Cross-Linked Chitosan as an Efficient Binder for Si Anode of Li-ion

Jan 8, 2016 - Effect of Binder Architecture on the Performance of Silicon/Graphite ... electrochemical performance as anode materials for Li-ion batte...
1 downloads 4 Views 6MB Size
Research Article www.acsami.org

Cross-Linked Chitosan as an Efficient Binder for Si Anode of Li-ion Batteries Chao Chen,†,§ Sang Ha Lee,† Misuk Cho,† Jaehoon Kim,‡ and Youngkwan Lee*,† †

School of Chemical Engineering and ‡School of Mechanical Engineering, Sungkyunkwan University, 440-746 Suwon, Korea § School of Materials and Textile Engineering, Jiaxing University, 314-001 Jiaxing, China S Supporting Information *

ABSTRACT: We investigate the use of chitosan (CS) as a new cross-linkable and water-soluble binder for the Si anode of Li-ion batteries. In contrast to the traditional binder utilizing a hydrogen bond and/or van der Waals force-linked anode electrodes, CS can easily form a 3D network to limit the movement of Si particles through the cross-linking between the amino groups of CS and the dialdehyde of glutaraldehyde (GA). Chemical, mechanical, and morphological analyses are conducted by Fourier transform infrared spectroscopy, tensile testing, and scanning electron microscopy. The cross-linked Si/CS-GA anode exhibits an initial discharge capacity of 2782 mAh g−1 with a high initial Coulombic efficiency of 89% and maintained a capacity of 1969 mAh g−1 at the current density of 500 mA g−1 over 100 cycles. KEYWORDS: Si, chitosan, cross-linking, binder, anode, Li-ion battery



INTRODUCTION In recent years, silicon (Si) has been introduced as an anode material for lithium ion batteries (LIBs) due to its high theoretical capacity (4000 mAh g−1). However, the major obstacle to the practical use of Si anodes in Li-ion batteries is the severe volume changes during the lithium insertion/ extraction process, which consequently leads to the pulverization of the active mass particles and permanent fade of the capacity.1 To overcome this problem, many researchers over the past decade have focused on the design of nanostructured Si with special morphology or the dispersion of Si particles in a lithium active/inactive matrix to restrain the volume change of Si.2 As an alternative solution, recent work has shown that the choice of binder is also very important to stabilize the cycling performance of electrodes for Li-ion batteries.3 Polyvinylidenefluoride (PVDF) has been widely used as a binder in commercial Li-ion batteries. However, PVDF does not provide satisfactory cycling stability when applied to a Si anode due to its linear chain structure.4 Furthermore, PVDF is generally dissolved in volatile, flammable, and explosive Nmethyl-2-pyrrolidone (NMP), which causes serious environmental pollution.5 Therefore, the development of a new ecofriendly, low-cost, and water-soluble binder is essential. Recent studies have shown that the water-soluble and naturederived polymers, such as carboxymethyl cellulose (CMC), alginate, and carboxymethyl chitosan (CMCS) exhibit promising properties as binders for Si-based anodes.6−9 It has been proposed that the hydrogen bonding between the carboxylic groups of these polymers and the Si surface is highly beneficial for reducing the damage induced by volume expansion and thus improving the cycling stability.10,11 Nonetheless, the linear © 2016 American Chemical Society

chain nature of these polymer binders is susceptible to sliding due to the continual volume change of Si during cycling. As a result, the polymer chain together with the deformed electrode cannot recover to its original state.12 For this structural limitation to be overcome, three-dimensional (3D) polymer networks, such as thermally cross-linked CMC-poly(acrylic acid) (PAA)13 and ionic cross-linked alginate binder,8 were applied for silicon anodes to effectively improve the cycle performance of Si anodes by suppressing the adverse effects from the large volume expansion. Chitosan (CS) is a linear polysaccharide composed of randomly distributed β-(1−4)-linked D-glucosamine (deacetylated unit) and N-acetyl-D-glucosamine (acetylated unit). Compared with CMC, which is synthesized by the alkalicatalyzed reaction of cellulose with chloroacetic acid, CS can be easily obtained from shrimp and other crustacean shells. In recent years, CS has been researched as a carbon-coating precursor, binder for cathode, and gel polymer electrolyte for LIBs.14−17 However, only one research group mentioned CS, PAA, amylose, and heparin as binders for Si anodes to compare with CMC until now.18 In our research, CS was systematically investigated as a new water-soluble and cross-linkable biopolymer binder for Si anodes of Li-ion batteries. Si and CS can be interconnected with each other through a hydrogen bond between the amino groups of CS and hydroxyl groups of the Si surface. Furthermore, CS is cross-linked to form a 3D network through the reaction between the amino groups of CS Received: November 5, 2015 Accepted: January 8, 2016 Published: January 8, 2016 2658

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665

Research Article

ACS Applied Materials & Interfaces

Figure 1. (a) FTIR spectra of CS and CS-GA film; (b) FTIR spectra of CS, Si, Si/CS, and Si/CS-GA powder.

and glutaraldehyde (GA),19,20 and then the cross-linked CS-GA limits the large movement of Si particles. Compared with PVDF and linear CS, the cross-linked CS-GA demonstrates much better cycle stability for Si-based anodes.



from 0.03 to 3 V using an electrochemical workstation (WonATech, WBCS3000L). The morphologies of Si nanoparticles and the electrodes were observed with a field emission scanning electron microscope (FE-SEM, JEOL7000F).



RESULTS AND DISCUSSION In the diluted acetic acid aqueous solution, CS underwent chemical cross-linking with GA. The solution was cast into a dish and turned into a film after solvent evaporation. Pure CS film was colorless, whereas the film cross-linked by GA was light yellow in color, which may result from a cross-linking reaction in the context of the Schiff mechanism.19−21 After being dried at 80 °C for 12 h, the CS and CS-GA films were characterized by FTIR (Figure 1a). No peaks of acetic acid were found in the FTIR spectrum, which confirmed complete removal of acetic acid. The chemical structures of CS and crosslinked CS-GA films were also evaluated by FTIR. The broad band located at 3360 cm−1 can be attributed to the hydrogen bonds of -OH and -NH stretching vibrations; the peak at 2870 cm−1 corresponded to the vibration of -CH2-, and the peak at 1566 cm−1 was related to the -NH2 bending vibration. For the cross-linked CS-GA film, a new band at 1647 cm−1 was generated that could be attributed to the formation of a CN bond. The cross-linking was further confirmed by a solubility test. The chitosan film and CS-GA film, which were dried at room temperature, were easily dissolved in the 2 wt % acetic acid solution (Figure S1a and b, see Supporting Information). However, the CS-GA film, which was thermally cross-linked at 80 °C, kept the film form in the solution, indicating that the thermal cross-link reaction was successful (Figure S1c). Once the cross-linking between CS with GA was confirmed, the interaction of Si with CS or cross-linked CS-GA was also characterized by FTIR. As shown in Figure 1b, the IR spectra showed a typical vibration band of Si−O−Si, asymmetric stretching band at 1147 cm−1, and a large vibration absorption band at 666 cm−1, which is a mixture of stretching wagging mode Si−Si and wagging mode Si−Hn (n = 1 and 2).22 Moreover, the broad peak located at 3000−3500 cm−1 was also clearly observed and was attributed to the hydrogen bonding of Si−OH groups.23 All of these characteristic peaks shown in the FTIR spectra support the existence of a SiO2 layer on the Si surface. It is also shown in Figure 1b that, compared with pure CS, the absorption band of NH2 (1584 cm−1) groups for CS shifted slightly to a lower wavenumber (1556 cm−1) for Si/CS and Si/CS-GA, as evidenced by the hydrogen bonding between the amino group of CS and the hydroxyl groups of SiO2 on the surface of the Si particles.9,24

EXPERIMENTAL SECTION

Si powder with an average particle size of ∼100 nm was obtained from Sigma. PVDF was purchased from Dupont Co. (USA). Two types of chitosan (CS) derived from crab shell chitin varying in molecular weight (MW) were used as a binder. The low molecular weight CS (LMW-CS, viscosity: 20−200 cP) was purchased from Dingying Trading Company (China). The high molecular weight CS (HMWCS, viscosity: 200−800 cP) was purchased from Sigma-Aldrich. Glutaraldehyde (GA) was also supplied from Sigma-Aldrich. All materials were used as received. CS (HMW-CS) was dissolved in a 2 wt % acetic acid aqueous solution to form a 3 wt % CS solution. Then, 1−5 wt % GA was injected into 10 mL of the CS solution and stirred for 2 h. The solutions were coded as CS-GA 1, CS-GA 3, and CS-GA 5 according to the amount of GA. The Si powder was mixed with Super P and CS or CS-GA in a weight ratio of 60:20:20 in 2 wt % acetic acid aqueous solution to form a homogeneous slurry. The slurries were spread onto the copper foil by doctor blade and dried at room temperature for 24 h. Then, the samples were further dried in an oven at 80 °C for 12 h to completely remove the solvent. After the heat treatment, chitosan was successful cross-linked by GA. The chemical composition was determined by Fourier transform infrared spectroscopy (FTIR, Bruker tensor 27). The swellability of the binder was studied through electrolyte absorption testing. Dry films were initially weighed (Wbefore) and then immersed in the mixed solvent ethylene carbonate (EC):diethyl carbonate (DC):dimethyl carbonate (DMC) (1:1:1 wt %) at room temperature for 6 h and weighed (Wafter) again after excess electrolyte was wiped from their surfaces. The swellability (S) was calculated as

S=

Wafter − Wbefore × 100% Wbefore

(1)

The tensile properties of the films were measured by a universal test machine (Lloyd/LR30K) using a load cell of 250 N at room temperature. All films were cut into 50 mm × 10 mm rectangles and then immersed in electrolyte for 6 h. The films were clamped onto the grips at a distance of 25 mm and loaded at a constant strain rate of 0.5 mm min−1 until failure. Five measurements were taken for each sample. The tensile stress and Young’s modulus were measured with relative errors of a maximum of 9%. Coin-type test cells were assembled in an argon-filled glovebox using the prepared films as a working electrode, Li foils as both a counter and reference electrode, 1 M solution of LiPF6 in the mixed solvent as an electrolyte, and polypropylene foil as the separator. The 2032-type coin cell was used. These cells were cycled at different rates 2659

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665

Research Article

ACS Applied Materials & Interfaces

Figure 2. (a) Percent of swellability of PVDF, LMW-CS, HMW-CS, and CS- GA. (b) Typical stress−strain curves recorded from PVDF, LMW-CS, HMW-CS, and CS-GA film.

Figure 3. Surface morphology of Si nanoparticles (a) and Si/CS-GA film (b).

modulus of the films increased first and then decreased. The cross-linked CS-GA 3 film showed the highest mechanical performance, and the tensile strength and Young’s modulus were 68.8 MPa and 4.7 GPa, respectively. CS films were mainly stabilized by weak ionic/hydrogen bonds. Covalent crosslinking with GA by inserting a covalent bond between CS strands led to a significant increase in the tensile strength and Young’s modulus of the films. The improved mechanical properties of the CS-GA 3 film were supposed to lead to the formation of a stable 3D network. However, when the GA amount was increased to 5%, high cross-linking density led to the CS-GA 5 film being too brittle, resulting in a decrease in the tensile stress (43.2 MPa). The morphologies of Si and Si/CS-GA anode were observed by FE-SEM. The size of Si particles was in the range of 50 to 150 nm (Figure 3a). In the SEM image of Si/CS-GA, the “bridges” of binder were observed between the adjacent Si particles (the red circles in Figure 3b), indicating that the Si particles were embedded in the cross-linked CS network. Electrochemical properties of the Si electrodes with different binders were evaluated in a half-cell configuration. Generally, high molecular weight polymers possess high mechanical properties, strong adhesion to the substrate, and low swellability in electrolyte solution, so the MW of the binder can greatly affect cycle performance. Therefore, two different MW CSs were used as binders for the cyclability test to optimize the MW of the CS binder. As shown in Figure S2, the Si anode with LMW-CS showed poor capacity retention. As expected, the Si anode with HMW-CS showed much higher

In general, electrolyte solvation of polymeric binders reduces the molecular interaction strength between binders or between the binder and Si particles. Thus, higher electrolyte uptake of the binder can eventually degrade battery performance due to softening of the binder and weak adhesion between the binder and the Si particle.25 The swellability of the films in the electrolyte is shown in Figure 2a. The swellability value of the PVDF film (∼136%) was much higher than that of CS-based films. Among these films, all cross-linked CS-GA films showed the lowest swellability value (∼12%). Presumably, the highly cross-linked CS-GA is likely to form a dense network, which prevents the electrolyte from infiltrating the film. For direct investigation of the mechanical properties of the binders in the presence of the electrolyte medium, the tensile strength of the films was used as an indicator of mechanical functionality after the swellability test. Figure 2b shows the typical stress−strain curves of the PVDF, LMW-CS, HMW-CS, and CS-GA films, which were immersed in the electrolyte for 6 h. As expected, the PVDF film had the lowest tensile strength (26.4 MPa) and Young’s moduli (0.28 GPa). The poor mechanical properties of the PVDF film can be attributed to its high swellability value. All of the chitosan-based films showed much higher tensile strengths and Young’s moduli than that of the PVDF film. The tensile strength and Young’s modulus of the HMW-CS film were significantly higher than those of the LMW-CS film, demonstrating that the mechanical strength of the film increased with an increase in the molecular weight of chitosan. This might be attributable to an entanglement network forming during film formation of HMW-CS.26 For the cross-linked CS films, the tensile strength and Young’s 2660

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665

Research Article

ACS Applied Materials & Interfaces

Figure 4. (a) Comparison of charge/discharge for the Si anode with different binders. (b) Cycle performance of Si anode with different binders at various charge/discharge rates. (c) Discharge capacities of the Si anode with different binders as a function of cycle number. (d) Illustration of the charge/discharge process of the Si/CS and Si/CS-GA anodes. All of the tests were taken between 0.03 and 3.0 V at room temperature.

corresponding to the dealloying process of Li−Si phases. The intensities of both cathodic and anodic peaks were increased with cycle numbers because of activation of crystal Si.28−30 The specific capacities of the Si/PVDF, Si/CS, and Si/CSGA 3 electrodes cycled between 0.03 and 3 V under various current densities are reported in Figure 4b. The Si/PVDF electrode delivered a capacity of ∼3420 mAh g−1 at a rate of 200 mA g−1 for the first cycle; this value quickly decreased to 67 mAh g−1 when the rate was increased to 1000 mA g−1. However, both the Si/HMW-CS and Si/CS-GA electrodes displayed higher reversible capacities of 2330 and 2704 mAhg−1 even at a rate of 2000 mA g−1, respectively. The cycle performance of the Si anode using different binders is shown in Figure 4c. The Si/PVDF electrode showed an initial discharge capacity of 3020 mAhg−1, and then a quick decrease to 114 mAh g−1 after 8 cycles. As a comparison, both the Si/CS and cross-linked Si/CS-GA electrodes showed improved cycling performance. The Si/CS and Si/CS-GA 3 retained 47 and 71% of the initial specific capacitor after 100 cycles at a current density of 500 mA g−1, respectively. The poor cycle stability of Si/PVDF can be attributed to the plastic deformation of PVDF binder in the silicon anodes, showing large volume change during Li ion insertion and extraction. Mechanical disintegration of the silicon electrode during cycling led to the degradation of the electrical conduction network, isolation of silicon nanoparticles, and finally capacity loss.12 For the Si/CS anode, the hydrogen bonding between CS and Si particles can create a self-healing effect for the Si particles, thus contributing to the improved cyclability of the electrode.10 Among all of the samples, the Si/CS-GA 3 anode exhibited the

capacity retention than that of the LMW-CS. Thus, HMW-CS was used as the binder for the whole experiment. The galvanostatic curves are shown in Figure 4a for the Si electrodes with different binders, including PVDF, HMW-CS, and cross-linked CS-GA 3. All of the electrodes showed an obvious charge/discharge plateau at approximately 0.5/0.05 V (vs Li/Li+), which was typically observed for the electrochemical lithiation of Si. No other reduction or oxidation peaks from the side reactions were observed. The initial charge/ discharge capacity of the Si electrode with PVDF, HMW-CS, and CS-GA 3 were 2190/3420, 2380/2904, and 2709/3023 mAhg−1, respectively. The corresponding initial Coulombic efficiencies (ICE) were 64.0, 82.0, and 89.6%, respectively. Compared with Si/PVDF, both Si/HMW-CS and Si/CS-GA 3 exhibited significantly enhanced reversible capacities during the first cycle. The superior ICE of the Si/CS-GA 3 compared to those of the other two electrodes suggests that the cross-linked CS network can decrease the formation of solid electrolyte interface (SEI) layers during the first cycle. Because of the hydrogen bonding of CS with Si, the surface of the Si particle is well-covered with CS-GA, so the decomposition of the electrolyte on the surface could be decreased.10,12,13,25,27 The formation of a network between the amino groups of CS and GA could also limit the movement of the Si, leading to a decrease in isolated Si during the first cycle.27 The CV curves for the first 5 cycles of Si/CS-GA 3 are also shown in Figure S3. For the cathodic processes, the broad peak at 0.19 V and the sharp peak at 0.02 V were attributed to the alloying process of the Li−Si phase. For the anodic processes, there are two broad peaks at 0.40 and 0.36 V, which were 2661

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665

Research Article

ACS Applied Materials & Interfaces

Figure 5. Nyquist plots of Si/PVDF, Si/CS, and Si/CS-GA anode (a) after 3 cycles, (b) after 100 cycles; (c) equivalent circuit model and (d) the plots of Zre vs ω−1/2.

Table 1. Impedance Parameters of Si/PVDF, Si/CS, and Si/CS-GA anode sample

Rs (Ω)

Rct (Ω)

δ (Ω cm2 s−1/2)

D (×10−13 cm2 s−1)

Si/PVDF 3rd Si/CS 3rd Si/CS-GA 3 third Si/PVDF 100th Si/CS 100th Si/CS-GA 3 100th

7.68 13.7 2.81 6.62 16.5 12.5

99 46.3 92.3 1781 405 268

330 212 292 2790 1309 1158

3.25 7.87 4.15 0.04 0.21 0.26

binders, such as PAA,33 PAA-CMC,13 PAA−PVA,12 β-cyclodextrin,10 and alginate.25 The results are listed in Table S1. The performance of cross-linked chitosan has reached the best level among those binders. For gaining further insight into Li-ion storage properties, EIS was conducted to compare the electrical conductivities of Si with PVDF, CS, and cross-linked CS-GA. Panels a and b in Figure 5 show Nyquist plots at the third and 100th discharged states, respectively. The Nyquist plots of Si with PVDF, CS, and cross-linked CS-GA 3 showed two depressed semicircles containing a high frequency semicircle (HFS) and a medium frequency semicircle (MHS) that overlapped with each other and a long low frequency line (LFL), all of which were relative to the SEI resistance, charge transfer resistance, and the Warburg impedance of Li+ diffusion in solid materials, respectively.34 The equivalent circuit was drawn in Figure 5c, where Rs and Rct represent the electrolyte resistance and the charge transfer resistance, W is the Warburg impedance of solid-phase diffusion, respectively.35 The values of Rs and Rct for these samples are reported in Table 1. For further calculating the Li-ion diffusion coefficient (D), the relationship between the real impedance (Zre) and the reciprocal square root of the

best cycle performance, which can be mainly attributed to the formation of a covalent bonding network as illustrated in Figure 4d. The chemical cross-linked CS-GA network effectively restrained any large movement of the silicon nanoparticles and resulted in a lower deformation of the composite electrode on volume change. This interconnected structure allowed a sustained electrical contact between the Si and conducting agent of carbon.7,25,31,32 For further understanding the crosslinking effect on the battery test, the cycling performances of Si/CS anodes with different amounts of GA are also shown in Figure 4c. For the Si/CS anode with 1 wt % of GA, the cycle performance was similar to that of Si/CS. The similar cycle stability implies that the cross-link density is too low to limit the large movement of Si particles. However, when the GA amount reached 5 wt %, the retention of specific capacity quickly decreased to 20.5% after 100 cycles. The decreased cycle performance may be attributed to the fact that high loading of the cross-linking agent made the chitosan film brittle. Ultimately, the Si/CS anode with 3% of GA showed the best cycle performance due to the balance of cross-link density and mechanical properties. The cycle performance of the Si/CS anode with 3% of GA was also compared with those of reported 2662

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665

Research Article

ACS Applied Materials & Interfaces

Figure 6. Surface morphology of Si/PVDF (a), Si/CS (b), and Si/CS-GA 3 (c) anode and surface morphology of Si/PVDF (d), Si/CS (e), and Si/ CS-GA 3 (f) anodes after 100 cycles.

angular frequencies (ω−1/2) is plotted in Figure 5d. Then, Warburg impedance coefficient (δ) can be obtained from the slope of the straight line according to eq 2. Zre = R s + R ct + δω−1/2

no obviously detectable morphology changes were found for the Si/CS and Si/CS-GA 3 electrode (Figure 6e and f). A large number of original Si particles could still be monitored for the Si/CS-GA 3 (Figure 6f), which can be explained as follows: the cross-linked CS-GA network accommodated the large volume change and limited the movement of Si particles, which persevered the original morphology and assisted in constructing a thin and uniform SEI for Si electrodes during the charge/ discharge process. The stable SEI can keep the integrity of the electrode and yield high utilization of Si and stability.7,12,39

(2)

Finally, the diffusion coefficient values of D can be obtained by eq 336 D=

2 1 ⎛ RT ⎞ ⎜ ⎟ 2 ⎝ AF 2C δ ⎠



(3)

CONCLUSIONS A stable cross-linked 3D network structure was successfully formed between Si and CS by the addition of cross-linking agent GA. The cross-linked CS-GA binder exhibited an improvement in its electrochemical performance with a high reversible capacity. The improved electrochemical performance was attributed to the cross-linked network of CS binder and its strong chemical bonding with silicon particles. Considering the facile, ecofriendly, and low cost fabrication, CS as a watersoluble and cross-linkable binder has great potential for use in high capacity silicon anodes in next generation Li-ion batteries and may also be extended to other metal oxide-based electrode materials that undergo large volume change.

where R, T, and F are the gas constant, absolute temperature, and Faraday’s constant, respectively, A is the electrode surface area, and C is the molar concentration of Li ions (mol cm−3). The values of δ and D were also listed in Table 1. After the 100 cycles, the charge transfer resistance of Si/PVDF was increased from 99 to 1781 Ω, and the diffusion coefficient was decreased from 3.25204 × 10−13 to 4.54962 × 10−15 cm2 s−1.The increased Rct and decreased D can be attributed to the aggregation of Si particles, which made the Si detach from substrate and hindered the diffusion of lithium ions. The Si/CSGA 3 electrode showed the lowest Rct (268 Ω) and highest D (2.64099 × 10−14 cm2 s−1) after 100 cycles among all of the anodes, demonstrating that the cross-linked CS-GA was more stable in the discharge/charge process, which was also confirmed by the cycle test. For better understanding of the improved electrochemical performance achieved by using the cross-linked CS binder, the Si anodes with different binders after the 100th cycle were characterized by FE-SEM. Panels a−c in Figure 6 show the morphology of pristine electrodes with PVDF, CS, and CS-GA binders, respectively. Si nanoparticles together with Super P were relatively uniformly dispersed in the matrix in the three samples before the cycle test. After 100 cycles, the topography of the cycled electrode using PVDF binder show a smooth surface, which can be attributed to the pulverization of Si and continues growth of SEI on newly exposed surfaces.12 Furthermore, huge crack, pulverization, and aggregation of the Si nanoparticles were also observed in Figure 6d due to the mechanical stress in the electrode during the volume expansion of the Si particles.37,38 Compared with the Si/PVDF electrode,



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.5b10673. Solubility of CS and CS-GA films; cycling performance of the Si anode with different molecular weight CS binders; CV curves of Si/CS-GA 2.5 anode; and cycle performances of Si anodes used with different kinds of binders (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +82-31-290-7248. Fax: +82-31290-7373. 2663

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665

Research Article

ACS Applied Materials & Interfaces Notes

(16) Prasanna, K.; Subburaj, T.; Jo, Y. N.; Lee, W. J.; Lee, C. W. Environment-Friendly Cathodes Using Biopolymer Chitosan with Enhanced Electrochemical Behavior for Use in Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 7884−7890. (17) Chupp, J.; Shellikeri, A.; Palui, G.; Chatterjee, J. Chitosan-based Gel Film Electrolytes Containing Ionic Liquid and Lithium Salt for Energy Storage Applications. J. Appl. Polym. Sci. 2015, 132, 8. (18) Bridel, J. S.; Azais, T.; Morcrette, M.; Tarascon, J. M.; Larcher, D. Key Parameters Governing the Reversibility of Si/Carbon/CMC Electrodes for Li-ion Batteries. Chem. Mater. 2010, 22, 1229−1241. (19) Rodriguez-Velazquez, E.; Silva, M.; Taboada, P.; Mano, J. F.; Suarez-Quintanilla, D.; Alatorre-Meda, M. Enhanced Cell Affinity of Chitosan Membranes Mediated by Superficial Cross-linking: A Straightforward Method Attainable by Standard Laboratory Procedures. Biomacromolecules 2014, 15, 291−301. (20) Schiffman, J. D.; Schauer, C. L. Cross-linking Chitosan Nanofibers. Biomacromolecules 2007, 8, 594−601. (21) Ma, J.; Sahai, Y. A Direct Borohydride Fuel Cell with Thin Film Anode and Polymer Hydrogel Membrane. ECS Electrochem. Lett. 2012, 1, F41−F43. (22) Rahmani, M.; Moadhen, A.; Zaïbi, M. A.; Elhouichet, H.; Oueslati, M. Photoluminescence Enhancement and Stabilisation of Porous Silicon Passivated by Iron. J. Lumin. 2008, 128, 1763−1766. (23) Kovalenko, I.; Zdyrko, B.; Magasinski, A.; Hertzberg, B.; Milicev, Z.; Burtovyy, R.; Luzinov, I.; Yushin, G. A Major Constituent of Brown Algae for Use in High-capacity Li-ion Batteries. Science 2011, 334, 75−79. (24) Laudenslager, M. J.; Schiffman, J. D.; Schauer, C. L. Carboxymethyl Chitosan as a Matrix Material for Platinum, Gold, and Silver Nanoparticles. Biomacromolecules 2008, 9, 2682−2685. (25) Yoon, J.; Oh, D. X.; Jo, C.; Lee, J.; Hwang, D. S. Improvement of Desolvation and Resilience of Alginate Binders for Si-based Anodes in a Lithium Ion Battery by Calcium-mediated Cross-linking. Phys. Chem. Chem. Phys. 2014, 16, 25628−25635. (26) Nunthanid, J.; Puttipipatkhachorn, S.; Yamamoto, K.; Peck, G. E. Physical Properties and Molecular Behavior of Chitosan Films. Drug Dev. Ind. Pharm. 2001, 27, 143−157. (27) Liu, J.; Zhang, Q.; Zhang, T.; Li, J. T.; Huang, L.; Sun, S. G. A Robust Ion-conductive Biopolymer as a Binder for Si Anodes of Lithium-ion Batteries. Adv. Funct. Mater. 2015, 25, 3599−3605. (28) Ge, M.; Rong, J.; Fang, X.; Zhou, C. Porous Doped Silicon Nanowires for Lithium Ion Battery Anode with Long Cycle Life. Nano Lett. 2012, 12, 2318−2323. (29) Hassan, F. M.; Chabot, V.; Elsayed, A. R.; Xiao, X.; Chen, Z. Engineered Si Electrode Nanoarchitecture: a Scalable Postfabrication Treatment for the Production of Next-generation Li-ion Batteries. Nano Lett. 2014, 14, 277−283. (30) Yang, Y.; Chen, D. Q.; Liu, B.; Zhao, J. B. Binder-free Si Nanoparticle Electrode with 3D Porous Structure Prepared by Electrophoretic Deposition for Lithium-ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 7497−7504. (31) Chou, S. L.; Pan, Y. D.; Wang, J. Z.; Liu, H. K.; Dou, S. X. Small Things Make a Big Difference: Binder Effects on the Performance of Li and Na Batteries. Phys. Chem. Chem. Phys. 2014, 16, 20347−20359. (32) Yim, T.; Choi, S. J.; Park, J. H.; Cho, W.; Jo, Y. N.; Kim, T. H.; Kim, Y. J. The Effect of an Elastic Functional Group in a Rigid Binder Framework of Silicon-graphite Composites on Their Electrochemical Performance. Phys. Chem. Chem. Phys. 2015, 17, 2388−2393. (33) Park, H. K.; Kong, B. S.; Oh, E. S. Effect of High Adhesive Polyvinyl Alcohol Binder on the Anodes of Lithium Ion Batteries. Electrochem. Commun. 2011, 13, 1051−1053. (34) Xiang, J. Y.; Tu, J. P.; Qiao, Y. Q.; Wang, X. L.; Zhong, J.; Zhang, D.; Gu, C. D. Electrochemical Impedance Analysis of a Hierarchical CuO Electrode Composed of Self-Assembled Nanoplates. J. Phys. Chem. C 2011, 115, 2505−2513. (35) Aurbach, D. Review of Selected Electrode−solution Interactions Which Determine the Performance of Li and Li Ion Batteries. J. Power Sources 2000, 89, 206−218.

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by Energy Efficiency & Resources of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korean government Ministry of Knowledge Economy (No. 20142010102690) and the Basic Science Research Program through the National Research Foundation of Korea Grant funded by the Ministry of Science, ICT & Future Planning (2009-0083540).



REFERENCES

(1) Su, X.; Wu, Q.; Li, J.; Xiao, X.; Lott, A.; Lu, W.; Sheldon, B. W.; Wu, J. Silicon-based Nanomaterials for Lithium-ion Batteries: A Review. Adv. Energy. Mater. 2014, 4, 1−23. (2) Jiang, J.; Li, Y. Y.; Liu, J. P.; Huang, X. T.; Yuan, C. Z.; Lou, X. W. Recent Advances in Metal Oxide-based Electrode Architecture Design for Electrochemical Energy Storage. Adv. Mater. 2012, 24, 5166−5180. (3) Magasinski, A.; Zdyrko, B.; Kovalenko, I.; Hertzberg, B.; Burtovyy, R.; Huebner, C. F.; Fuller, T. F.; Luzinov, I.; Yushin, G. Toward Efficient Binders for Li-ion Battery Si-based Anodes: Polyacrylic Acid. ACS Appl. Mater. Interfaces 2010, 2, 3004−3010. (4) Komaba, S.; Yabuuchi, N.; Ozeki, T.; Han, Z. J.; Shimomura, K.; Yui, H.; Katayama, Y.; Miura, T. Comparative Study of Sodium Polyacrylate and Poly(vinylidene fluoride) as Binders for High Capacity Si-graphite Composite Negative Electrodes in Li-ion Batteries. J. Phys. Chem. C 2012, 116, 1380−1389. (5) Chai, L. L.; Qu, Q. T.; Zhang, L. F.; Shen, M.; Zhang, L.; Zheng, H. H. Chitosan, a New and Environmental Benign Electrode Binder for Use with Graphite Anode in Lithium-ion Batteries. Electrochim. Acta 2013, 105, 378−383. (6) Bridel, J. S.; Azais, T.; Morcrette, M.; Tarascon, J. M.; Larcher, D. In Situ Observation and Long-term Reactivity of Si/C/CMC Composites Electrodes for Li-ion Batteries. J. Electrochem. Soc. 2011, 158, A750−A759. (7) Bridel, J.-S.; Azais, T.; Morcrette, M.; Tarascon, J.-M.; Larcher, D. Key Parameters Governing the Reversibility of Si/Carbon/CMC Electrodes for Li-ion Batteries. Chem. Mater. 2009, 22, 1229−1241. (8) Liu, J.; Zhang, Q.; Wu, Z. Y.; Wu, J. H.; Li, J. T.; Huang, L.; Sun, S. G. A High-performance Alginate Hydrogel Binder for the Si/C Anode of a Li-ion Battery. Chem. Commun. 2014, 50, 6386−6389. (9) Yue, L.; Zhang, L.; Zhong, H. Carboxymethyl Chitosan: A New Water-soluble Binder for Si Anode of Li-ion Batteries. J. Power Sources 2014, 247, 327−331. (10) Jeong, Y. K.; Kwon, T. W.; Lee, I.; Kim, T. S.; Coskun, A.; Choi, J. W. Hyperbranched Beta-cyclodextrin Polymer as an Effective Multidimensional Binder for Silicon Anodes in Lithium Rechargeable Batteries. Nano Lett. 2014, 14, 864−870. (11) Vogl, U. S.; Das, P. K.; Weber, A. Z.; Winter, M.; Kostecki, R. M.; Lux, S. F. The Mechanism of Interactions between CMC Binder and Si Single Crystal Facets. Langmuir 2014, 30 (34), 10299−10307. (12) Song, J. X.; Zhou, M. J.; Yi, R.; Xu, T.; Gordin, M. L.; Tang, D. H.; Yu, Z. X.; Regula, M.; Wang, D. H. Interpenetrated Gel Polymer Binder for High-performance Silicon Anodes in Lithium-ion Batteries. Adv. Funct. Mater. 2014, 24, 5904−5910. (13) Koo, B.; Kim, H.; Cho, Y.; Lee, K. T.; Choi, N.-S.; Cho, J. A Highly Cross-linked Polymeric Binder for High-performance Silicon Negative Electrodes in Lithium Ion Batteries. Angew. Chem., Int. Ed. 2012, 51, 8762−8767. (14) Kim, K. W.; Kim, J. S.; Lee, S. W.; Lee, J. K. Employment of Chitosan-linked Iron Oxides as Mesoporous Anode Materials for Improved Lithium-ion Batteries. Electrochim. Acta 2015, 170, 146− 153. (15) Liu, J.; Du, C. Q.; Tang, Z. Y.; Yang, M.; Zhang, X. H. In Situ Nickel/Carbon Coated Lithium Titanium Oxide Anode Material with Improved Electrochemical Properties. Electrochim. Acta 2014, 143, 56−62. 2664

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665

Research Article

ACS Applied Materials & Interfaces (36) Ding, N.; Xu, J.; Yao, Y. X.; Wegner, G.; Fang, X.; Chen, C. H.; Lieberwirth, I. Determination of the Diffusion Coefficient of Lithium Ions in Nano-Si. Solid State Ionics 2009, 180, 222−225. (37) Vogl, U. S.; Das, P. K.; Weber, A. Z.; Winter, M.; Kostecki, R.; Lux, S. F. Mechanism of Interactions between CMC Binder and Si Single Crystal Facets. Langmuir 2014, 30, 10299−10307. (38) Liu, N.; Lu, Z. D.; Zhao, J.; McDowell, M. T.; Lee, H. W.; Zhao, W. T.; Cui, Y. A Pomegranate-inspired Nanoscale Design for Largevolume-change Lithium Battery Anodes. Nat. Nanotechnol. 2014, 9, 187−192. (39) Ming, J.; Ming, H.; Kwak, W.-J.; Shin, C.; Zheng, J.; Sun, Y.-K. The Binder Effect on An Oxide-based Anode in Lithium and Sodiumion Battery Applications: The Fastest Way to Ultrahigh Performance. Chem. Commun. 2014, 50, 13307−13310.

2665

DOI: 10.1021/acsami.5b10673 ACS Appl. Mater. Interfaces 2016, 8, 2658−2665