Nikolas Királya,
Dominika Capkováb,
Miroslav Almášia,
Tomáš Kazdac,
Ondej Čechc,
Pavel Čudekc,
Andrea Straková Fedorkováb,
Maxim Lisnichukd,
Vera Meynene and
Vladimír Zeleňák*a
aDepartment of Inorganic Chemistry, Faculty of Sciences, Pavol Jozef Šafárik University in Košice, Moyzesova 11, 04154, Košice, Slovak Republic. E-mail: vladimir.zelenak@upjs.sk
bDepartment of Physical Chemistry, Faculty of Sciences, Pavol Jozef Šafárik University in Košice, Moyzesova 11, 04154, Košice, Slovak Republic
cDepartment of Electrical and Electronic Technology, Faculty of Electrical Engineering and Communication, Brno University of Technology, Technická 10, 616 00, Brno, Czech Republic
dInstitute of Physics, Faculty of Sciences, Pavol Jozef Šafárik University in Košice, Park Angelinum 9, 04001 Košice, Slovak Republic
eLaboratory of Adsorption and Catalysis (LADCA), Department of Chemistry, University of Antwerp, Universiteitsplein 1, 2610 Wilrijk, Belgium
First published on 24th August 2022
Lithium–sulphur batteries attract increasing interest due to their high theoretical specific capacity, advantageous economy, and “eco-friendliness”. In this study, a metal–organic framework (MOF) GaTCPP containing a porphyrinic base ligand was used as a conductive additive for sulphur. GaTCPP was synthesized, characterized, and post-synthetically modified by the transition metal ions (Co2+/Ni2+). The doping of GaTCPP ensured an increase in the carbon dioxide adsorption capacities, which were measured under different conditions. Post-synthetic modification of GaTCPP with Co2+/Ni2+ ions has been shown to increase carbon dioxide storage capacity from 22.8 wt% for unmodified material to 23.1 wt% and 26.5 wt% at 0 °C and 1 bar for Co2+ and Ni2+-doped analogues, respectively. As a conductive part of cathode material, MOFs displayed successful sulphur capture and encapsulation proven by stable charge/discharge cycle performances, high-capacity retention, and coulombic efficiency. The electrodes with pristine GaTCPP showed a discharge capacity of 699 mA h g−1 at 0.2C in the fiftieth cycle. However, the doping of GaTCPP by Ni2+ has a positive impact on the electrochemical properties, the capacity increased to 778 mA h g−1 in the fiftieth cycle at 0.2C.
Carbon dioxide, as one of a group of greenhouse gases generated by the burning of fossil fuels, is being produced at an alarming rate by human society. To stabilize the level of CO2 in the environment, it is important to create a cost-effective group of carbon dioxide adsorbents. In this process, liquid organic amines are used, which have a high adsorption capacity for carbon dioxide, as it is chemisorption. However, their disadvantage is regeneration, which requires a high temperature (∼50 °C) and high pressure (20–60 bar). There is currently a growing interest in the adsorption of carbon dioxide by means of highly porous solids, and MOFs compounds are at the forefront of this interest due to their large specific surfaces and the ability to functionalize the pore walls.30 The first key criterion for choosing suitable MOFs for CO2 adsorption is that the MOFs' pores have to be compatible with the kinetic diameter of the CO2 molecules. The second criterion is the adsorption capacity. MOFs with polar (hydroxy, azo, amino, imino, etc. groups) pores have a higher CO2 adsorption capacity, with compare to MOFs without these groups, due to the quadrupole moments of CO2 molecules.
Electrochemical energy storage technologies, with their numerous advantages such as “eco-friendly”, high efficiency, and wide applicability, have attracted tremendous attention as important possible solutions to the dramatic increase in environmental pollution.31–33 Lithium-ion (Li-ion) batteries are the most widespread energy storage systems used in portable electronic devices, electric vehicles (EVs), stationary energy storage and satellites.34–38 Nevertheless, as energy storage requirements expand, the energy density and specific capacity of commercial Li-ion batteries currently in use are proving to be insufficient for future applications. One of the possible candidates to replace Li-ion batteries are lithium-sulphur (Li–S) batteries thanks to their high theoretical specific capacity of 1675 mA h g−1, high energy density (2600 W h kg−1), natural abundance, and the low cost of sulphur.39,40 However, there are fundamental challenges regarding the sulphur cathode, including the low conductivity of sulphur and the final discharge product of lithium sulphide, the dissolution and diffusion of polysulphide intermediates in the electrolyte, and the large volumetric variations (∼80%) of the sulphur and polysulfides during cycling. During discharging, sulphur reacts with lithium ions, and the intermediates are higher (soluble; Li2Sx, 4 < x < 8) and lower (insoluble; Li2S2, Li2S) polysulfides. The reverse reaction is present during charging.41–43 The soluble polysulfides are involved in the “shuttle effect” where higher polysulfides migrate between the electrodes.44,45 To avoid these problems, various improvements were published,46–49 e.g. conductive additives with high stabilities50,51 and organic binders42,43 in the cathode material, separator modifications52 and insertion of an interlayer.53 Nevertheless, the search for functional materials for Li–S battery improvement remains urgently required.54,55 Improving the electrochemical behaviour of the sulphur cathode can be achieved with various carbon materials with high conductivity, tunability, and porosity.56 As such, several types of materials have been used to improve cyclic stability: functionalized carbon,57,58 metal oxides/sulphides/nitrides,59–61 covalent-organic frameworks62 and metal–organic frameworks,63–69 all of the mentioned materials have a high affinity for polar polysulphides.
To be employed as sulphur conductive additives in Li–S batteries, first of all, MOFs need to be electrochemically stable during cycling to ensure a permanent porous structure and thus confine sulphur and polysulfides inside. Second, the pore structures need to be well-designed, with suitably large pore size. Third and most importantly, these pores need to offer suitable chemical environments to effectively interact with sulphur and polysulfides.70–72 For example, ZIF-8 is one of the most widely studied materials in Li–S batteries application, showing a different specific capacity depending on the preparation of the composite used as a cathode in a range between 400–900 mA h g−1.73
Herein, we report the example of porphyrinic MOF self-assembled from porphyrin H2TCPP4− ligands, which was post-synthetically doped by Co2+/Ni2+ ions as adsorbent of carbon dioxide and the sulphur conductive additive for Li–S batteries. Accessible open metal sites of Co2+/Ni2+ ions increase adsorption of carbon dioxide from 22.8 wt% for undoped material (GaTCPP) to 26.5 wt% for Ni2+ analogue. The initial discharge capacity of the electrode with unmodified GaTCPP reaches the value of 667 mA h g−1 at 0.2C, and after fifty cycles, the capacity increases up to 699 mA h g−1. The doping by Ni2+ ions increases the conductivity of the electrode material, which increases the capacity, although the addition of Co2+ ions resulted in a capacity decrease and instability of the electrode. In addition, the cathode material with a pristine GaTCPP exhibited stable cycle performance, low fading rate per cycle (0.07% at 0.5C after 200 cycles), and high coulombic efficiency (94.6% at 0.5C during 200 cycles). The capacity fading rate per cycle during long-term cycling at 0.5C for GaTCPP(Co) was 0.14% and for the GaTCPP(Ni) electrode reached the value of 0.11%.
Post-synthetically modified materials with Co2+ or Ni2+ ions were obtained by soaking 200 mg GaTCPP(AS) in appropriate nitrate salt solutions in 10 mL of DMSO with concertation of 20 mg mL−1 at 90 °C for 72 hours under atmospheric conditions. After the reaction time, the reaction mixture was cooled down to ambient temperature and centrifugated with a 50% mixture of the suspension and acetone. The suspension was filtrated, washed with acetone and dried at 55 °C in an oven in an ambient atmosphere. Corresponding yields were 180 mg, 90% for GaTCPP(Co) and 150 mg, 75% for and GaTCPP(Ni) (both based on GaTCPP(AS)).
The infrared spectra of the samples were measured at laboratory temperature and recorded using an Avatar FTIR 6700 spectrometer in the range of wavenumbers 4000–400 cm−1 with 64 repetitions for a single spectrum, using the ATR (attenuated total reflectance) technique.
The thermal behaviour of prepared samples was studied by thermogravimetric analysis (TGA) and differential thermal analysis (DTA) with a sample weight of approximately 20 mg, using corundum crucibles. Samples were heated in the temperature range of 25–900 °C with a heating rate of 9 °C min−1 in a dynamic air atmosphere with a flow rate of 50 cm3 min−1, using a Netzsch 409-PC STA apparatus.
The powder X-ray diffraction (PXRD) patterns were obtained on a Rigaku MiniFlex 600 powder diffractometer using Cu Kα radiation (λ = 1.5406 Å) in 2 theta range 5–60°.
The morphology of the MOF materials was investigated using transmission electron microscopy (TEM) and energy-dispersive X-ray spectroscopy (EDS) on a JEOL JEM 2100F UHR, equipped with an EDS analyser. The surface characterization and elemental mapping of the electrode materials were performed by scanning electron microscope TESCAN VEGA 3, equipped with an EDS analyser. The surface areas and pore volumes of the samples were measured by argon sorption at −186 °C using a Quantachrome AUTOSORB-1-MP automated gas sorption system. Prior to the measurements, the samples were degassed in a vacuum at 100 °C for 2 hours followed by an increase in temperature to 140 °C, which was held for 14 hours. The total surface area was calculated via the Brunauer Emmett Teller (BET) equation, and the micropore volume was obtained using the DFT method (NLDFT kernel). Adsorptions of carbon dioxide at 0 °C and 20 °C were measured using a Quantachrome AUTOSORB-iQ-C combined volumetric and dynamic sorption system. Before measurements, the samples were activated in a vacuum at 100 °C for 2 hours followed by an increase in temperature to 140 °C which was held for 14 hours. The composition of MOF materials was analyzed via X-ray photoelectron spectroscopy (XPS) with a Kratos Axis Supra with an Al Kα X-ray source.
All electrochemical measurements were performed using a BioLogic VMP3 potentiostat. Cyclic voltammetry (CV) was measured in a potential window from 1.8 V to 3.0 V (vs. Li/Li+) and the scan rate was set to 0.1 mV s−1. Galvanostatic cycling was carried out within the potential range of 1.8–2.8 V (vs. Li/Li+). Electrochemical impedance spectroscopy (EIS) was measured in the frequency range of 1 MHz to 100 mHz with an amplitude of 10 mV.
After successful synthesis of GaTCPP(AS), the material was post-synthetically modified with Co2+ and Ni2+ ions. Mentioned ions were bound to the central porphyrin ring of the TCPP ligand throughout the modification process. Two pyrrole aromatic scaffolds were deprotonated during the coordination process to compensate for the positive charge of included metal ions.
Vibration | Wavenumber | ||
---|---|---|---|
GaTCPP(AS) [cm−1] | GaTCPP(Co) [cm−1] | GaTCPP(Ni) [cm−1] | |
ν(OH)H2O | 3401 | 3373 | 3378 |
ν(NH)pyrrole | 3314 | — | — |
ν(CH)aliph. DMF | 2971/2867 | 3002/2906 | 2999/2903 |
ν(CO)DMF | 1706 | 1708 | 1707 |
νas(COO) | 1585 | 1585 | 1585 |
ν(CC) | 1546 | 1546 | 1546 |
νs(COO) | 1419 | 1417 | 1417 |
ν(SO) | — | 1032 | 1032 |
γ(C–H)1,4-subst. | 796 | 795 | 796 |
γ(C–H)out of plane | 709 | 709 | 709 |
δ(CC)skeleton | 504 | 506 | 506 |
The thermogravimetric curves of the prepared materials, GaTCPP(AS) (black line in Fig. 2b), GaTCPP(Co) (blue line in Fig. 2b), and GaTCPP(Ni) (red line in Fig. 2b), display thermal stability of prepared materials. In the TG curves, we can observe three characteristic steps of weight loss. The first small weight loss, about 3.3 wt%, is observed up to 120 °C and can be attributed to the release of lattice water molecules from the channel system. The second weight loss observed between 120 °C and 300 °C can be attributed to the removal of DMF (∼25 wt%) or DMSO (from ion-modified samples) from the pores. The materials were thermally stable up to 300 °C. Above this temperature, the decomposition of the framework (∼50 wt%) started at around 350 °C. The residual solid product represented 20 wt%, 24 wt%, and 27 wt%, corresponding to the gallium oxide (calculated 18 wt%) and the mixture of Ga2O3 with cobalt/nickel oxides. An increasing percentage of residual masses confirms the presence of metal oxides of corresponding ions in the modified samples. The amounts of released solvent molecules differ slightly due to sample storage conditions before measurements and the synthetic procedure for preparing GaTCPP(Co) and GaTCPP(Ni) (possible presence of DMSO).
The crystallinity and bulk composition of all prepared materials were studied by powder X-ray diffraction (PXRD), and obtained PXRD patterns are depicted in Fig. 2c. Fig. 2c shows the comparison of the experimental and calculated PXRD patterns from the single-crystal X-ray data for GaTCPP(AS).74 Both patterns of GaTCPP(Co) (red curve) and GaTCPP(Ni) (blue curve) are almost identical, evidencing the phase purity of the prepared samples. The differences in intensities of diffraction lines can be attributed to the variation in the preferred orientation of the crystallites in the analysed powdered materials and the presence of solvent molecules in frameworks. Moreover, it can be concluded from the results obtained using PXRD measurements, that the modification of GaTCPP(AS) with Co2+/Ni2+ ions does not degrade the quality of GaTCPP framework.
The particle size, morphology, and successful incorporation/distribution of metal ions in the prepared samples were investigated by TEM measurements. Fig. 3 shows TEM images of the GaTCPP(Co) and GaTCPP(Ni) samples, which showed that the crystals of the compounds have a rod-like shape with different sizes, approximately from 60 to 300 nm for both materials and tend to agglomerate. The EDS mapping was used to confirm the elemental composition and distribution of Co2+/Ni2+ in prepared materials (yellow dots in Fig. 3a-Co and b-Ni). According to the obtained results from EDS mapping, we can conclude the presence of the desired ions and their regular distribution in the samples. SEM images of GaTCPP(AS), GaTCPP(Co), and GaTCPP(Ni) are available in ESI (see Fig. S1†).
Fig. 3 TEM images of the modified materials: (a) GaTCPP(Co) and (b) GaTCPP(Ni) with elemental mapping of Co2+/Ni2+ (yellow), Ga3+ (green) and N (red) using EDS analysis. |
In order to further investigate the composition of GaTCPP MOF materials and the chemical state of the present metals, XPS characterization was carried out. The wide-scan XPS spectra of GaTCPP(AS), GaTCPP(Co), and GaTCPP(Ni) are shown in Fig. 4. The peak located at 20 eV is not possible to unequivocally determine, it may be attributed to C 2s, Ga 3d, or O 2s. All observed peaks are slightly shifted to lower binding energy for modified GaTCPP(Ni) compared to GaTCPP(AS) and GaTCPP(Co). The XPS peak of Ga shows the presence of Ga in the MOF materials as it was observed in the TEM image. The peaks corresponding to Ga 3p3/2 and Ga 3p1/2 are observed at 103 and 106 eV for GaTCPP(Ni), and at 105 and 109 eV for GaTCPP(AS)/(Co), respectively. Ga 3s is found at 158 eV for GaTCPP(Ni) and 161 eV for GaTCPP(AS)/(Co). The Auger peak shapes Ga LMM can be seen around 423 eV. The peaks related to Ga 2p3/2 and Ga 2p1/2 are accessible at 1115 eV and 1142 for GaTCPP(Ni), and at 1119 eV and 1146 eV for GaTCPP(AS)/(Co), respectively. The presence of carbon (C 1s) can be observed at 281 eV for GaTCPP(Ni) and 285 eV for GaTCPP(AS)/(Co). The peak corresponding to N 1s is found at 396 eV for GaTCPP(Ni) and 399 eV for GaTCPP(AS)/(Co). Oxygen O 1s is seen in GaTCPP(Ni) spectrum at 528 eV and GaTCPP(AS)/(Co) spectra at 533 eV. The Auger peak O KLL is located around 978 eV. The presence of cobalt in GaTCPP(Co) is confirmed by a peak at 779 eV corresponding to Co 2p3/2. Nickel in GaTCPP(Ni) is determined by a peak at 852 eV, which is attributed to Ni 2p3/2. The atomic and mass concentration of the elements in GaTCPP materials is summarized in Table S1 in ESI.†
After pressing, the surface of the sulphur-containing cathodes with the GaTCPP samples, and its modifications with Co2+ and Ni2+, were also studied by SEM and EDS analyses (see Fig. 5). The SEM images are depicted at a view field of 41.5 μm, and elemental mapping was probed on the area of 400 μm × 400 μm. The mapped elements on the cathode surface were S (green colour), C (red colour), and Ga (blue colour), respectively. The elements Co2+ a Ni2+ were not analysed in the electrode by EDS, as they were previously analysed in the pure material. The electrode materials containing GaTCPP(AS) and GaTCPP(Ni) (Fig. 5a and c) are homogenous with uniform distribution of the elements. However, the electrode prepared with GaTCPP(Co) (Fig. 5b) shows less homogeneity. The sulphur particles are aggregated, and this phenomenon is not visible in samples GaTCPP(AS) and GaTCPP(Ni).
Material | Adsorbate | |||||
---|---|---|---|---|---|---|
Ar | CO2 | |||||
−186 °C | 20 °C; 1 bar | 0 °C; 1 bar | ||||
SBET (m2 g−1) | Vp (cm3 g−1) | |||||
wt% | mmol g−1 | wt% | mmol g−1 | |||
GaTCPP(AS) | 1010 | 0.74 | 19.0 | 4.31 | 22.8 | 5.17 |
GaTCPP(Co) | 1177 | 0.73 | 15.8 | 3.59 | 23.1 | 5.26 |
GaTCPP(Ni) | 1188 | 0.88 | 25.5 | 5.77 | 26.5 | 6.00 |
Carbon dioxide measurements were performed at two temperatures, 0 °C and 20 °C (see Fig. S2a and b in ESI†). Carbon dioxide adsorption isotherms at 20 °C (see Fig. S2a in ESI† and Table 2) showed that the as-synthetized material GaTCPP(AS) adsorbed 103 cm3 g−1 of CO2 up to 1 bar, which corresponds to the storage capacity of 19.0 wt% (4.31 mmol g−1) CO2. The modified materials GaTCPP(Co) and GaTCPP(Ni) adsorbed 86 cm3 g−1 and 138 cm3 g−1 CO2 up to 1 bar, corresponding to 15.8 wt% (3.59 mmol g−1) and 25.5 wt% (5.77 mmol g−1) CO2, respectively. As the adsorption phenomenon is an exothermic process, the adsorbed amount of CO2 increases with decreasing temperature. The carbon dioxide adsorption measurements at 0 °C (see Fig. S2b in ESI† and Table 2) showed that materials adsorbed 115, 117, and 134 cm3 g−1 of CO2, corresponding to the maximal uptake of 22.8, 23.1, and 26.5 wt% (5.17, 5.26 and 6 mmol g−1) at 1 bar, respectively for GaTCPP(AS), GaTCPP(Co) and GaTCPP(Ni). Differences in the influence of doped materials, with the lowering of the temperature from 20 °C to 0 °C should be caused by the difference in binding energy or accessibility of binding sites. One of the most significant effects on increasing the capacity and selectivity of adsorptive, especially carbon dioxide over other gases are open metal sites (OMS). In the activation process, excess solvents are removed from the metal sites of the lattice, leading to the formation of open metal sites (OMSs). Doping with metal ions (Co2+/Ni2+) increases the number of OMSs in the framework, leading to the increased CO2 adsorption capacity of modified materials. Since the characteristic coordination number of the incorporated ions is six, we assume that they are coordinated in the cavity of the TCPP ligand by the orbital (in the equatorial plane), while coordinated solvent molecules occupy the orbital (axial position). Activation of materials results in removing solvents and the formation of a free orbital, which represents OMS. In fact, these metal centres form the skeleton's surface as active sites for the capture of CO2 molecules and their binding by dipole–quadrupole interactions.78 The increased CO2 sorption capacity of GaTPCC(Ni) compared to GaTCPP(Co) can be explained by the presence of highly polarizing Ni2+ adsorption sites, leading to the dense packing of carbon dioxide within the framework and large binding enthalpies.73
The rate performance of the S/GaTCPP(AS), S/GaTCPP(Co), and S/GaTCPP(Ni) electrodes at different current densities is illustrated in Fig. 7b. The C-rate value increases from 0.2 to 2C and back to 0.2C as displayed in the figure. The initial discharge capacities at 0.2C of the S/GaTCPP(AS), S/GaTCPP(Co), and S/GaTCPP(Ni) electrodes are 666.5, 578.2, and 733.8 mA h g−1, respectively. The electrode with pristine, unmodified GaTCPP(AS) exhibits highly stable cycle performance with sufficient discharge capacities even at a high C-rate, the discharge capacity at 2C was 378.0 mA h g−1. The electrode containing modified GaTCPP with Co achieved lower discharge capacities, and the cycle performance was less stable than that containing the unmodified material. The highest discharge capacities, despite multi-current cycling, were observed for the S/GaTCPP(Ni) electrode. The discharge capacity at 2C acquires the value of 439.6 mA h g−1, which demonstrates good cycle performance and reversibility. The average value of coulombic efficiency at different C-rates of the S/GaTCPP(AS), S/GaTCPP(Co), and S/GaTCPP(Ni) electrodes was as follows 92.9, 95.3, and 91.7% respectively. The comparison of charge/discharge curves at the last cycles of each C-rate is depicted in Fig. 7c–e. The charge/discharge voltage plateaus correspond well with the redox peaks in the CVs. Both plateaus are visibly suppressed for the S/GaTCPP(Co) electrode indicating a higher impact of the shuttle effect during cycling than in the other samples. The most obvious difference in the shape of the charge/discharge curves was observed from cycling at 2C. The half-capacity during discharging and charging at 0.2C for the S/GaTCPP(AS) and S/GaTCPP(Ni) electrodes were located at 2.00 and 2.40 V, respectively. However, the value of half-capacity for the S/GaTCPP(Co) electrode was shifted to 1.99 and 2.42 V, respectively. The shift of the half-capacity to lower potential during discharging and higher potential during charging may indicate lower stability of the electrode material and enhanced material degradation.
Comparison literature of MOF based on 1,4-BDC (benzenedicaboxylate) linker was applied as cathode support in a Li–S battery where the initial discharge capacity at 0.2C was 392 mA h g−1 with 46% retention after 100 cycles.79 Also, HKUST-1 was studied as a host for sulphur in ref. 80, the discharge capacity at 0.1C achieved a value of around ∼550 mA h g−1 and at 1C ∼300 mA h g−1. After 100 cycles, the discharge capacity decreased to 300 mA h g−1 at 0.1C. It can be noted that the structure of GaTCPP is more suitable for capturing sulphur and polysulphides during cycling, which results in higher and more stable capacity than mentioned MOF materials reported in the literature.
To further study the electrochemical properties of the electrode materials improved by GaTCPP and its modifications with Co2+ and Ni2+ ions in Li–S batteries, 200 cycles of charge/discharge test were carried out at 0.5C. As shown in Fig. 8, the initial discharge capacity for the S/GaTCPP(AS) electrode was 600.6 mA h g−1, and after 200 cycles the capacity retention reached the value of 86.1% which corresponds with the fading rate of 0.07% per cycle. Coulombic efficiency during long-cycling at 0.5C was 94.6%. For comparison, the S/GaTCPP(Co) electrode shows the discharge capacity in the 200th cycle of 374.0 mA h g−1 with a capacity retention of 71.4% and fading rate of 0.14% per cycle respectively. Coulombic efficiency during cycling was decreasing due to the instability of the electrode material and the average value was around 88.4%. Finally, the initial discharge capacity of the S/GaTCPP(Ni) electrode reached the value of 609.9 mA h g−1, and the capacity retention after 200 cycles was 78.5% with a fading rate of 0.11% per cycle. Beyond, the specific capacity of the S/GaTCPP(Ni) electrode was higher than S/GaTCPP(AS) up to the 85th cycle, then the capacity was comparable up to the 100th cycle and it was followed by a gentle decrease of capacity for the S/GaTCPP(Ni) electrode which resulted in the highest capacity of the S/GaTCPP(AS) electrode after 200 cycles. Coulombic efficiency was lower compared to the S/GaTCPP(AS) electrode and reached the value of 88.3%. In addition, discharge profiles during cycling at 0.5C for all electrodes are presented in Fig. 8b–d. The most significant decrease in high voltage plateau and the reduced amount of higher polysulphides after 200 cycles is observed for the S/GaTCPP(Co) electrode indicating the weakest ability to trap polysulphides. The S/GaTCPP(Ni) electrode achieved the highest capacity in the high voltage plateau, the stability of the high voltage plateau was noticeably higher than that of S/GaTCPP(Co). The modification of GaTCPP with Ni2+ ions showed enhanced polysulfide trapping than modification with Co2+ ions. The S/GaTCPP electrode exhibit the most stable cycle performance and efficient capture of polysulphides, as indicated by the high stability of the high voltage plateau throughout the galvanostatic cycling. To sum up, the S/GaTCPP(AS) and S/GaTCPP(Ni) electrodes exhibit high and comparable discharge capacities during long-cycling at 0.5C. Moreover, coulombic efficiency of the S/GaTCPP(Ni) electrode decreases with cycle number, while it is stable at a high value for the S/GaTCPP(AS) electrode. However, the obvious difference between capacity retention and coulombic efficiency is the higher stability of the S/GaTCPP(AS) electrode and efficient polysulphide confinement in the GaTCPP material structure. Chen et al.81 presented MOF Ce-UiO-66-BPDC (benzenebiphenyldicaboxylate) as a cathode host of the Li–S battery. The initial discharge capacity at 0.1C was 757 mA h g−1, the decrease of capacity in the first 5 cycles was rapid and after 100 cycles the capacity retention was only around 20%. Also, MOF-5 was applied as a matrix for sulphur, which had an initial discharge capacity that was higher compared to our materials (874 mA h g−1 at 0.5C), although after fifty cycles the capacity retention was unsatisfactory and exhibit the value of 31.1%.82 The inferior cycling stability of the S/GaTCPP(Co) electrode compared to the S/GaTCPP(Ni) electrode may be due to weaker coordination between Co2+ and polysulfides than Ni2+ and polysulfides. The length of the high-voltage plateau does not shrink significantly for the S/GaTCPP(Ni) electrode, which is in agreement with improved coordination between Ni2+ and polysulfides. Moreover, the results are in accordance with the order of stability constant of complexes by bivalent ions of the first–row transition metals Mn2+ < Fe2+ < Co2+ < Ni2+ < Cu2+(II).83 The comparison of parameters obtained from multi-current and long-term cycling for all electrodes based on GaTCPP structure is summarized in Table 3.
Parameter | S/GaTCPP(AS) | S/GaTCPP(Co) | S/GaTCPP(Ni) |
---|---|---|---|
Capacity [mA h g−1] in 1st cycle at 0.2C | 666.5 | 578.2 | 733.8 |
Capacity [mA h g−1] in 20th cycle at 0.2C | 677.4 | 626.1 | 746.6 |
Capacity [mA h g−1] in 25th cycle at 0.5C | 587.4 | 543.2 | 648.0 |
Capacity [mA h g−1] in 30th cycle at 1C | 502.9 | 448.6 | 557.1 |
Capacity [mA h g−1] in 35th cycle at 2C | 378.0 | 331.0 | 436.9 |
Capacity [mA h g−1] in 50th cycle at 0.2C | 699.2 | 590.6 | 778.0 |
Coulombic efficiency [%] – various C-rates | 92.9 | 95.3 | 91.7 |
Capacity [mA h g−1] in 1st cycle at 0.5C | 600.6 | 523.8 | 609.9 |
Capacity [mA h g−1] in 100th cycle at 0.5C | 597.9 | 534.8 | 597.9 |
Capacity [mA h g−1] in 200th cycle at 0.5C | 516.9 | 374.0 | 478.9 |
Capacity retention after 200 cycles [%] | 86.1 | 71.4 | 78.5 |
Coulombic efficiency [%] – 0.5C | 94.6 | 88.4 | 88.3 |
In order to investigate the effect of various GaTCPP MOF materials on the electrochemical behaviour of the electrode, electrochemical impedance spectroscopy (EIS) was performed before cycling and after cycling at 2C 35 cycles. The EIS spectra of fresh and aged Li–S cells are shown in Fig. 9. All EIS curves consist of two semi-circles and a straight line in low frequencies. The applied equivalent circuit for the simulation of the EIS spectra was proposed in ref. 84 and depicted in Fig. 9. The electrolyte resistance (Re) represents the contribution of ohmic resistance. The first loop is expressed by Rint||CPEint representing interphase contact resistance and the capacitance of the sulphur cathode bulk. The second loop is represented by Rct||CPEdl, the charge transfer resistance, and the capacitance of the double layer on the electrode. The last part of the spectra is described as Rdiff||CPEdiff with Warburg element (W) and reflects the diffusion process. The electrolyte resistance was shifted to a slightly higher resistance for the S/GaTCPP electrode (4.5 Ω compared to 2.2 Ω for S/GaTCPP(Co) and 2.1 Ω S/GaTCPP(Ni)). After cycling, the Re of the S/GaTCPP electrode decreased to 2.9 Ω and raised to 4.4 Ω for S/GaTCPP(Co) and S/GaTCPP(Ni). The decrease of the electrolyte resistance after cycling for the S/GaTCPP may be associated with improved contact of sulphur with the conductive support for the GaTCPP than in its modified forms. The interphase contact resistance is associated with the position of sulphur in the porous structure of the cathode. The Rint was comparable for the fresh cells (5.0 Ω for S/GaTCPP, 5.9 Ω for S/GaTCPP(Co), and 4.6 Ω for S/GaTCPP(Ni)). After cycling, the Rint decreased for the S/GaTCPP (3.0 Ω) and S/GaTCPP(Ni) (0.6 Ω) but increased for the S/GaTCPP(Co) (6.8 Ω). This phenomenon may be related to the improved position of sulphur in GaTCPP and GaTCPP(Ni). The reaction kinetics influence the charge transfer resistance. Insignificant differences were observed for the fresh cells for charge transfer resistance (0.7 Ω for S/GaTCPP, 0.5 Ω for S/GaTCPP(Co), and 0.8 Ω for S/GaTCPP(Ni)). The cycled cells showed higher charge transfer resistance, the highest value showed the S/GaTCPP(Co) electrode (16.7 Ω), lower was observed for the S/GaTCPP (13.8 Ω), and the lowest was for the S/GaTCPP(Ni) (9.4 Ω). It may be assumed that the reaction kinetics was improved for the S/GaTCPP and S/GaTCPP(Ni) electrodes.
Fig. 9 The equivalent circuit and EIS spectra of the S/GaTCPP(AS), S/GaTCPP(Co), and S/GaTCPP(Ni) electrodes (a) before cycling and (b) after 35 cycles. |
The comparison of the electrochemical performance of presented GaTCPP-based electrodes in Li–S batteries with the state-of-the-art is presented in Table 4. The initial discharge capacity of GaTCPP-based electrode materials is mostly lower than the compared MOF materials, but their stability is worse and their capacity retention at the end of cycling is lower compared to our electrode materials.
MOF | Initial capacity [mA h g−1] | Number of cycles | Final capacity [mA h g−1] | C-Rate | S content [wt%] | Capacity retention [%] | References |
---|---|---|---|---|---|---|---|
MOF-5 | 874 | 50 | 272 | 0.5C | 58 | 31.1 | 82 |
ZIF-67 | 1225 | 100 | 422 | 0.1C | 75 | 34.5 | 85 |
Ce-UiO-66-BPDC | ∼700 | 50 | ∼300 | 0.1C | 25 | 42.3 | 81 |
Mg-1,4-BDC | 392 | 100 | 180 | 0.2C | 69 | 46.0 | 79 |
HKUST-1 | 431 | 300 | 286 | 0.5C | 30 | 66.4 | 65 |
GaTCPP(Co) | 524 | 200 | 374 | 0.5C | 60 | 71.4 | This work |
GaTCPP(Ni) | 610 | 200 | 479 | 0.5C | 60 | 78.5 | This work |
GaTCPP | 601 | 200 | 517 | 0.5C | 60 | 86.1 | This work |
Footnote |
† Electronic supplementary information (ESI) available. See https://doi.org/10.1039/d2ra03301a |
This journal is © The Royal Society of Chemistry 2022 |