Exploring a green route for recycling spent lithium-ion batteries: Revealing and solving deep screening problem

Exploring a green route for recycling spent lithium-ion batteries: Revealing and solving deep screening problem

Journal Pre-proof Exploring a green route for recycling spent lithium-ion batteries: Revealing and solving deep screening problem Jiadong Yu, Quanyin ...

3MB Sizes 2 Downloads 161 Views

Journal Pre-proof Exploring a green route for recycling spent lithium-ion batteries: Revealing and solving deep screening problem Jiadong Yu, Quanyin Tan, Jinhui Li PII:

S0959-6526(20)30316-4

DOI:

https://doi.org/10.1016/j.jclepro.2020.120269

Reference:

JCLP 120269

To appear in:

Journal of Cleaner Production

Received Date: 11 November 2019 Revised Date:

31 December 2019

Accepted Date: 25 January 2020

Please cite this article as: Yu J, Tan Q, Li J, Exploring a green route for recycling spent lithium-ion batteries: Revealing and solving deep screening problem, Journal of Cleaner Production (2020), doi: https://doi.org/10.1016/j.jclepro.2020.120269. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier Ltd.

Credit author statement Jiadong Yu: Conceptualization, Methodology, Visualization, Writing Original draft preparation, revising; Quanyin Tan: Formal analysis, co- supervision, co-project administration; Jinhui

Li:

Conceptualization,

administration, funding acquisition.

Resources,

supervision,

project

Graphical abstract

1

Exploring a green route for recycling spent Lithium-ion batteries:

2

revealing and solving deep screening problem

3

Jiadong Yu a, Quanyin Tan *,a, Jinhui Li *,a

4

a

State Key Joint Laboratory of Environment Simulation and Pollution Control, School of

5

Environment, Tsinghua University, Beijing 100084, China

6 7

Abstract:

8

The conventional recycling methods of spent lithium-ion batteries (LIBs) regard Al/Cu foils as

9

impurities, which are usually removed by deep screening or alkali leaching. Deep screening refers to

10

the process of over-screening the electrode materials by passing the crushed materials of spent LIBs

11

through a screen with a pore size of only 0.075 mm. However, the serious hole-blocking

12

phenomenon of deep screening and the environmental hazards of alkali leaching restrict their

13

practical application. Herein, a combination of mild screening, reduction leaching and selective

14

purification is proposed to achieve green and sustainable recycling of spent LIBs. Specifically, spent

15

LIBs are subjected to discharge, crushing and mild screening to obtain fine-mixture materials under

16

4-mm sieving, leaving large pieces of Cu/Al foils on the screen mesh, which can be separated as

17

crude products by color sorting. Then, dilute H2SO4 is employed to dissolve the valuable elements in

18

underflow fractions, where residual Cu/Al foils serve as reductants to assist leaching. Eventually,

19

through multi-step precipitates and hydrothermal treatment, a series of promising products (CuO,

20

NaAlCO3(OH)2, precursor and Li2CO3) are obtained, at high quality. This route provides a clever

21

strategy for fine management and cascade recovery of the valuable components from spent LIBs.

22 1

23

Key words: Spent Lithium-ion battery, Deep screening, Reductive leaching, Ternary precursor,

24

NaAlCO3(OH)2.

25 26

1. Information

27

With the representative codes of high energy density, high battery voltage, long life span and no

28

memory effect, lithium-ion batteries (LIBs) have dominated the global market for smart devices and

29

electric vehicles (Golmohammadzadeh, R., et al. 2018). According to the latest research report, the

30

global market value of LIBs in 2017 was US$ 25 billion, and it is predicted to reach US$ 47 billion

31

by 2023, exhibiting a CAGR of around 11% during 2018-2023(Wood, L., 2018). The flourishing

32

development of LIBs stems from the robust demand for safer and stronger rechargeable batteries

33

(Hannan, M.A., et al. 2017), but this popularity will also lead to a surge in the amount of LIBs

34

entering the waste stream, in the near future. It is estimated that in the European Union alone, there

35

will be 13,828 tons of spent LIBs in need of efficient recycling by 2020 (Träger, T., et al. 2015).

36

Spent LIBs contain many heavy metals, such as Mn, Ni, Co and Cu, and poor recycling treatments

37

for these wastes will result in irreparable damage to the environment (Yu, J., et al. 2017). Conversely,

38

responsible recycling of end-of-life LIBs can not only facilitate the development of environmental

39

protection and resource sustainability, but also achieve considerable economic benefits.

40

Typical spent LIBs usually include electrode materials, Cu/Al foils, shells, separators and

41

electrolytes (Yu, J., et al. 2018; Wang, M., et al. 2018). The recycling methods for spent LIBs focus

42

on the valuable cathode materials, and can be primarily divided into hydrometallurgy (Nayaka, G.P.,

43

et al. 2016; Vieceli, N., et al. 2018) and pyrometallurgy (Li, J., et al. 2016; Xiao, J., et al. 2017). 2

44

High-temperature smelting is a simple and efficient approach, with the advantages of less

45

consumption of chemical reagents and short recycling route (Dos Santos, C.S., et al. 2019), but it

46

could produce some toxic sludge containing lithium, aluminum and other heavy metals (Li, L., et al.

47

2017; Roshanfar, M., et al. 2019). On the other hand, based on the special functions in precise

48

separation, effective recovery and selective preparation of high-value products, hydrometallurgy has

49

attracted more attention (Joulié, M., et al. 2014). Although the two companies, Sungeel Hi-Metal Co.

50

(South Korea) and Umicore Co. (Belgium), have claimed to master the key technology for

51

commercial application of spent LIBs, it is difficult for developing countries to learn and transform

52

locally due to technical barriers and equipment manufacturing defects. Therefore, the urgent vision is

53

to explore facile technologies that can be replicated and promoted worldwide. Furthermore, cathode

54

materials are often composed of trivalent cobalt and tetravalent manganese, and these high-valent

55

transition metal oxides are poorly soluble in hydrometallurgy (Sattar, R., et al. 2019). In order to

56

alleviate this problem, an additional reductant (Chen, X., et al. 2015; Hu, J., et al. 2017) or a

57

reducing organic acid (Gao, W., et al. 2017; Li, L., et al. 2017) should be employed, to enhance the

58

leaching effect of cathode materials. Nevertheless, the purchase, transportation and storage of

59

reducing agents will increase operating costs (Ma, X., et al. 2018). And since most acid leaching

60

methods regard Al as an impurity element, deep screening (the passing of powders through a screen

61

with the pore size less than 75 microns (Zhang, X., et al. 2018; Zheng, X., et al. 2018)) was studied

62

to forcibly remove the residual Al foils from the cathode electrode powders after efficient crushing,

63

and even an alkali leaching step (Gratz, E., et al. 2014; Ferreira, D.A., et al. 2009) could be added to

64

fully remove Al element. However, deep screening is almost impossible to realize in industrial 3

65

application, and the sudden addition of the alkali leaching will significantly increase the acid

66

consumption in subsequent processes. In addition, some scholars used deep eutectic solvent (Wang,

67

M., et al. 2019) or toxic NMP solvent (Contestabile, M., et al. 2001) to separate the cathode powder

68

from the Al foils completely by removing the organic binder, so as to avoid the Al pollution. But this

69

may increase the complexity and cost of the recycling process. Therefore, reusing the Al foils in

70

spent LIBs as a reductant to assist in the leaching of cathode materials is a promising strategy.

71

Joulié (Joulié, M., et al. 2017) first put forward the idea that Cu and Al could be utilized as

72

reducing agents to dissolve NCM cathode materials by acid leaching, and gave a simple dynamic

73

explanation. Peng (Peng, C., et al. 2019) also found that Cu, Al and Fe in spent LIBs could

74

significantly improve the leaching effect of Li and Co up to almost 100%. Unfortunately, they did

75

not consider the effective amount of Cu and Al foils, nor did they specify the corresponding

76

purification methods. It is precisely these two key factors that limit the practical application of Cu/Al

77

foils as reductants in waste battery recovery.

78

In this study, we are committed to consuming the least amount of Cu/Al foils as reductants to

79

assist in leaching and recovering the most valuable metal elements. This meticulous management

80

strategy is applied to cleverly solve the deep-screening bottleneck problem in the industrial

81

application of LIB recycling. Compared with traditional methods, this technology will bring the

82

following benefits: (1) This technology refuses to use caustic alkali or toxic solvents in the recycling

83

process; (2) It is unnecessary to add additional reducing agents, such as H2O2 and Na2S2O5, which

84

reduces the consumption of chemical reagents and greenhouse gas emissions; (3) This technology

85

realizes the recovery of all components of valuable materials in spent LIBs. The general framework 4

86

of this research is as follows: the feasibility of Cu/Al powder as reductants was firstly verified in

87

simulation experiments. Secondly, thermodynamic analysis gave the reduction mechanism and

88

determined the theoretical demand of Cu and Al foils. In the third section, fine management of Cu/Al

89

foils was conducted under the actual accumulative effect of mild screening. Then, the systematic

90

exploratory experiment of actual screening products revealed the optimum leaching conditions. Next,

91

through progressive precipitation and hydrothermal process, some high value products were prepared,

92

and their crystallographic and microscopic morphologies were discussed in detail. Finally, a green

93

and fully mechanized process was distilled, removing an important obstacle to industrial recycling of

94

spent LIBs.

95 96

2. Experimental materials and methods

97

2.1 Materials and reagents

98

Spent ternary lithium-ion batteries from electric vehicle battery packs were donated by

99

TES-AMM, an environmental protection company in Beijing. As shown in Figure 1, the diffraction

100

analysis results of electrode materials from discarded LIBs gave that the cathode material was a

101

high-nickel ternary material with LiNiO2 crystal phase. Furthermore, the element analysis showed

102

that the electrode materials were composed of 5.19 wt.% lithium, 30.77 wt.% nickel, 12.23 wt.%

103

cobalt and 17.21 wt.% manganese and 36.4 wt.% graphite. Therefore, the waste cathode material in

104

this experiment was Li(Ni0.5Co0.2Mn0.3)O2, namely, 523 ternary cathode materials. The chemical

105

reagents used in the experiments, including Cu, Al, H2SO4, H2O2, Na2CO3, NaOH, and

106

diaminomethanal, were of analytical grade. Deionized water was used as the experimental water. 5

107 108

2.2 Mild screening experiments

109

In order to avoid the fire caused by the short circuit in the crushing process, three (600g)

110

lithium-ion batteries were first immersed in 5% sodium chloride solution for 48 hours to discharge

111

completely. Besides, this could also allow the residual Li ions on the anode surface to return to the

112

cathode materials (Liu, C., et al. 2016), improving the recovery rate of Li element in the whole

113

recovery process. All batteries were thoroughly crushed in the fume hood by the hand-held universal

114

crusher. The crushed products were screened in turn by mild screening with sieve hole sizes of 4 mm,

115

2 mm, 1.45 mm, 0.9 mm and 0.5 mm. Specifically, mild screening means that only 5 minutes of

116

screen vibration is needed to make materials permeable, without water pressure or other external

117

force to assist in the sieving.

118 119

2.3 Leaching and purification experiments

120

The leaching experiments were completed in a 500-ml three-necked flask with a reflux

121

condenser. In addition, it was necessary to use a heating bath equipped with magnetic stirring to

122

assist in material agitation and temperature control. During the experiments, a certain amount of raw

123

materials was added to 100 ml dilute sulfuric acid. The leaching experiments in this study were

124

divided into simulation experiments and conditional optimization experiment. In order to illustrate

125

the feasibility of using Cu and Al as reducing agents, simulation experiments were first performed.

126

Different combinations of leaching agents (H2SO4 alone, H2SO4 + Cu/Al, H2SO4 + 6 vol % H2O2)

127

with analytically pure grade were employed to leach 523 high-nickel cathode materials. Because the 6

128

mass ratio of NCM / Cu and NCM / Al in spent LIBs were 1:1.2 and 1:0.7 respectively (Doberdò et

129

al., 2014; Loeffler et al., 2014), the proportion of NCM / Cu / Al was simplified to 2:1:1 in the

130

simulation experiment. The operating parameters of simulation experiments were 2.5 mol/L H2SO4,

131

200 mL/g, 60℃ and 6h of reaction time. For conditional optimization experiment, the raw materials

132

were the mixture obtained under 4 mm mild screening. Different from the simulation experiments,

133

the raw materials and reducing agents were the electrode materials and Cu/Al foils inherent in the

134

spent LIBs. In this section, the temperature was controlled at 0℃, 30℃, 45℃, 60℃ and 75℃, the

135

liquid-solid ratio was adjusted within the range of 20 mL/g, 50 mL/g, 100 mL/g, 200 mL/g and 400

136

mL/g, and the sulfuric acid concentration range was 0.5- 4.0 mol / L. When leaching experiments

137

were finished, the residue and leachate were separated using a filter paper with a mesh opening of

138

0.02 µm.

139

Depending on the difference of solubility product (Ksp) and pH range of precipitation, Cu2+ and

140

Al3+ ions in leachate can be precipitated thoroughly. As shown in Table S1, the pH values of Cu2+ and

141

Al3+ ions reaching the minimum concentration were 4.49 and 6.65, respectively. In order to prevent

142

the simultaneous precipitation of Ni element at low pH, the key is to dilute the concentration of Ni2+

143

to less than 1 M (Lain M., et al. 2001; Zou, H., et al. 2013). In this experiment, the concentrations of

144

all the metal ions (Al, Cu, Ni, Co and Mn) were first adjusted to below 0.5 M after acid leaching, and

145

then the pH value could increase slowly to 6.65 by adding 1 M of NaOH. The Cu(OH)2 and Al(OH)3,

146

separated through filtration, were stirred in other NaOH solution at pH=12 for 2 h. After re-filtration

147

and heating, Cu(OH)2 residue and NaAlCO3(OH)2 filtrate could be translated into CuO and

148

NaAlCO3(OH)2 powders, respectively. 7

149

The leachate retrieved after removing Cu2+ and Al3+ is a neutral solution for hydrothermal

150

reaction. The specific steps were that 50 ml leachate and 10 g diaminomethanal were placed in a

151

100-ml hydrothermal reactor and reacted for 12 hours at 160℃. The NCM precursors could be

152

obtained by vacuum filtration with large surface morphology. The pH of residual solution after

153

hydrothermal reaction was adjusted to 12, to remove all the heavy metal ions in the solution. The

154

Li2CO3 precipitate was obtained with the adding of Na2CO3 and washed with boiling ultra-pure

155

water.

156 157

2.4 Analytical methods

158

The concentration of metal elements in solution were measured with an inductively coupled

159

plasma-optical emission spectrometer (ICP-OES, OPTIMA 2000, PerkinElmer, USA). In addition,

160

the leaching rate of each element was calculated with the following equation:

161

W wt % =

× 100%

162

where

163

the solid material before acid leaching.

(1)

is the final content of metals in solution, and

denotes the original content of metals in

164

The crystal structures and phase analysis of the solid powders were characterized by X-ray

165

diffraction (XRD, Philips PW 1700, USA) using Cu Kα radiation, and the data were analyzed by

166

MDI Jade 6.0 software. The microscopic morphology and content distribution of surface elements

167

were characterized by field emission scanning electron microscopy, which contained an energy

168

dispersive spectrometer (FE-SEM and EDX, Carl Zeiss MERLIN Compact, Germany).

169 8

170

3. Results and discussion

171

3.1 Feasibility analysis

172

High-nickel batteries are distinguished from conventional NCM batteries in that they have better

173

electrochemical performance and structural stability, but this also increases the difficulty of recycling.

174

The use of Cu and Al as reducing agents for leaching Li(Ni1/3Co1/3Mn1/3)O2 (namely 111 ternary

175

cathode materials) with sulfuric acid has proved to be effective, but no similar research has been

176

done on the feasibility of recovering high-nickel ternary materials. The combination of H2SO4 and

177

H2O2, though, is generally considered to have a good leaching effect. In this simulation experiment,

178

mass ratio of cathode material to Cu and Al was 2:1:1. The final leaching results were shown in

179

Figure 2, and suggested that only 40% of the ternary materials could be dissolved by pure dilute

180

H2SO4; its effect was poor. The addition of H2O2 could help dissolve 90% of the cathode materials.

181

More gratifying was that the addition of Cu and Al could recover almost 100% of all the metal

182

elements, by leaching, mainly due to the reduction of insoluble Mn4+ and Co3+ to soluble Mn2+ and

183

Co2+. However, determining the reason Al and Cu powders have a better reduction effect requires an

184

in-depth analysis of their Gibbs free energy in leaching process.

185 186

3.2 Thermodynamic analysis

187

The good electrochemical performance of the NCM battery derives from its stable layered

188

structure, which mainly includes transition metal oxides with embedded lithium elements. It can be

189

determined through assimilation that the NCM ternary material is composed of a stack of transition

190

metal oxides—namely NiO, MnO2 and Co2O3—with inserted lithium (Joulié, M., et al. 2017). In 9

191

particular, bivalent nickel is easily soluble in dilute H2SO4 and does not participate in reduction

192

reaction. Therefore, the reaction of the NCM electrode material with the above reducing agents (Cu,

193

Al and H2O2) can be approximated by the reaction of MnO2 and Co2O3. In the case of Co2O3, three

194

different redox reactions may occur, and their standard Gibbs free energies are listed below (Wang L.,

195

2013). 3Co2O3 + 2Al + 9H2SO4 → 6CoSO4 + Al2(SO4)3 + 9H2O = −477.636 − 0.02684 ,

0~1000!

Co2O3 + Cu + 3H2SO4 → 2CoSO4 + CuSO4 + 3H2O = −69.497 − 0.01245 ,

0~1000!

Co2O3 + H2O2(l) + 2H2SO4 → 2CoSO4 + 3H2O + O2(g) = −37.262 − 0.04583 ,

0~1000!

(1) (2) (3) (4) (5) (6)

196

Similarly, trivalent manganese ions also need to be reduced to divalent transition metal ions; the

197

corresponding standard Gibbs free energies were calculated by HSC Chemistry 6.0 as shown in

198

follows: 3MnO2 + 2Al + 6H2SO4 → 3MnSO4 + Al2(SO4)3 + 6H2O = −491.349 + 0.20149 ,

0~1000!

MnO2 + Cu + 2H2SO4 → MnSO4 + CuSO4 + 2H2O = −74.024 + 0.0635 ,

0~1000!

MnO2 + H2O2(l) + H2SO4 → MnSO4 + 2H2O + O2(g) = −26.677 − 0.03372 , 199

0~1000!

(7) (8) (9) (10) (11) (12)

Figure 3 shows the relationship between standard Gibbs free energy and temperature in different 10

200

reactions (1-12). In Figure 3, the reactions of Al, Cu and H2O2 as reductants are marked in green,

201

blue and black, respectively. It is well known that if the Gibbs free energy is lower than zero and no

202

other potential energy is generated, the reaction can proceed spontaneously; otherwise the reaction

203

cannot occur. According to the curve trend in Figure 3, only the standard Gibbs free energy of

204

reaction 9 may be higher than 0, within 1000 K. Therefore, based on Eq. 13, the reaction (9) was

205

subjected to further quantitative calculation to determine its Gibbs free energy at specific

206

temperatures. G =

+ R ln

)*+ )

(13)

207

where R is the molar gas constant, 8.314 J/(mol· K); T is the thermodynamic temperature, degrees

208

Kelvin; )*+ is the partial pressure of the gas product; and )

209

When the reaction is carried out in an air atmosphere without applied pressure, this

210

approximately equal to 0.21, and Eq. 10 should change to: G = −74.024 + 0.0508 ,

is the standard atmospheric pressure.

0~1000!

, -+ ,.

value is

(14)

G < 0, 4-valent manganese ions may theoretically be reduced to

211

On the basis of Eq.14, when

212

2-valent manganese ions by copper, within 1457.165 K (1184℃).

213

The leaching reaction of NCM materials occurs in water-based solution, so the range of reaction

214

temperature is usually 273.15-373.15 K (0-100℃). The variation of Gibbs free energy for each

215

reaction in this temperature range is presented in Figure 3 with a champagne background. It indicates

216

that the absolute value order of Gibbs free energy for each reaction within the temperature range of 0

217

-100 ℃ is (1) > (7) > (3) > (9) ≈ (5) > (11). In other words, the order of reducibility is Al > Cu >

218

H2O2, and the reaction rate is Al > Cu > H2O2. Since there is a large gap between Al and Cu both in 11

219

the Gibbs free energy and the reductivity sequence of the metals (Al > hydrogen (H) > Cu), we can

220

conjecture that the cathode materials preferentially react with Al foils. Only after the Al foils are

221

completely consumed, the cathode materials begin to react with the Cu foils in dilute sulfuric acid.

222

On the whole, the reduction of high-valent transition metal ions in NCM materials by Al and Cu is

223

efficient, and the reduction effect is better than that of H2O2, in theory, a result consistent with those

224

of previous experiments (Joulié, M., et al. 2017).

225

More importantly, according to the reaction equations (1) and (7), 2 moles of Al can react with 3

226

moles of Co2O3 or MnO2, implying that 1 mole of Al can induce at least 1.5 moles of NCM cathode

227

materials to be completely dissolved in 3 moles of H2SO4 by the redox reaction: in other words, 1

228

gram of Al can catalyze about 5.36 grams of ternary cathode material in 10.89 grams of H2SO4.

229

Similarly, 1 gram of Cu can catalyze about 1.52 grams of NCM powders in 3.08 grams of H2SO4.

230

For simplicity of calculation, it can be considered that 1 g Al is equivalent to 5 g NCM material and

231

10 g H2SO4, while 1 g Cu is equivalent to 1.5 g NCM material and 3 g H2SO4. It should be noted that

232

in this calculation, the total content of Co2O3 and MnO2 in the NCM material is approximately 100%,

233

but their actual content is only 30-70% in 111, 523, 532, 622, 811 cathode material. Therefore, this

234

result can be applicable to all kinds of NCM cathode materials.

235 236

3.3 Fine management of Cu/Al foils through mild screening

237

As the simplest method of separating Cu/Al foils and electrode materials, mechanical crushing

238

has the most practical application prospects. In this section, we hoped to find the best proportion of

239

Cu/Al foils to cathode materials by adjusting the sieving size of the mild screening, based on the 12

240

above thermodynamic analysis.

241

After efficient crushing and mild screening of spent LIBs, the morphology and element

242

distribution of each component are shown in Figure 4. It indicates that 60.51 wt.% of the crushed

243

materials were enriched in the -0.5 mm particle size, while the material contents of +0.5-0.9 mm,

244

+0.9-1.45 mm, +1.45-2 mm, +2-4 mm and +4 mm were 12.24%, 4.89%, 3.61%, 5.68% and 5.88%,

245

respectively. Besides, the metal content in -0.5 mm fine materials was 57.24%, and the remaining

246

42.76% was non-metallic graphite. It should be noted that -0.5mm particles mean that the particle

247

size is less than 0.5mm, while +0.5-0.9 mm particles mean that the particle size is greater than

248

0.5mm and less than 0.9mm. Furthermore, the experimental results show that the content of electrode

249

materials in the -0.5 mm particles is as high as 98.94%, and the content of Cu/Al foils in the +4 mm

250

crude material could reach 94.19%. This remarkable selective crushing behavior is consistent with

251

other scholars’ research (Zhang T., et al. 2013). In order to determine the optimum screening size, the

252

Al and Cu contents in Figure 4 were converted into NCM cathode material equivalents: the materials

253

were then compared by sieving the cumulative specimens. The results are shown in Figure 5 and

254

Table S2, and it illustrates that the cumulative content of actual NCM materials increased slightly,

255

from 33.99% in -0.5mm level to 41.29% in total. Nevertheless, with the increasing accumulative

256

amount of Cu/Al foils, the NCM equivalent content (initially 2.28%) gradually caught up with the

257

actual content of NCM, finally reaching 73.36%. More importantly, when mild screening was carried

258

out with 4 mm sieve holes, the contents of Al/Cu foils in the screening product were 6.30 wt.% and

259

9.98 wt.% respectively, converted into NCM equivalents of 31.49 wt.% and 14.97 wt.% respectively.

260

In fact, only 41.13 wt.% of the products under mild screening were NCM cathode materials, while 13

261

the total content of NCM equivalent from Al/Cu foils was 46.46 wt.%, which was 5.33 wt.% higher

262

than the actual content. Therefore, 4 mm sieve holes are the optimal screening conditions.

263

Furthermore, if 100g of this mild screening products need to process, containing 6.3 g Al, 9.98 g Cu,

264

and 41.13g NCM materials, the demand for H2SO4 is 92.94 g. Consequently, the mass ratio of

265

sulfuric acid to sieving product is about 1:1, while the specific mass ratio of H2SO4 to Cu, Al and

266

NCM materials is about 2:3:15:30.

267

Large pore size not only reduces the difficulty of materials’ passing through the mesh, but also

268

controls the proportion of reductant for subsequent leaching reactions. This fine management

269

strategy of Cu/Al foils subtly solve the industrialization restriction brought by 0.075 mm deep

270

screening, and paves the way for the cascade recycling of copper and aluminum elements.

271 272

3.4 Optimum leaching condition

273

In this section, the effects of temperature, acid concentration and liquid-solid ratio on leaching

274

results were considered comprehensively, as shown in Figure 6. In order to explore the optimal

275

conditions of liquid-solid ratio, the experimental temperature was controlled at 60°C and the

276

concentration of sulfuric acid at 3 mol/L. As shown in Figure 6A, the liquid-solid ratio of 20 mL/g

277

could only dissolve about 25% of the valuable metal elements. In contrast, a higher liquid-solid ratio,

278

such as 200 mL/g, could dissolve more than 90% of the cathode materials. When the liquid-solid

279

ratio reached 400 mL/g, the NCM cathode materials could be completely dissolved. In terms of total

280

component recovery and economic value, the best liquid to solid ratio was 200 mL/g. When

281

exploring the optimum temperature conditions, the liquid-solid ratio of the reaction was kept at 200 14

282

mL/g, and the acid concentration was 3 mol/L. Temperature is an important condition for reducing

283

the dissolution energy barrier. In this experiment, the reaction temperatures were set to 0℃, 30℃,

284

45℃, 60℃, 75℃ and 90℃, and the results are shown in Figure 6B. Due to the increased diffusivity

285

and decreased viscosity of liquid, the leaching rate of each metal increased with the increase of

286

temperature, and reached a maximum at 60℃. As the temperature continued to increase, the leaching

287

rate showed a small fluctuation, but remained stable on the whole. Accordingly, considering the

288

economics and security of the heating process, the optimal temperature was 60℃.

289

The optimal concentration of dilute sulfuric acid was explored at 60℃ and 200 mL/g. As seen in

290

Figure 6C, with the increase of dilute sulfuric acid concentration, the leaching rates of Li, Ni, Co, Mn

291

first increased and then decreased. The main reason may be that there is a side reaction here and the

292

reaction equation is as follows. 4Al + 3O3 + 6H3 SO6 → 2Al3 SO6 = −885.442 + 0.17961 ,

+ 6H3 O

(15)

0~1000!

(16)

8

293

When the H2SO4 concentration is low (0.5-2.5 mol/L), most Al foils play a role of reducing agent to

294

promote the leaching of cathode materials. Nevertheless, NCM cathode material is a kind of

295

spherical secondary particles composed of Li2O, NiO, Co3O4 and MnO2 micro-particles. The

296

leaching and release of these metal ions into liquid phase are relatively slow. When the H2SO4

297

concentration is high (2.5-4.0 mol/L), a large number of hydrogen ions will consume some Al foils,

298

resulting in the decrease of reductant in the system. Thus, the leaching of metal ions from ternary

299

materials is insufficient at the same time unit. Figure 6C shows that the optimum leaching rate

300

appeared at 2.0-2.5 mol/L H2SO4. From the viewpoint of cost savings and environmental load 15

301

reduction, dilute sulfuric acid of 2.0 mol/L could be the optimum leaching condition.

302

Therefore, the optimum conditions for leaching 523 NCM cathode materials with Cu/Al foils

303

inherent in spent LIBs are: 4 mm mild screening, liquid-solid ratio of 200 mL/g, reaction temperature

304

of 60℃ and dilute sulfuric acid concentration of 2 mol/L.

305 306

3.5 Purification and characterization

307

H+, Li+, Ni2+, Co2+, Mn2+, Al3+, Cu2+ and SO42- were all present in the leaching solution. The

308

interactions of these six metal cations made the solution system more chaotic, resulting in a great

309

challenge of multi-metal separation and purification. Fortunately, there is a distinct pH gap between

310

Al3+, Cu2+ and Ni2+, Co2+, Mn2+ during precipitation, Namely, Al3+ and Cu2+ could be pre-removed

311

first, at pH less than 6.5, and then Ni2+, Co2+, Mn2+ could be co-precipitated completely when the pH

312

was more than 12. For the mixture of Al(OH)3 and Cu(OH)2, Al(OH)3 precipitation can be converted

313

into NaAlCO3(OH)2 solution by adding NaOH solution, so that copper-rich solid and aluminum-rich

314

solution can be separated by filtration. The possible dissolution reaction equations are as follows. Al OH

8

+ 9:3 + ;<:= → ;<>?9:8 :=

= −36.768 + 0.03472 ,

3

+ =3 :

0~1000!

(17) (18)

315

For further surface morphology and phase composition analysis, the blue-violet liquid

316

(NaAlCO3(OH)2 solution) was dried and crystallized, and then subjected to SEM and XRD tests,

317

whose results were shown in Figures 7B and C. The SEM image indicates that the basic carbonates

318

are mainly flocculent and flaky, with some rod-like crystals. The XRD pattern suggests that only the

319

diffraction peaks of the NaAlCO3(OH)2 match those of the materials exactly, and the standard card 16

320

number is 45-1359. Since the detection limit of XRD is 5% and no other impurity peaks occurred,

321

the purity of the NaAlCO3(OH)2 can be considered to be higher than 95%. In this recovery system,

322

the Al foil powder can be recovered in the form of crude Al sheets after mild screening and

323

high-quality NaAlCO3(OH)2 after reductive leaching, which realizes the cascade recovery to some

324

extent.

325

After application of the co-precipitation and the hydrothermal method, the ternary precursor

326

was produced. In order to analyze the quality of the product, the precursors were characterized by

327

FE-SEM, EDX and XRD in detail, and the results are shown in Figure 7. Due to changes in the

328

contents of nickel-cobalt-manganese elements and crystalline hydrate, there was no unified chemical

329

structure for the precursor, and no single XRD phase standard card corresponding to it. From the

330

XRD pattern in Figure 7D, three diffraction characteristic peaks of MnO·OH, CoO and Ni(OH)2

331

existed in the precursor phase, and analyzed under the standard cards of 18-0804, 42-1300 and

332

14-0117, respectively. In general, these ternary precursor materials belong to a mixture of metal

333

oxides or metal hydroxides, which can be understood as a form of NiO·MnO·CoO·(OH)x. Figure 7A,

334

E and F show the microscopic surface topography of the precursor. These indicate that the precursor

335

prepared by the diaminomethanal-hydrothermal method is a type of chrysanthemum-cluster spherical

336

particle. The obvious wrinkles and clear stripes on the surface of the particles prove that the

337

precursor will have a good layered structure after calcination and hold a good electrochemical

338

performance potential, as shown in S1 (He, Q., et al. 2017; Zhang, F., et al. 2017; Chen, M., et al.

339

2019). In order to observe the element distribution on the surface of particles more intuitively, EDX

340

energy spectrum analysis was carried out, as shown in Figure 7G-L. The distributions of nickel, 17

341

cobalt, manganese and oxygen elements on the surface of the particles are relatively uniform,

342

consistent with the XRD results. Relative quantitative analysis of the elements shows that the

343

contents of nickel, cobalt, manganese and oxygen were 25.9 wt.%, 17.2 wt.%, 13.8 wt.% and 43.1

344

wt.%, respectively. All these characterizations indicate that the precursors have good performance

345

potential.

346 347

3.6 A green recycling route of spent ternary lithium-ion batteries for industrial application

348

With the preparation of the key products NaAlCO3(OH)2 and NCM precursors, we were able to

349

introduce a simple technical flowchart for manufacturing advanced materials from spent LIBs.

350

However, the main factors determining if a technical route can be industrialized are whether it can

351

recover as much valuable product as possible, whether the input cost of equipment is appropriate,

352

and whether labor costs can be contained. Taking the above factors into consideration, this study

353

optimized the main technology of mild screening and multi-step precipitation, and obtained a new

354

green recycling route, as shown in Figure 8.

355

The mild screening technology breaks through the bottleneck of LIB recycling, and is one of the

356

core technical points of the entire route. The process developed in this study was as follows: the

357

crushed products of spent LIBs were screened slightly by a 4-mm sieve; the coarse materials with

358

particle size larger than 4 mm were spherical-like or sheet-shaped Al and Cu foils, which could be

359

separated by color sorting (Zhong, X., et al. 2019). The fine powder of Cu/Al foils, under mild

360

screening, entered the leaching step, along with the electrode material, and were eventually

361

recovered in the form of CuO and NaAlCO3(OH)2, respectively. It is here that practice the concepts 18

362

of fine management and cascade recycling of Cu/Al foils. For the leaching process, all the metal

363

elements entered into an aqueous solution; only the graphite from the anode electrode remained as

364

solid residue, and this could be separated out by filtering. The literature (Guo, Y., et al. 2016) shows

365

that acid leaching can modify the layered structure of graphite, so that it can also be sold as a single

366

product. In addition, graphite can also be regard as adsorbent to purify organic electrolyte impurities

367

in waste liquids produced by this technology and to promote the recycling of alkali liquor in the

368

system (Rodrigues T.M., et al. 2016; Natarajan, S., et al. 2015). Therefore, the technological route

369

can realize the recovery of Cu and Al foils, graphite, CuO, NaAlCO3(OH)2, NCM precursor and

370

lithium carbonate from spent LIBs. It is noteworthy that the entire technology given in Figure 8

371

requires no manual assistance; the entire route can be designed as a fully mechanized production line

372

or facility. Since it does not involve intense chemical reactions and the main processes are

373

conventional technologies such as crushing, sieving, room-temperature leaching and precipitation,

374

equipment investment keeps at a very low level.

375

Last but not least, the global recovery rates of main metals from the spent LIBs have been listed

376

in Table 1. In this research, two precipitation processes were involved, namely, precipitate Ⅰ of

377

Cu-Al at pH 6.5 and precipitate Ⅰ of Ni-Co-Mn at a pH of 12. As a matter of fact, it is difficult to

378

completely separate Ni2+ from Cu2+ based on solubility differences. It suggests that the total removal

379

of Cu ions in solution requires an increase in pH to 6.65, while Ni ions begin to precipitate when pH

380

reaches 5.156. This shows that if we want to recover more Ni metal, the pH of precipitate Ⅰ only

381

needs to be controlled at about 5.0. And if we want to recover high-quality Ni metal and avoid the Cu

382

impurities, the pH of precipitate Ⅰ needs to be controlled at about 6.5, but this will inevitably lead to 19

383

the loss of some Ni metal. From the economic point of view, we have chosen the latter scheme. It

384

needs to be stated that if the pH of precipitate Ⅰ is controlled at about 6.5, 0.05 wt.% of Cu will still

385

be allowed to enter precipitate Ⅰ and then co-precipitate with the precursor. Fortunately, (Yang, L., et

386

al. 2018) showed that Cu doping can optimize the electrochemical performance of NCM cathode

387

materials. Therefore, although the recovery rate of Ni element in this process is relatively low and

388

the major loss is in precipitate Ⅰ, the recovery rates of main metal in spent LIBs have remained

389

above 95% from the overall perspective. It proves that this green sustainable technology also meets

390

the requirement of economic feasibility.

391 392

4. Conclusion

393

Fine management and cascade recovery of Cu/Al materials are proposed for the first time to

394

assist in high-quality recycling of full valuable materials from spent ternary LIBs. The key problem

395

of deep screening has been revealed and solved to avoid alkali leaching or any other toxic solution

396

(NMP) to remove Cu and Al impurities before acid leaching, practicing the concept of green and

397

sustainable development. Compared with the inadequate management and recycling of Cu/Al foils in

398

the past, this study takes full advantages of their selective crushing behavior and excellent reduction

399

performance, involving direct recovery of coarse globular-like particles and refinement extraction of

400

fine powders. Using thermodynamic theories and experimental verification, the strategy of leaching

401

NCM electrode materials assisted by fine Cu/Al foils is superior to using traditional additive

402

reductants. Almost 100% of the valuable metal elements in the fine crushed powders are dissolved

403

into aqueous solution and eventually converted into CuO, NaAlCO3(OH)2, NCM precursor, and 20

404

Li2CO3 products. In addition, further characterization of the novel products indicates that the purity

405

of NaAlCO3(OH)2 was more than 95%, and that the surface of the NCM precursor was evenly folded,

406

indicating a promising potential to form a good layered structure. This strategy breaks through the

407

long-standing gap between theoretical research and practical application, and may help reduce the

408

consumption of chemicals in the waste battery recycling industry. The optimal sintering conditions of

409

ternary precursors and the final electrochemical properties need further study.

410 411 412 413

Acknowledgments This research is supported by financial supports by “National Natural Science Foundation of China” (71804085) and “National Key Technology R&D Program of China” (2018YFC1900101).

414 415

Reference:

416

Chen, M., Zheng, Z., Wang, Q., Zhang, Y., Ma, X., Shen, C., Xu, D., Liu, J., Liu, Y., Gionet, P., O

417

Connor, I., Pinnell, L., Wang, J., Gratz, E., Arsenault, R., Wang, Y., 2019. Closed Loop Recycling of

418

Electric Vehicle Batteries to Enable Ultra-high Quality Cathode Powder. Sci. Rep., 9 1654-1662.

419

Chen, X., Chen, Y., Zhou, T., Liu, D., Hu, H., Fan, S., 2015. Hydrometallurgical recovery of metal

420

values from sulfuric acid leaching liquor of spent lithium-ion batteries. Waste Manage. 38, 349-356.

421

Contestabile, M., Panero, S., Scrosati, B., 2001. A laboratory-scale lithium-ion battery recycling

422

process. J. Power Sources 92, 65-69.

423

Doberdò, I., Löffler, N., Laszczynski, N., Cericola, D., Penazzi, N., Bodoardo, S., Kim, G., Passerini,

424

S., 2014. Enabling aqueous binders for lithium battery cathodes – Carbon coating of aluminum

425

current collector. J. Power Sources 248, 1000-1006. 21

426

Dos Santos, C.S., Alves, J.C., Da Silva, S.P., Evangelista Sita, L., Da Silva, P.R.C., de Almeida, L.C.,

427

Scarminio, J., 2019. A closed-loop process to recover Li and Co compounds and to resynthesize

428

LiCoO2 from spent mobile phone batteries. J. Hazard. Mater. 362, 458-466.

429

Ferreira, D.A., Prados, L.M.Z., Majuste, D., Mansur, M.B., 2009. Hydrometallurgical separation of

430

aluminium, cobalt, copper and lithium from spent Li-ion batteries. J Power Sources, 187, 238-246.

431

Gao, W., Zhang, X., Zheng, X., Lin, X., Cao, H., Zhang, Y., Sun, Z., 2017. Lithium carbonate

432

recovery from cathode scrap of spent lithium-ion battery: a closed-loop process. Environ. Sci.

433

Technol. 51, 1662-1669.

434

Golmohammadzadeh, R., Faraji, F., Rashchi, F., 2018. Recovery of lithium and cobalt from spent

435

lithium ion batteries (LIBs) using organic acids as leaching reagents: A review. Resour. Conserv.

436

Recy. 136, 418-435.

437

Gratz, E., Sa, Q., Apelian, D., Wang, Y., 2014. A closed loop process for recycling spent lithium ion

438

batteries. J. Power Sources 262, 255-262.

439

Guo, Y., Li, F., Zhu, H., Li, G., Huang, J., He, W., 2016. Leaching lithium from the anode electrode

440

materials of spent lithium-ion batteries by hydrochloric acid (HCl). Waste Manage. 51, 227-233.

441

Hannan, M.A., Hoque, M.M., Mohamed, A., Ayob, A., 2017. Review of energy storage systems for

442

electric vehicle applications: Issues and challenges. Renew. Sustain. Energy Rev. 69, 771-789.

443

He, Q., Gu, S., Wu, T., Zhang, S., Ao, X., Yang, J., Wen, Z., 2017. Self-supported mesoporous

444

FeCo2O4 nanosheets as high capacity anode material for sodium-ion battery. Chem. Eng. J. 330,

445

764-773.

446

Hu, J., Zhang, J., Li, H., Chen, Y., Wang, C., 2017. A promising approach for the recovery of high 22

447

value-added metals from spent lithium-ion batteries. J. Power Sources 351, 192-199.

448

Joulié, M., Laucournet, R., Billy, E., 2014. Hydrometallurgical process for the recovery of high value

449

metals from spent lithium nickel cobalt aluminum oxide based lithium-ion batteries. J. Power

450

Sources 247, 551-555.

451

Joulié, M., Billy, E., Laucournet, R., Meyer, D., 2017. Current collectors as reducing agent to

452

dissolve active materials of positive electrodes from Li-ion battery wastes. Hydrometallurgy, 169,

453

426-432.

454

Lain M., Recycling of lithium ion cells and batteries. J. Power Sources, 97(2001) 736-738.

455

Li, L., Fan, E., Guan, Y., Zhang, X., Xue, Q., Wei, L., Wu, F., Chen, R., 2017. Sustainable Recovery

456

of Cathode Materials from Spent Lithium-Ion Batteries Using Lactic Acid Leaching System. ACS

457

Sustain. Chem. Eng. 5, 5224-5233.

458

Li, J., Wang, G., Xu, Z., 2016. Environmentally-friendly oxygen-free roasting/wet magnetic

459

separation technology for in situ recycling cobalt, lithium carbonate and graphite from spent LiCoO2

460

/graphite lithium batteries. J. Hazard. Mater. 302, 97-104.

461

Liu, C., Neale, Z.G., Cao, G., 2016. Understanding electrochemical potentials of cathode materials in

462

rechargeable batteries. Mater. Today 19, 109-123.

463

Loeffler, N., von Zamory, J., Laszczynski, N., Doberdo, I., Kim, G., Passerini, S., 2014. Performance

464

of LiNi1/3Mn1/3Co1/3O2/graphite batteries based on aqueous binder. J. Power Sources 248, 915-922.

465

Ma, X., Ma, Y., Zhou, J., Xiong, S., 2018. The Recycling of Spent Power Battery: Economic

466

Benefits and Policy Suggestions. IOP Conference Series: Earth and Environmental Science 159,

467

12017. 23

468

Natarajan, S., Shanthana Lakshmi, D., Bajaj, H.C., Srivastava, D.N., 2015. Recovery and utilization

469

of graphite and polymer materials from spent lithium-ion batteries for synthesizing polymer–graphite

470

nanocomposite thin films. J. Environ. Chem. Eng., 3, 2538-2545.

471

Nayaka, G.P., Pai, K.V., Manjanna, J., Keny, S.J., 2016. Use of mild organic acid reagents to recover

472

the Co and Li from spent Li-ion batteries. Waste Manage. 51, 234-238.

473

Peng, C., Liu, F., Aji, A.T., Wilson, B.P., Lundström, M., 2019. Extraction of Li and Co from

474

industrially produced Li-ion battery waste – Using the reductive power of waste itself. Waste

475

Manage. 95, 604-611.

476

Rodrigues T.M., Kavashima L., Adas E.G., Antoniassi B., Chaves M.R., 2016. Adsorption of

477

rhodamine B from aqueous effluents on graphite from spent lithium-ion battery anode. Int. J. Res.

478

Eng. Sci. 4, 27-36.

479

Roshanfar, M., Golmohammadzadeh, R., Rashchi, F., 2019. An environmentally friendly method for

480

recovery of lithium and cobalt from spent lithium-ion batteries using gluconic and lactic acids. J.

481

Environ. Chem. Eng. 7, 102794.

482

Sattar, R., Ilyas, S., Bhatti, H.N., Ghaffar, A., 2019. Resource recovery of critically-rare metals by

483

hydrometallurgical recycling of spent lithium ion batteries. Sep. Purif. Technol. 209, 725-733.

484

Träger, T., Friedrich, B., Weyhe, R., 2015. Recovery Concept of Value Metals from Automotive

485

Lithium-Ion Batteries. Chem. Ing. Tech. 87, 1550-1557.

486

Vieceli, N., Nogueira, C.A., Guimarães, C., Pereira, M.F.C., Durão, F.O., Margarido, F., 2018.

487

Hydrometallurgical recycling of lithium-ion batteries by reductive leaching with sodium

488

metabisulphite. Waste Manage. 71, 350-361. 24

489

Wang, J., Pyo, J., Ahn, S., Choi, D., Lee, B., Lee, D., 2018. A study on the recovery of Li2CO3 from

490

cathode active material NCM(LiNiCoMnO2) of spent lithium ion batteries. J. Korean Powder Metall.

491

Inst. 25, 296-301.

492

Wang L., Research on thermodynamics and influencing factors of reaction of LiCoO2 in the acid

493

roasting conditions, in: Master Dissertation, Lanzhou University of Technology, Lanzhou, 2013 (in

494

Chinese).

495

Wang, M., Tan, Q., Li, J., 2018. Unveiling the Role and Mechanism of Mechanochemical Activation

496

on Lithium Cobalt Oxide Powders from Spent Lithium-Ion Batteries. Environ. Sci. Technol. 52,

497

13136-13143.

498

Wang, M., Tan, Q., Liu, L., Li, J., 2019. A low-toxicity and high-efficiency deep eutectic solvent for

499

the separation of aluminum foil and cathode materials from spent lithium-ion batteries. J. Hazard.

500

Mater. 380, 120846.

501

Wood, L., 2018. Global Lithium-ion Battery Market Report 2018: Market is Expected to Reach $47

502

Billion

503

https://www.prnewswire.com/news-releases/global-lithium-ion-battery-market-report-2018-market-i

504

s-expected-to-reach-47-billion-by-2023-from-25-billion-in-2017-300744613.html.

505

November 2018)

506

Xiao, J., Li, J., Xu, Z., 2017. Novel Approach for in Situ Recovery of Lithium Carbonate from Spent

507

Lithium Ion Batteries Using Vacuum Metallurgy. Environ. Sci. Technol. 51, 11960-11966.

508

Yang, L., Ren, F., Feng, Q., Xu, G., Li, X., Li, Y., Zhao, E., Ma, J., Fan, S., 2018. Effect of Cu

509

Doping on the Structural and Electrochemical Performance of LiNi1/3Co1/3Mn1/3O2 Cathode

by

2023,

from

$25

25

Billion

in

(Accessed

2017.

06

510

Materials. J. Electron. Mater. 47, 3996-4002.

511

Yu, J., He, Y., Li, H., Xie, W., Zhang, T., 2017. Effect of the secondary product of semi-solid phase

512

Fenton on the flotability of electrode material from spent lithium-ion battery. Powder Technol. 315,

513

139-146.

514

Yu, J., He, Y., Ge, Z., Li, H., Xie, W., Wang, S., 2018. A promising physical method for recovery of

515

LiCoO 2 and graphite from spent lithium-ion batteries: Grinding flotation. Sep. Purif. Technol. 190,

516

45-52.

517

Zhang, F., Bao, Y., Ma, S., Liu, L., Shi, X., 2017. Hierarchical flower-like nickel phenylphosphonate

518

microspheres and their calcined derivatives for supercapacitor electrodes. J. Mater. Chem. A, 5,

519

7474-7481.

520

Zhang, T., He, Y., Ge, L., Fu, R., Zhang, X., Huang, Y., 2013. Characteristics of wet and dry

521

crushing methods in the recycling process of spent lithium-ion batteries. J. Power Sources, 240,

522

766-771.

523

Zhang, X., Li, L., Fan, E., Xue, Q., Bian, Y., Wu, F., Chen, R., 2018. Toward sustainable and

524

systematic recycling of spent rechargeable batteries. Chem. Soc. Rev. 47, 7239-7302.

525

Zheng, X., Zhu, Z., Lin, X., Zhang, Y., He, Y., Cao, H., Sun, Z., 2018. A Mini-Review on Metal

526

Recycling from Spent Lithium Ion Batteries. Engineering-PRC 4, 361-370.

527

Zhong, X., Liu, W., Han, J., Jiao, F., Qin, W., Liu, T., Zhao, C., 2019. Pyrolysis and physical

528

separation for the recovery of spent LiFePO4 batteries. Waste Manage. 89, 83-93.

529

Zou, H., Gratz, E., Apelian, D., Wang, Y., 2013. A novel method to recycle mixed cathode materials

530

for lithium ion batteries. Green Chem. 15, 1183-1191. 26

Table 1. Global recovery rates of main metals from the spent LIBs Element

Al

Cu

Li

Ni

Co

Mn

Recovery rate / %

97.85

99.42

98.23

95.45

98.37

98.64

Products

Al foil/ NaAlCO3(OH)2

Cu foil/ CuO

Li2CO3

Precursor

Figure 1. The XRD pattern of the electrode materials of spent Lithium-ion battery

Figure 2. Leaching rate of transition metals using different methods (Especially, Al and Cu powder were analytical reagent)

Figure 3. The relationship between Gibbs free energy and temperature in different reactions

Figure 4. The morphology and element distribution of each component of spent LIBs after efficient crushing and mild screening

Figure 5. Actual content and equivalent content of NCM cathode material under different screening sizes

Figure 6. Effect of different experimental conditions on leaching efficiency (A. Liquid to solid ratio; B. Temperature; C. Acid concentration;

Especially, the -4 mm crushing powders of spent LIBs were the raw materials.)

Figure 7. Crystal structure, surface morphology and element distribution analysis of novel products (For NCM precursors: A.E.F. Surface morphology by FE-SEM; D. Crystal structure by XRD; G-L. Element distribution and quantitative analysis by EDX; For NaAlCO3(OH)2: B. Surface morphology by FE-SEM; C. Crystal structure by XRD)

Figure 8. A green recycling route of spent ternary lithium batteries for industrial application

1. The fine management and cascade recovery of Cu/Al foils are first proposed. 2. Thermodynamic analysis of Cu/Al foils as reducing agent was carried out. 3. Multi-step precipitation and hydrothermal method were used for purification. 4. The deep screening problem has been solved in a green and sustainable way.

Declaration of interest statement The publication of this work has been approved by all authors. This paper has not been published/submitted or being submitted to another journal. We believe that this manuscript will be of exceptional interest to the specialist in related research field. If accepted, it will not be published elsewhere in the same form, in English or in any other language, without the written consent of the Publisher. We confirm that this research is supported by financial supports by “National Natural Science Foundation of China” (71804085) and “National Key Technology R&D Program of China” (2018YFC1900101). All financial support for this project comes from legal channels. All sources of funds do not involve other enterprises, research institutions and other units, and do not involve property rights disputes.