Guiying
Tian‡
a,
Zijian
Zhao‡
a,
Angelina
Sarapulova
a,
Chittaranjan
Das
a,
Lihua
Zhu
a,
Suya
Liu
bc,
Aleksandr
Missiul
d,
Edmund
Welter
e,
Julia
Maibach
a and
Sonia
Dsoke
*af
aInstitute for Applied Materials (IAM), Karlsruhe Institute of Technology (KIT), Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany. E-mail: sonia.dsoke@kit.edu
bInternational Center for New-Structured Materials (ICNSM), Zhejiang University (ZJU), Zheda Rd 38, Hangzhou, 310027, P. R. China
cInstitute of Nanotechnology (INT), Karlsruhe Institute of Technology (KIT), Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany
dCELLS-ALBA, Carrer de la Llum 2-26, 08290 Cerdanyola del Vallès, Barcelona, Spain
eDeutsches Elektronen-Synchrotron DESY, Notkestrasse 85, D-22607 Hamburg, Germany
fHelmholtz-Institute Ulm for Electrochemical Energy Storage (HIU), Helmholtzstrasse 11, 89081 Ulm, Germany
First published on 6th April 2019
Sulfide compounds are interesting conversion electrode materials for Li-ion batteries, due to their high theoretical capacity. However, they suffer from large volumetric changes and fast capacity fading. To overcome these issues, nanosized zinc sulfide (ZnS) modified with polyelectrolytes and graphene (ZnS-C/G) has been synthesized and investigated as an enhanced conversion-alloying anode material. In situ synchrotron X-ray diffraction and X-ray absorption spectroscopy are used to elucidate the Li storage process during the 1st cycle. In addition, the evolution of internal resistance and the corresponding solid electrolyte interphase (SEI) formation during the 1st cycle are discussed based on electrochemical impedance spectroscopy and X-ray photoelectron spectroscopy. The results reveal that the formation of lithiated products and the SEI layer at different voltages can influence Li+ diffusion into the electrode. Moreover, an artificial carbon layer can not only facilitate Li+ transport but also avoid the direct formation of the SEI layer on the surface of active particles. Compared to bare ZnS, the ZnS-C/G electrode shows outstanding rate capability and cycling capacity (571 mA h g−1 after 120 cycles at a specific current of 1.0 A g−1 with a retention rate of 94.4%). The high capacity at elevated current density is ascribed to the contribution of capacitive charge storage.
In general, sulfur atoms possess a large atomic radius and they are easily polarized.6 Metal–sulfur bonds in sulfide conversion materials are more easily broken/formed during the electrochemical Li+ insertion/deinsertion compared to metal–oxygen bonds. Thus, they effectively enhance the reaction kinetics and decrease the overpotential for lithium storage.7 In the conversion reaction, the transition metal can be fully and reversibly reduced to its metallic state, leading to high theoretical capacity.7–11 Thus, many sulfide conversion materials (e.g., FeS2, NiS, CoS2, MnS, MoS2, and ZnS) have been studied as anode materials for LIBs.12–17 Differing from other transition-metals, lithium storage in the ZnS electrode relies not only on the redox conversion mechanism but also on the alloying mechanism. The formation of a Zn–Li alloy leads to a higher theoretical capacity (824.9 mA h g−1), compared to the one only based on the conversion mechanism (549.9 mA g−1). In addition, since ZnS is the main form of zinc which can be found in the nature (as sphalerite and wurtzite), it can be used as low-cost electrode material.
Based on the conversion mechanism, the lithiation process leads to the formation of nanosized metal grains embedded in amorphous Li2S. Due to the large amount of Li+ ions being inserted/deinserted, conversion materials undergo significant volumetric expansion, leading to pulverization during cycling. The reconstructive phase transition can also cause dissolution of polysulfide intermediates in an organic electrolyte, when sulfide anodes are directly exposed to the electrolyte.17 Thus, the problems of poor cycling stability and rapid capacity fading are unavoidable for sulfide anodes. At the same time, the low electronic conductivity of ZnS severely restricts the electrochemical kinetics. To solve these issues, carbon modification was proposed to enhance the bulk conductivity and cycling stability.7,13 Most importantly, carbon modification can further improve the structural stability by capturing/adsorbing polysulfide intermediates.18,19 However, there is still a need to elucidate the structural evolution and electrochemical kinetics of sulfide-type conversion materials with a carbon coating during the cycling process.
Currently, many efforts are focused on the development of conversion materials with novel synthesis techniques to create unique morphologies and achieve excellent performance.15,20–25 However, only a few studies have provided an understanding of the detailed reaction mechanism, such as phase transitions and resistance changes. Thus, the crystal change and SEI formation during the electrochemical reaction need to be further investigated.26 In this study, ZnS nanoparticles modified with a carbon coating derived from polyelectrolytes and graphene (ZnS-C/G) are studied as a typical conversion-alloying electrode material. The electrochemical performance, evolution of the crystal phase and simultaneous SEI formation are analysed systematically using a series of in situ and ex situ methods (synchrotron radiation diffraction (SRD), X-ray absorption spectroscopy (XAS), X-ray photoelectron spectroscopy (XPS) and electrochemical impedance spectroscopy (EIS)).
A three-electrode EL-CELL® was built for EIS tests with a working electrode, a thin glass fiber separator, a lithium foil counter electrode, LP30 electrolyte, and a Li-ring reference electrode placed exactly around the working and counter electrodes. This cell was cycled with potential steps of ±0.185 V in the voltage range of open-circuit voltage (OCV) → 0.01 V → 3.00 V. At each potential, the cell was held for 3 h for charge equilibrium followed by the EIS test using an alternating current signal of a small amplitude of 5 mV over a frequency range from 500 kHz to 10 mHz.
In situ synchrotron radiation diffraction (SRD) data were collected using synchrotron radiation (λ = 0.41266(2) Å) at the materials science and powder diffraction beamline (MSPD), ALBA, Barcelona (see Fig. S1†). During the in situ measurements, the cells were cycled at 80 mA g−1 in the potential range of 0.01–3.0 V vs. Li/Li+ at room temperature, and Si and LaB6 were used as standard samples for calibration. In situ XAS measurements were performed at the P65 beamline (photon flux > 1011 to 1012 s−1) at PETRA III in DESY, Hamburg (see Fig. S2†).
In situ XAS spectra were recorded in transmission geometry with the conventional step-scan mode at the Zn K-edge (9659 eV). The double-crystal fixed exit monochromator equipped with a Si (111) crystal and the ionization chamber were optimized for the Zn edge. The cells were cycled at 150 mA g−1 in the potential range from OCV to 0.01 V vs. Li/Li+ at room temperature using a VMP3 potentiostat (Bio-Logic), which was taken from the P02.1 beam line (PETRA III, DESY, Hamburg). The spectra were processed using the Demeter software package.29
HRTEM images are collected to further confirm the carbon coated ZnS-C/G composite. The HRTEM image in Fig. 2a shows that nanosized ZnS particles are homogeneously covered by an ∼2 nm carbon layer. The inserted image provides a detailed view of crystal lattice fringes of ZnS with a spacing of 0.307 nm, which corresponds to the (111) lattice planes. Furthermore, the phase purity of the ZnS-C/G composite is evaluated based on the XRD pattern using Rietveld refinement of the structure model. In Fig. 2b, the refinement confirms that the main reflections are well indexed to the ZnS cubic phase (space group: F3m). Satisfactory agreement factors (Rwp = 4.78% and Rp = 6.20%) with the cell parameters (a = b = c = 5.399(2) Å and cell volume = 157.4(1) Å3) are determined in this crystal structure refinement.
![]() | ||
Fig. 2 (a) HRTEM images of the ZnS-C/G composite; (b) Rietveld refinement of the structure model, based on the XRD pattern of the ZnS-C/G composite. |
As discussed above, carbon coating is an effective strategy to improve the bulk conductivity of conversion-type electrodes. To determine the degree of graphitization of the coated carbon, Raman measurements of ZnS and the ZnS-C/G were carried out. As presented in Fig. S3a,† ZnS displays a series of modes at 259.4, 347.7 and 700.4 cm−1, corresponding to the longitudinal optical (LO) phonon–plasmon coupled L− mode, the LO phonon mode associated with the Brillouin zone center, and the E1(LO) mode at Γ.33–35 Together with the XRD refined result, Raman analysis confirms that ZnS is a single cubic phase. In contrast, the typical ZnS modes are not observed in the ZnS-C/G, due to the uniform carbon coating. The intensity ratio of G (graphitic) and D (defective) bands (IG/ID) can be used to determine the graphitization level of coated carbon. A higher IG/ID ratio corresponds to a higher degree of graphitization. Here, the IG/ID ratio of ∼0.62 for the ZnS-C/G is higher than the threshold (0.52) for electron-conductive carbon,36,37 indicating that the surface of the ZnS nanoparticles is wrapped with a conductive carbon layer.
The CV scans of the carbon-coated ZnS-C/G material show in principle the same anodic and cathodic peaks as the uncoated material. However, the fade of the redox peaks is much less, indicating improved structural and electrochemical stability of the ZnS-C/G electrode (see Fig. 3b). In addition, the polarization of the ZnS-C/G electrode is reduced to 0.56 V compared to that of bare ZnS (0.63 V), which is ascribed to the improvement of the bulk conductivity of the ZnS-C/G electrode via the carbon coating. The improved reversibility of the ZnS-C/G electrode can also be viewed on the consecutive galvanostatic cycling (Fig. 3d). The irreversible capacity at the 1st cycle (213 mA h g−1) is much lower than that of ZnS (910 mA h g−1), indicating that fewer Li+ ions are consumed in this irreversible SEI formation.43 This can indicate that the carbon coating acts as a protection shell on the ZnS particles preventing SEI formation directly on the ZnS particles. In addition, the ZnS-C/G electrode maintains high capacity during the subsequent cycles (800 mA h g−1 at the 2nd cycle and 780 mA h g−1 at the 10th cycle). These electrochemical results suggest that it is promising to achieve long-term reversible cycling with the ZnS-C/G material.
In order to further determine the mechanism of the electrochemical reaction, in situ SRD analysis was performed. Fig. 4a shows the initial galvanostatic cycle of the ZnS-C/G electrode vs. lithium foil in a voltage range of 0.01–3.0 V. The corresponding SRD patterns (Fig. S4a and b†), consecutively recorded with 133 scans, provide an overview of the structural evolution. To clarify the multiple steps during (de)lithiation processes, the procedure is divided into four steps (steps I–IV).
In step I (scan 1–scan 84, 2.37–0.21 V), Fig. 4b and S5a† show that the ordered cubic structure of ZnS is still maintained at the beginning of the lithiation process, and Rietveld refinement determines the cubic structure of scan 20 at 0.86 V (space group: F3m, see Fig. S5b†); and further lithiation can reduce the content of the cubic structure, as confirmed by the gradually decreased intensity corresponding to the ZnS cubic phase (marked with green arrows). At the same time, a gently sloping voltage profile around 0.4 V can be observed in step I, confirming that a large number of Li+ ions are consumed or stored here, which is due to the initial conversion reaction, SEI formation and pseudocapacitive storage. Therefore, no new phase related to conversion products (i.e., reduced Zn and Li2S) appears here. This is also possible because the particle size of the formed Zn grains is below a few nanometers, and the amorphous Li2S cannot be seen in the SRD pattern either.
In step II (scan 85–scan 104, 0.21–0.01 V), new reflections at about 11.1°, 15.7° and 19.4° (marked with red arrows) start to appear since scan 85 (see Fig. 4b and S5c†). The previously formed Zn starts to alloy with Li+ ions, forming a LixZn alloy with a low x value (x ≈ 0). A refined analysis of scan 110 confirms the trigonal crystal structure (space group: Pm1) of Li0.62Zn0.87 (Li0.71Zn), indicating further lithiation and formation of the LixZn alloy with increasing x value (see Fig. S5d†). The higher lithiation level leads to a gradual shift of the reflections towards lower 2θ angles, because inserted Li+ ions expand the unit cell of LixZn, which was also observed in the lithiation process of ZnO.40 When delithiation is performed from scan 105 to scan 120, the reflections shift back to higher 2θ angles (step III, 0.01–0.89 V), and this phenomenon corresponds to the dealloying process.
In step IV (scan 121–scan 133, 0.89–3.00 V), Fig. 4b shows that a new phase appears (marked with black arrows). This is related to the formation of a LiyZn alloy with a low y value during the dealloying process.40 The SRD patterns of scan 121 are compared with references (Fig. S5e, f and S6†). This comparison proves the formation of Li0.25Zn, indicating that the y in the LiyZn alloy becomes less again. Corresponding to the last steep increase of potential, the reflections of hexagonal ZnS appear (space group: P63mc). At the same time, the LiyZn-related reflections completely vanish (see Fig. S5g†).
To further determine the evolution of the zinc species in the ZnS-C/G electrode during the 1st lithiation, in situ XAS at the Zn K-edge (9659 eV) are measured during a galvanostatic cycle performed at 150 mA g−1 from OCV to 0.05 V vs. Li/Li+ (Fig. 5a). Fig. 5b shows that the Zn K-edge of the pristine ZnS-C/G electrode systematically shifts to lower energies as the lithiation level increases, confirming the conversion reaction Zn2+ → Zn0. However, the mismatch of scan 12 and the Zn-foil reference suggests that the conversion reaction to metallic Zn does not completely finish. Zn K-edge EXAFS Fourier transform (FT) curves are plotted in Fig. 5c. The pristine ZnS-C/G shows one distinct peak at ∼1.84 Å and one broad peak at 3.4–4.1 Å, corresponding to the 1st Zn coordination shell of 4 S and overlapping of the 2nd and 3rd shells with 12 Zn and 9 S, respectively.44 As more Li+ ions are inserted and the voltage decreases to ∼0.15 V (scan 9), the three peaks become weaker and weaker, and shift to higher distances. This corresponds to step I in the in situ SRD analysis, where ZnS gradually deforms as more Li+ are inserted. The estimation of the oxidation state of Zn in the ZnS-C/G (according to reference spectra of Zn metal foil and pristine ZnS) is achieved through linear combination fitting of the Zn K-edge XANES spectra (see Table S1 and Fig. S7†). The result proves that not all Zn is involved in the conversion reaction due to the fast lithiation, and 31.6 wt% of residual ZnS still remains unreacted in the ZnS-C/G electrode (see Fig. 5d). Previous reports attempted to elucidate the electrochemical mechanism of Li-insertion in ZnS. He et al.24 for the first time used ex situ XRD patterns to confirm the formation of Zn + LixZn (1st lithiation at 0.05 V), and ZnS (1st delithiation at 2.0 V). Park et al.45 applied ex situ XRD and XAS to determine the formation of LiZn4 (1st lithiation at 0.4 V), LiZn (1st lithiation at 0.0 V) and ZnS (1st delithiation at 2.0 V). As we know, in situ technologies have the advantage to avoid the influence of external factors affecting the sensitive cycled electrolyte, and also to provide consecutive information during the (de)lithiation process. In this study, it is the first time that the (de)lithiation evolution of the ZnS electrode during the 1st cycle is elucidated in detail by in situ analysis (XRD and XAS). In situ XRD clearly shows the formed intermediate alloying phases: there is no evidence of the pure Zn phase, but of LixZn with an extremely low x value. The fitting of in situ XAS confirms the Zn generation (68.4% of total Zn source in the electrode). After the 1st cycle, hexagonal ZnS is regenerated, which is different from the initial cubic ZnS phase. The multiple processes (conversion and alloying reactions) can be described with the following reactions:
Step I (lithiation conversion, OCV ∼ 0.21 V):
cubic ZnS + 2Li+ + 2e− → Li2S + Zn |
Step II (lithiation alloying, 0.21–0.01 V):
Zn + xLi+ + xe− → trigonal LixZn |
Step III (delithiation dealloying, 0.01–0.89 V):
LixZn → LiyZn + (x − y)e− + (x − y)Li+ (x > y) |
LiyZn → Zn + ye− + yLi+ |
Step IV (delithiation conversion, 0.89–3.00 V):
Li2S + Zn → hexagonal ZnS + 2e− + 2Li+ |
DLi = 0.5 × (RT/Az2F2CLiσ)2 | (1) |
In general, the variation of impedance and Li+ diffusion with the potential depends on interfacial changes induced by the SEI formation and phase transition caused by the redox reactions.47 To clarify this, an equivalent circuit (Fig. S8b†) has been used to fit these EIS spectra. In this equivalent circuit, Rel accounts for the contribution of the liquid electrolyte, separator and internal connections of the cell, corresponding to the intercept with the real axis at high-frequency. At high frequency, the interface resistance (Rinter) mainly consists of SEI resistance (RSEI) and contact resistance (Rcon), combined with the related double layer capacitance (Cdl1 and Cdl2). At low frequency, charge transfer resistance (Rct) combined with the related double layer capacitance (Cdl3) corresponds to the third semicircle. Also in the low frequency region, it is possible to recognize a 45° inclined line as Warburg diffusion related to Li+ diffusion into the electrode. As summarized in Fig. 7a, there is no obvious change in the Rel during the 1st cycle, which means that the liquid electrolyte is electrochemically stable.
In detail, Fig. 6a displays the spectra related to step A (related to the voltage range of 2.5 V → 1.2 V). In step A, Rct is extremely high since no lithiation reaction (no charge transfer) is observed above 1.2 V, as confirmed by the 1st CV curve in Fig. 3b. At the same time, a relatively high Rinter can be obtained in Fig. 7b due to the fresh status of the electrode. Thereafter, in step B, the semicircle size of the Rct drops rapidly from 1.2 V to 0.1 V (see Fig. 6b), and this corresponds to the cathodic peaks of the 1st CV curve at ∼0.75, ∼0.5 and ∼0.1 V (see Fig. 3b), meaning the occurrence of Li+ ion insertion into the electrode. A decrease in the Rinter can be ascribed to the SEI formation, because a thin SEI layer can facilitate Li+ to go through the interface of the active particles. Hence, the Rinter and Rct are further combined into one semicircle. The increase in Rct after lithiation at ∼0.1 V suggests that the lithiation process in the ZnS bulk finishes; the Rcon and RSEI are reduced due to the accumulation of metallic Zn and the appearance of the LixZn alloy which can increase the electrode conductivity. In step C (see Fig. 6c), the Rct decreases again at around 1.2 V due to the occurrence of delithiation.
In step D, the increase in Rinter could be ascribed to the fact that the pulverized nanograins are surrounded by the SEI layer, leading to a high resistance above 2.5 V (Fig. 6d). The increasing Rct suggests that the delithiation process ends at 1.8 V. In sum, a relatively low Rct is obtained at 0.8–0.2 V in the lithiation process and at 0.9–1.6 V in the delithiation process, in accordance with the main conversion and (de)alloying processes. Furthermore, the change of DLi can be clearly observed in Fig. 7b, where a relatively high DLi is observed from 1.0 V to 0.01 V in the lithiation process and below 1.5 V in the delithiation process. In step D, Li ions diffuse slowly again because of the reconstruction of the nonconductive ZnS. Basically, the occurrence of (de)lithiation in certain potential ranges can directly influence the variation of the Rct and the DLi.
Because the Li storage in the anode material is normally accompanied by SEI formation,42,48 XPS analysis was performed on the ZnS-C/G electrodes to uncover the SEI components involved in redox reactions (comparing a pristine electrode and lithiated electrodes at 1.0 V and 0.01 V during the 1st lithiation process). The pristine electrode and the lithiated electrode at 1.0 V show similar spectra of Zn 2p peaks at 1022.1 and 1045.3 eV, corresponding to Zn 2p3/2 and Zn 2p1/2, respectively (Fig. 8a).49 No reduction of Zn2+ → Zn0 can be observed in the XPS data above 1.0 V, while these typical Zn peaks disappear on the lithiated electrode at 0.01 V. This indicates that the surface of the active material (ZnS) is fully covered by an SEI layer when the potential is further reduced from 1.0 V to 0.01 V. This is also confirmed by the S 2p1/2 and 2p3/2 peaks (162.6 eV and 161.4 eV) observed only at the pristine and 1.0 V lithiated electrodes, corresponding to the Zn–S bonds (Fig. 8b). In contrast to Zn 2p, the S 2p still remains visible for the lithiated electrode at 0.01 V. The sulfur signal may come either from the active material underneath the SEI layer as it has higher kinetic energy compared to Zn 2p or from sulfur compounds in the SEI layer. On considering the shift due to the dipole of the SEI layer,50 the S 2p peak is found to be at the lower binding energy of 160.5 eV which indicates the formation of Li2S underneath the SEI layer during lithiation.51 However, the limited probing depth of 10 nm in XPS measurement restricts a further comment on the formation of LixZn. Compared to the Li-free pristine ZnS electrode, the peaks of Li 1s appearing at 56.0 eV for the lithiated electrode at 1.0 V and at 55.2 eV for the lithiated electrode at 0.01 V can be assigned to LiF and Li-carbonates, respectively (Fig. 8c).52
![]() | ||
Fig. 8 XPS spectra of ZnS-C/G electrodes at pristine, lithiated 1.0 V and 0.01 V states: Zn 2p (a), S 2p (b), Li 1s (c), C 1s (d), O 1s (e) and F 1s (f). |
In the carbon spectra in Fig. 8d, the high signal intensity of graphitic carbon (at 284.4 eV) is ascribed to the added carbon (i.e., Super P and graphene). The active carbon and PVDF binder can introduce some oxygen and fluorine containing groups as observed at higher binding energies in the spectrum of the pristine electrode. As lithiation proceeds, the peak intensity in the C 1s, corresponding to hydrocarbon at 285.3 eV, ester group at 286.7 eV and carbonate at 289.5/290.4 eV, gradually increases, which shows the presence of a mixture of PEO-type oligomers (–CH2–CH2–O–)n, and ROCO2Li/Li2CO3 in the SEI layer.53 The clearly reduced intensity for graphitic carbon and the absence of the Zn 2p signal at 0.01 V show that the surface of the ZnS-C/G electrode is covered with an SEI layer with a thickness of more than 5 nm, corresponding to the information depth for C 1s using Al Kα radiation. In O 1s spectra (Fig. 8e), peaks at 532.1 and 534.4 eV in both the lithiated samples are observed, which correspond to the formation of LiOH (or Li2CO3) and ROCO2Li induced by electrolyte decomposition, respectively.53–55 In addition, the O 1s spectrum of the highly lithiated sample at 0.01 V shows the formation of Li2O (at 528.2 eV) due to the further decomposition of LiOH or Li2CO3 at a low potential.56 As shown in Fig. 8f in the F 1s spectra, a peak at 687.8 eV from PVDF binder appears for all three electrodes. However, a peak at lower binding energy appears at 685.2 eV for the electrode lithiated at 1.0 V and becomes more intensive for the electrode lithiated at 0.01 V. The appearance of the lower binding energy peak in the F 1s spectra of the lithiated samples confirms the formation of LiF at the surface, which comes from the side reaction of LiPF6 (LiPF6 → LiF + PF5 or LiPF6 + xe− + xLi+ → (1 + x)LiF + LiPF6−x).57 From the XPS analysis, one can observe the formation of the SEI layer above the lithiation potential at 1.0 V, which is composed of mainly organic materials (PEO-type oligomers (–CH2–CH2–O–)n and ROCO2Li) and inorganic Li salts (Li2CO3, LiOH, LiF and Li2O), while the amount of the inorganic species obviously increases with further lithiation to lower potential (0.01 V). Although producing irreversible capacity at the 1st lithiation process, the formed SEI layer increases the wettability of the active material and the surface Li+ concentration, so that the Rcon and RSEI decrease at low potentials (≤1.5 V), and therefore facilitates the lithiation reaction.
In order to determine the effect of the Li–Zn alloy on the cycling performance, a higher cut-off voltage (0.25 V) for the lithiation process was used in galvanostatic cycling at a current density of 1.0 A g−1 (see Fig. S11†). Here the maximum capacity of the ZnS-C/G electrode is achieved at the 34th cycle (402 mA h g−1). Interestingly, the capacity fading of the ZnS-C/G electrode is suppressed. The capacity (339 mA h g−1) at the 200th cycle is still higher than the first value (324 mA h g−1). This implies that the alloy process is detrimental during long cycling, although inducing the additional alloying capacity.
To understand the capacity fade during long-term cycling, EIS were recorded at the delithiated state of 3.0 V during the cycling test. The resistance evolution in the lithium half-cell can reveal the change of internal resistance. It is noteworthy that the resistance of the ZnS-C/G pristine electrode (∼190 Ω cm) is lower than that of the pristine ZnS (∼240 Ω cm), which is due to the carbon coating, as shown in Fig. S12.† On the other hand, the steeper inclined line in the Nyquist plots indicates a faster ion diffusion in the ZnS-C/G electrode than in the ZnS electrode. Furthermore, a gradual decrease in the size of two depressed semicircles in the medium–high frequency region (except for the 1st Nyquist plot) can be observed in the ZnS electrode (see Fig. 9c), suggesting a slight decrease in the resistance. In contrast, it is observed in the ZnS-C/G electrode that the resistance remains low until the 110th cycle and then increases continuously in later cycles, which corresponds to the capacity decrease from its highest value in the galvanostatic cycling test (see Fig. 9d). This can be ascribed to the fact that the coated carbon effectively improves the ZnS-C/G redox reaction and protects the ZnS bulk from direct exposure to the electrolyte during the initial hundred cycles. However, in the later cycles, the protection of the coated carbon becomes weak due to the repeated volume change of the ZnS grains.
To elucidate the charge storage mechanism, either surface-controlled (capacitive) or diffusion-controlled (bulk) dominates the charge storage process, the CV rate test was performed.61,62Fig. 10a and b display the CV curves measured at scan rates from 0.1 to 5.0 mV s−1. By increasing the scan rate, the CV peaks become broader but keep a similar shape. According to the power-law dependence relationship (ip = avb) between the generated peak current (ip) in the CV curves and scan rate (ν), the value of b can indicate the charge storage mechanism, where b = 1 corresponds to the capacitive process and b = 0.5 is related to the diffusion-controlled process.63 As shown in Fig. 10c, b was calculated based on eqn (2). The linear fitting of both of the cathodic peak (b = 0.76) and the anodic peak (b = 0.71) for the ZnS-C/G indicates a mixed diffusive/capacitive-controlled storage mechanism; whereas the ZnS electrode displays a predominance of the diffusion-controlled storage process (cathodic b = 0.56 and anodic b = 0.68). This trend confirms that the coated carbon can increase the capacitive storage proportion in the ZnS-C/G composite. Since the capacitive process shows higher rate performance than the diffusive-controlled process, the ZnS-C/G composite shows a much higher rate performance as a result.
log|ip| = log![]() ![]() ![]() | (2) |
|iv|/v0.5 = k1v0.5 + k2 | (3) |
In addition, the contribution of capacitive energy storage in the ZnS-C/G electrode can be calculated according to eqn (3), where k1 and k2 are the coefficients denoting the proportion of the surface-controlled process and the diffusion-controlled process, respectively. k1 can be determined from the slope of iv/v0.5vs. v0.5, and the intercept corresponds to the k2 value. As shown in Fig. 10d, ∼64.2% of the total capacity is derived from the capacitive process at a scan rate of 0.25 mV s−1. The capacitive contribution increases as the scan rate increases, i.e., as high as 65.9% at 0.5 mV s−1 (Fig. 10e) and 73.9% at 1.0 mV s−1 (Fig. 10f). Therefore, the result confirms that the capacitive contribution dominates the charge storage process at high current density.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c9ta01382b |
‡ G. Tian and Z. Zhao contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2019 |