ, Wu BR, Xia ZH, Wei XY and Qin GR, Characteristics of NiMH power batteries and its applications to electric vehicles. J Alloys and Compounds 293295:804808 (1999). 6 Hoshino H, Uchida H, Kimura H, Takamoto K, Hiraoka K and Matsumae Y, Preparation of a nickelmetal hydride (NiMH) rechargeable battery and its application to a solar vehicle. Int J Hydrogen Energy 26:873877 (2001). 7 Gifford P, Adams J, Corrigan D and Venkatesan S, Development of advanced nickel/metal hydride batteries for electric and hybrid vehicles. J Power Sources 80:157163 (1999). 8 Ovshinsky SR, Dhar SK, Fetcenko MA, Young K, Reichman B, Fierro C, Koch J, Martin F, Mays W, Sommers B, Ouchi T, Zallen A and Young R, presented at 17th International Seminar & Exhibit on Primary and Secondary Batteries, Ft Lauderdale, Florida, March 69 (2000). 933

4 CONCLUSIONS A combination of hydrometallurgical and electrochemical processes can potentially provide an efcient process for recycling valuable metal materials from
J Chem Technol Biotechnol 79:927934 (online: 2004)

N Tzanetakis, K Scott
9 Noreus D, Substitution of rechargeable NiCd batteries. University of Stockholm, August (2000). available at: http://europa.eu.int/comm/environment/waste/nicd.pdf. 10 Stempel RC, in Third Annual Merrill Lynch Global Energy Technology Conf New York, May (2002). 11 Lyman JW and Palmer GR, Recycle of nickel hydride battery scrap. J Metals 3235 (1993). 12 Lyman JW and Palmer GR, in Third International Symposium on Recycling of Metals and Engineered Materials, ed by Queneau PB and Peterson RD. The Minerals, Metals and Materials Society, Warrendale, PA, pp 131144 (1995). 13 Yoshida T, Ono H and Shirai R, Recycling of used NiMH rechargeable batteries in Third International Symposium on Recycling of Metals and Engineered Materials, ed by Queneau PB and Peterson RD. The Minerals, Metals and Materials Society, Warrendale, PA, pp 145152 (1995). 14 Zhang P, Yokoyama T, Itabashi O, Wakui Y, Suzuki TM and Inoue K, Hydrometallurgical process for recovery of metal values from spent nickelmetal hydride secondary batteries. Hydrometallurgy 50:6175 (1998). 15 Zhang P, Yokoyama T, Itabashi O, Wakui Y, Suzuki TM and Inoue K, Recovery of metal values from spent nickelmetal hydride rechargeable batteries. J Power Sources 77:116122 (1999). 16 Tenorio JAS and Espinosa DCR, Recovery of Ni-based alloys from spent NiMH batteries. J Power Sources 108:7073 (2002). 17 Wang R, Yan J, Zhou Z, Gao X, Song D and Zhou Z, Regeneration of hydrogen storage alloy in spent nickelmetal hydride batteries. J Alloys and Compounds 336:237241 (2002). 18 Pietrelli L, Bellomo B, Fontana D and Montereali MR, Rare earths recovery from NiMH spent batteries. Hydrometallurgy 66:135139 (2002). 19 Gomez E and Valles E, Electrodeposition of Co + Ni alloys on modied silicon substrates. J Appl Electrochem 29:805812 (1999). 20 Golodnitsky D, Rosenberg Y and Ulus A, The role of anion additives in the electrodeposition of nickelcobalt alloys from sulfamate electrolyte. Electrochim Acta 47:27072714 (2002). 21 Lupi C and Pilone D, Electrodeposition of nickelcobalt alloys: the effect of process parameters on energy consumption. Minerals Eng 14:14031410 (2001). 22 Burzynska L and Rudnik E, The inuence of electrolysis parameters on the composition and morphology of CoNi alloys. Hydrometallurgy 54:133149 (2000). 23 Correia AN and Machado SAS, Electrodeposition and characterisation of thin layers of NiCo alloys obtained from dilute chloride baths. Electrochim Acta 45:17331740 (2000). 24 Peppard DF, Mason GW, Driscoll WJ and Sironen RJ, Acidic esters of orthophosphoric acid as selective extractants for metallic cations-tracer studies. J Inorg Nucl Chem 7:276285 (1958). 25 Pierce TB and Peck PF, The extraction of the lanthanide elements from perchloric acid by di-(2-ethylhexyl)hydrogen phosphate. Analyst 88:217221 (1963). 26 Grimm R and Kolarik Z, Extraction of Cu(II), Co(II), Ni(II), Zn(II) and Cd(II) by di(2-ethylhexyl)phosphoric acid. J Inorg Nucl Chem 36:189192 (1974). 27 Sarangi K, Reddy BR and Das RP, Extraction studies of cobalt(II) and nickel(II) from chloride solutions using NaCyanex 272. Separation of Co(II)/Ni(II) by the sodium salts of D2EHPA, PC 88A and Cyanex 272 and their mixtures. Hydrometallurgy 52:253265 (1999). 28 Tzanetakis N and Scott K, Recycling of nickelmetal hydride batteries. I: Dissolution and solvent extraction of metals. J Chem Technol Biotechnol 79:919926 (2004). 29 Orhan G, Arslan C, Bombach H and Stelter M, Nickel recovery from the rinse waters of plating baths. Hydrometallurgy 65:18 (2002).

934

J Chem Technol Biotechnol 79:927934 (online: 2004)

Você também pode gostar