Next Article in Journal
Folic Acid Ameliorates Renal Injury in Experimental Obstructive Nephropathy: Role of Glycine N-Methyltransferase
Next Article in Special Issue
NMR-Relaxometric Investigation of Mn(II)-Doped Polyoxometalates in Aqueous Solutions
Previous Article in Journal
Recent Advances in Manganese-Based Materials for Electrolytic Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Behavior of Reduced Graphene Oxide Supported Germanium Oxide, Germanium Nitride, and Germanium Phosphide as Lithium-Ion Battery Anodes Obtained from Highly Soluble Germanium Oxide

by
Alexey A. Mikhaylov
1,
Alexander G. Medvedev
1,
Dmitry A. Grishanov
2,
Timur M. Fazliev
1,
Vasilii Chernyshev
1,
Elena A. Mel’nik
1,
Tatiana A. Tripol’skaya
1,
Ovadia Lev
2,3 and
Petr V. Prikhodchenko
1,*
1
Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, Leninskii prosp. 31, 119991 Moscow, Russia
2
The Casali Center of Applied Chemistry, The Institute of Chemistry, The Hebrew University of Jerusalem, Jerusalem 9190401, Israel
3
The Harvey M. Krueger Family Center for Nanoscience and Nanotechnology, The Hebrew University of Jerusalem, Edmond J. Safra Campus, Jerusalem 9190401, Israel
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6860; https://doi.org/10.3390/ijms24076860
Submission received: 13 March 2023 / Revised: 3 April 2023 / Accepted: 4 April 2023 / Published: 6 April 2023
(This article belongs to the Special Issue State-of-the-Art Materials Science in Russia—2nd Edition)

Abstract

:
Germanium and germanium-based compounds are widely used in microelectronics, optics, solar cells, and sensors. Recently, germanium and its oxides, nitrides, and phosphides have been studied as active electrode materials in lithium- and sodium-ion battery anodes. Herein, the newly introduced highly soluble germanium oxide (HSGO) was used as a versatile precursor for germanium-based functional materials. In the first stage, a germanium-dioxide-reduced graphene oxide (rGO) composite was obtained by complete precipitation of GeO2 nanoparticles on the GO from an aqueous solution of HSGO and subsequent thermal treatment in argon at low temperature. The composition of the composite, GeO2-rGO (20 to 80 wt.% of crystalline phase), was able to be accurately determined by the HSGO to GO ratio in the initial solution since complete deposition and precipitation were achieved. The chemical activity of germanium dioxide nanoparticles deposited on reduced graphene oxide was shown by conversion to rGO-supported germanium nitride and phosphide phases. The GeP-rGO and Ge3N4-rGO composites with different morphologies were prepared in this study for the first time. As a test case, composite materials with different loadings of GeO2, GeP, and Ge3N4 were evaluated as lithium-ion battery anodes. Reversible conversion–alloying was demonstrated in all cases, and for the low-germanium loading range (20 wt.%), almost theoretical charge capacity based on the germanium content was attained at 100 mA g−1 (i.e., 2595 vs. 2465 mAh g−1 for Ge3N4 and 1790 vs. 1850 mAh g−1 for GeP). The germanium oxide was less efficiently exploited due to its lower conversion reversibility.

1. Introduction

Germanium and binary inorganic germanium compounds (GeO2, GeP, Ge3N4) are widely used in various fields of science and technology [1]. Germanium is a critically important semiconductor for microelectronics, optics, solar cells, and sensing. The applications of germanium dioxide are determined by its optical and electrical properties [2] and are characterized by a wide band gap (more than 5 eV), high transparency in the visible and infrared regions, and a high refractive index (1.6–1.7) [3,4]. Germanium phosphide has emerged as an attractive candidate for broad-band and mid-infrared photonics [5]. Germanium nitride has attracted interest due to its unique properties [6,7,8]. It can be used as a thin film material for the passivation of semiconductors [9,10], for plasmonic devices [1], and as an effective non-oxide photocatalyst for water splitting [11]. Germanium has higher electrical and ionic conductivity than silicon, which makes it an attractive alternative for high-performance anode material [12,13,14]. Despite its high cost, elemental germanium and its compounds are promising electrode materials in metal-ion batteries [15,16,17,18] due to the large theoretical gravimetric and volumetric specific capacities of germanium (with the formation of Li15Ge4), equal to 1384 mAh g−1 and 7366 mAh cm−3, respectively. Unfortunately, large volume changes (by more than 300%) during Li alloying/de-alloying lead to the pulverization of electrode particles and destabilization of solid electrolyte interphase (SEI) films, which hinders the use of germanium-based materials in LIBs [19,20]. Various approaches have been proposed to prevent a dramatic capacitance decrease in Ge-based anodes. For example, the synthesis of carbon-containing composites [21,22,23,24], which endow elasticity and conductive network, can significantly increase the long-cycling stability of the anode material. In addition, fabricating nanosized particles or particles of a tailored morphology [20,25,26] is an effective strategy to avoid degradation during cycling. Nowadays, different germanium compounds and their composites are studied as active materials for electrodes of lithium-ion batteries: GeO2 [27,28,29], Ge3N4, GeP of different stoichiometry [30,31,32,33,34,35,36], and germanium chalcogenides (sulfides, selenides, tellurides) [37,38,39].
Germanium phosphide, which combines the benefits of both germanium and phosphorus, is considered a potential anode for lithium- and sodium-ion batteries. Currently, the main approaches to synthesizing nanosized germanium phosphide are limited to ball milling and separation of GeP particles. The ball milling leads to the formation of irregularly shaped particles with a large size spread (from several tens of nanometers to several microns) [40]. Meanwhile, exfoliated GeP showed relatively low capacity and cyclic stability [35]. Microspheres of germanium phosphide prepared by a solvothermal method showed attractive electrochemical properties as a potential anode material, such as a high first-cycle Coulombic efficiency of 83% and high reversible capacity (1400 mAh g−1 after 150 cycles at 0.2 C) [41]. Germanium phosphide particles were obtained by red phosphorus evaporation–condensation on the germanium substrate. The authors demonstrated that the morphology was determined by the morphology of the initial germanium reactant [42].
Compared to the phosphide and oxide, germanium nitride has been less studied as an LIB anode material, perhaps since the first publication showed that only 40% of the material participates in the reaction with lithium ions, and unreacted Ge3N4 remains in the core. Therefore, it was concluded that the conversion reaction forming Li3N is irreversible [43].
Germanium nitride is usually obtained by nitridation with gaseous ammonia at high temperatures and a long annealing time [11,44]. Kim et al. described a new technique for obtaining the Ge3N4@C composite by partial oxidation of germanium and nitridation by gaseous NH3 at 700 °C for 1 h. Subsequent carbonization with acetylene at 800 °C results in Ge3N4@C powder with a specific capacity of 600–700 mAh g−1 at 550 mA g−1 for 300 cycles as a LIB anode [45].
The introduction of carbon components, such as graphene or carbon nanotubes, can enhance the performance of the germanium oxide LIB anode. The observed reversible capacity was close to or even higher than the theoretical reversible capacity of germanium oxide [23,26,29,46,47,48,49,50]. These observations show the possible partial reversibility of Li2O formation. Lv et al. calculated that 44.4–44.6% of Li2O matrix participates in the reverse alloying during cycling at 100 mA g−1 due to intimate contact between graphene layers and GeOx nanoparticles [51]. However, the significant irreversible capacity loss of the first cycle remains a severe problem for germanium oxide anode materials [52].
We have previously reported on the green-hydrogen-peroxide-assisted sol–gel processing route to deposit thin films of p-block peroxo compounds on various substrates [53,54]. The thermal or chemical treatment of these coatings on graphene oxide layers results in the formation of the corresponding oxides, sulfides, and tellurides, which are effective as lithium-, sodium-, and potassium-ion battery anodes [50,55,56,57,58,59,60,61,62,63,64,65]. In particular, a peroxogermanate thin film was deposited in high yield at room temperature on graphene oxide from peroxogermanate sols [50]. Thermal treatment of the filtered material produced amorphous germanium dioxide, crystalline germanium dioxide, and elemental germanium films supported by reduced graphene oxide (rGO). The obtained composites exhibited high cycling stability and good rate performance as LIB anodes.
The germanium compounds are expensive reagents, which calls for processing protocols with a quantitative yield of germanium. For this, there is a need for new convenient starting germanium compounds with higher reactivity and better water solubility compared to known precursors. Herein, we report an environmentally friendly method for the preparation of germanium-based composite materials from a highly soluble germanium oxide (HSGO) and the performance of the Ge composite products as anode materials in lithium-ion batteries. Germanium phosphide and nitride were selected as target materials for two reasons: (i) The high lithium conductivities of lithium nitride and phosphide—the lithiation by-products of GeP and Ge3N4. Lithium nitride and lithium phosphide are known lithium superconductors [66,67], and amorphous Li3N [68,69] and Li3P [67,70] were also reported to exhibit high ionic conductivity (10−3 S·cm−1). (ii) To improve the performance of germanium compounds as LIB anodes, it is, of course, essential to exhibit high Ge-Li alloying performance, but the performance can be further enhanced by exploiting the charge capacity of the conversion reaction associated with the counter element. Due to the lower electronegativity of nitrogen and phosphorous compared to oxygen, the respective bond cleavage and reversibility of the reactions of lithium phosphide and nitride should be better than Li2O.

2. Results and Discussion

2.1. Synthesis

The choice of an effective germanium precursor in the form of a highly soluble form of amorphous germanium oxide made it possible accurately dose the amount of germanium introduced into the system and opened new possibilities for controlling the morphology and composition of composite materials. HSGO was obtained by rapid decomposition of ammonium peroxogermanate (NH4)6[Ge6(μ-OO)6(μ-O)6(OH)6].6H2O at 300 °C [71]. We have already demonstrated that HSGO has a high solubility (up to 100 g per liter of water) and reactivity and, therefore, is a convenient starting reagent for the preparation of various germanium compounds [71,72]. However, until now, HSGO has not been used to obtain germanium-based composite functional materials or electrode materials. The first stage of synthesis was the deposition of a film of germanium oxide nanoparticles on the surface of graphene oxide sheets from an aqueous solution of HSGO. High solubility of precursors in water allows for easy regulation of the loading of germanium dioxide in obtained materials. Three different loadings of germanium dioxide were chosen for subsequent thermal and chemical processing, namely, 20, 50, and 80 wt.% of GeO2. The addition of aqueous ammonia to the solution of HSGO increased the pH from 5.5 to about 8. Precipitation of germanium oxide species on the graphene oxide was then accomplished by the addition of ethanol, resulting in uniform amorphous germanium-based coating (Figure S1). According to X-ray powder diffractometry, the crystalline phase of germanium dioxide was not formed even for the highest loading of germanium (Figure 1a). Subsequent heat treatment at a relatively low temperature of 300 °C for 3 h resulted in the reduction of graphene oxide and the crystallization of nanosized germanium dioxide. The XRD of the product revealed broad low-intensity peaks that were assigned to the GeO2 phase (PDF 01-073-9108) (Figure 1c). After the heat treatment, the coating morphology did not contain large agglomerates on the rGO surface and retained the morphology of the initial material (Figure 2a,b and Figure S1).
According to the CHN analysis, the amount of carbon in GeO2-rGO composites corresponded to the initial ratio of HSGO to GO, resulting in 16.7, 42.9, and 75.5 wt.% for 20, 50, and 80 wt.% of GO, respectively. The reduction of graphene oxide, accompanied by the removal of oxygen-containing functional groups, decreased the carbon content in the heated materials, GeO2-rGO (Table 1). The CHN confirms the high yield of composite material in terms of germanium. Thus, the described procedure was a simple, low-cost, and highly effective method for the preparation of GeO2-rGO composite.
The GeO2-rGO composites are used as a suitable initial material for preparing crystalline germanium nitride and phosphide-decorated rGO. As far as we know, this is the first attempt to obtain rGO-supported Ge3N4 and GeP composites. Previously, only the corresponding carbonized phases were obtained and characterized [34,45]. The lack of research can be attributed to the limited number of wet methods for obtaining germanium oxide–graphene oxide composites due to the low solubility of germanium oxide.
Prolonged treatment of GeO2-rGO-80 by gaseous NH3 at 700 °C for 2 h yields long wires of germanium nitride (Figure S2). Reduced treatment time (up to 5–10 min) results in Ge3N4 agglomerates with an average size of 50–200 nm, which decorate the rGO surface (Figure 2c,d). It should be noted that a part of the crystalline phase appears as a uniform coating on the surface of reduced graphene oxide (Figure S3). An attempt to control the particle size of germanium nitride was made by adjusting the ratio between the loading of germanium dioxide and graphene oxide. Decreasing of GeO2 loading to 20 wt.% (GeO2-rGO-20) resulted in a uniform coating of Ge3N4 crystalline phase, as shown by the SEM and EDX analyses (Figure S4). Treatment of GeO2-rGO-50 by ammonia formed a Ge3N4-rGO-50 composite consisting of rGO decorated with about 100 nm nitride particles. EDX mapping (Figure S5) showed that a thin layer of the Ge3N4 phase uniformly coated the rGO surface. We were able to estimate the mass content of the germanium nitride phase on the basis of the CHN analysis. The calculated loadings of germanium nitride were 58.7, 45.7, and 33.5 wt.% for Ge3N4-rGO-80, Ge3N4-rGO-50, and Ge3N4-rGO-20, respectively (Table 1). The calculated values indicated the presence of a residual germanium dioxide phase in the Ge3N4-rGO-80 sample. The powder diffractograms of Ge3N4-rGO samples reflected the structure of β-Ge3N4 (PDF 00-011-0554, Figure 1d–f). Both samples contained an admixture of a low-crystallinity phase of germanium nitride that appeared as a background and broad peaks according to the corresponding diffractograms. Thus, the Ge3N4-rGO-50 and Ge3N4-rGO-80 were refined (by the methodology in reference [73] as composites with bimodal crystallite size distributions. The average crystallite size for the germanium nitride phase, which appeared as a uniform coating, was around 2 nm in all three samples, while the anchored particles were characterized by 20.5 and 31.5 nm in Ge3N4-rGO-50 and Ge3N4-rGO-80, respectively. The large particle parameters were able to be significantly affected by crystallite shape.
Heat treatment of GeO2-rGO material with red phosphorus in a sealed quartz tube led to the formation of a germanium phosphide phase, which was clearly confirmed by X-ray powder diffraction data (Figure 1g–i). The scanning electron microscopy and energy-dispersive X-ray spectroscopy demonstrated the formation of both a uniform coating of germanium phosphide on the surface of reduced graphene oxide and relatively large particles of germanium phosphide with an average crystallite size of 168 nm (computed by the Scherrer equation), as in the case of germanium nitride with a high loading of germanium dioxide in the starting material (Figure 2e,f and Figure S6). Similar to the synthetic procedure of germanium nitride, the decrease in the content of GeO2 to 50% formed particles with a smaller crystallite size of 38.4 nm. Subsequent decreasing of GeO2 loading to 20% revealed a uniform coating of reduced graphene oxide by germanium phosphide with an average crystallite size of 2.2 nm (Figures S7 and S8), similar to values obtained by Rietveld refinement of experimental data for materials with higher Ge loading. According to the EDX analysis, the Ge/P ratio corresponded to the GeP crystalline phase (Table 1).
Using HSGO as a precursor allowed us to, firstly, accurately control the amount of active material loaded into the system and, secondly, to form a uniform coating of germanium oxide on graphene oxide. In addition, regulating the germanium oxide-graphene oxide ratio and the reaction time allowed for the obtaining of materials of different morphologies. Thus, highly soluble germanium oxide is a suitable and effective precursor for fabricating germanium-based materials of various compositions and morphologies, which was demonstrated by synthesizing germanium nitride and phosphide composites of reduced graphene oxide. As a test case, all nine obtained composites were evaluated as active materials of anodes in lithium-ion batteries to demonstrate the usefulness, uniqueness, and versatility of the HSGO-based synthesis approach.

2.2. Electrochemical Evaluation

Cyclic voltammetry was carried out to determine the electrochemical properties of the composites at room temperature. The CV plot of GeO2-rGO-80 is shown in Figure 3a. The first reduction scan shows a shallow shoulder at 1.07, which was previously attributed to minor lithiation of the germanium oxide to form LiyGeO2. An intensive peak at 0.4 V was attributed to the formation of the SEI layer overlapping a conversion reaction (LiyGeO2 to Ge(0)). Then, a cathodic wave, at E < 0.2 V, was attributed to a Ge-Li alloying process: (xLi + Ge → LixGe (0 ≤ x ≤ 4.4)). Similarly, the discharge peaks on subsequent cycles at about 0.98 V and 0.46 V vs. Li/Li+ and the cathodic wave at <0.2 V were attributed to minute alloying to form LiyGeO2, a conversion reaction that produced elemental Ge and then Ge-Li alloying [24,50]. The charging peaks were present at 0.44 and 1.14 V. The first related to dealloying, and the second indicated the (re)oxidation of Ge to GeO2 [50,51,74].
The CV plot of the Ge3N4-rGO-80 anode was similar to the GeO2-rGO electrode (Figure 3b) and reflected the alloying–conversion reactions. As indicated by Pereira et al. [43], there is practically only one reduction wave corresponding to the conversion and alloying reaction (Equation (1)).
Ge3N4 + 3(4 + x)Li+ + 3(4 + x)e → 3LixGe + 4Li3N,
where 0≤ x ≤ 4.4. Pereira et al. [43] attributed the overlap of the alloying and conversion waves to the rapid reductive alloying of Ge(0) to form LixGe alloy phases. The small shoulder, at 0.2 V, on the first discharging scan may also be attributed to this alloying reaction. The two anodic peaks at 0.4 and 0.9 V were attributed to the delithiation reaction (LixGe → Ge + xLi+ + xe) and the partial recovery of lithium germanium nitride LizGeyN, although full recovery of the original germanium nitride Ge3N4 was unable to be proven by X-ray, and Pereira et al. [43] did not identify this phase. The persistence of the 0.9 V anodic peak (Figure 3b) showed that the conversion reaction was partially reversible, and nitride (re)formation contributed significantly to the observed charge capacity. The same can be concluded on the basis of the relatively moderate (around 18%) capacity loss in the first cycle. Irreversible formation of the inactive Li3N phase was noted [45], and Periera et al. [43] even reported that the 0.9 V peak disappeared after long cycling. Therefore, we carried out cyclic voltammetry of the Ge3N4-rGO-80 electrode after 150 discharge–charge cycles at 100 mA g−1 (Figure S10). The figure clearly indicates the presence of the 0.9 V peak and confirms the capacity contribution of reversible Li3N formation even after long cycling. The intimate contact between the germanium nitride coating and graphene oxide could be responsible for the favorable reversibility of the germanium nitride conversion reaction in this study. The CV plots of GeO2-rGO-20 and Ge3N4-rGO-20 are presented in Figure S9. They indicate the validity of our conclusions regarding the lower potential at which the conversion of germanium nitride took place compared to germanium oxide and the higher contribution of the conversion reaction in the former.
Figure 3c depicts the CV plots of the germanium phosphide anode, GeP-rGO-80. The study was conducted under the same conditions used to investigate the germanium oxide and germanium nitride electrodes. Comparison of the CV plots with the literature allows us to assign the following processes [30,32,35,40]. The interval 1.0–0.68 V corresponded to lithium intercalation with the formation of the LixGeP phase. The subsequent stage of conversion with the formation of Li3P and Ge appeared as peaks in the range of 0.68–0.40 V. The 0.4–0.005 V peaks were assigned to the alloying stage with LiyGe formation. At the charging scan, the active electrode demonstrated electrochemical oxidation peaks at the 0.005–0.68 V interval assigned to the LiyGe dealloying process. The next potential window at 0.68–1.1 V was attributed to the reversal of the conversion reaction with LixGeP formation. The broad, low-intensity peak at higher potentials was assigned to the deintercalation reaction, which led to the GeP formation. The lower germanium phosphide loading (GeP-rGO-20) provided a similar CV plot (Figure S9c) but with much lower currents and less distinct peaks. The germanium phosphide anodes, similar to the germanium nitride electrodes, exhibited significant reversibility and charge contribution to the conversion reaction. However, in all cases, the alloying reactions contributed a much larger charge capacity than the conversion reactions.
The galvanostatic charge–discharge curves of the three different germanium compounds are presented in Figure 3d–f. The first and second cycles at 100 mA g−1 showed good first cycle efficiency of about 68, 80, and 63% for GeO2-rGO-80, Ge3N4-rGO-80, and GeP-rGO-80, respectively, which were not inferior and mostly exceeded the best-published values, including those tested at a narrower potential window of up to 1.5 V [33,43,45,50].
The specific capacities depended on the Ge content in the active phase loadings (Figure 4). We carried out long cycling stability tests for 150 cycles at 100 mA g−1 and evaluated the rate performance at 100–3000 mA g−1 at a potential window of 0.005–3.0 V vs. Li/Li+. The decreases in specific capacity after 150 charge–discharge cycles compared to the second cycle were 44% (from 1404 to 785 mAh g−1), 66% (from 1696 to 575 mAh g−1), and 58% (from 990 to 408 mAh g−1) for the GeO2-rGO-80, Ge3N4-rGO-80, and GeP-rGO-80 anodes, respectively. The obtained results were not inferior to the recently reported data for GeO2 [23,24,50], Ge3N4 [43,45], and GeP anodes [35]. This can be attributed to the large volume expansion of the inorganic phase.
The electrode materials with lower germanium content demonstrated good cycling stability for germanium oxide, nitride, and phosphide. The specific cycling capacities of the Ge3N4-rGO-20 and GeP-rGO-20 were higher than for GeO2-rGO-80 after long cycling at 100 mA g−1. The capacity drops after 150 cycles were only 17% (from 714 to 592 mA g−1) and 23% (from 610 to 468 mA g−1) for Ge3N4-rGO-20 and GeP-rGO-20, respectively, while for GeO2-rGO-20 material, it was 51% (from 710 to 345 mA g−1). The increased stability at low loading can be attributed to the higher physical buffering of the larger fraction of the (flexible) rGO or to the better adhesion of the small nanoparticles to the rGO surface, having a larger specific contact area with the rGO support (i.e., contact area to the rGO per crystallite weight). Support for the second explanation was indirectly provided by the performance of the GeO2 electrodes, which exhibited monomodal particle size distribution (even for the 80 wt.% GeO2 loading) but exhibited similar cyclic stabilities at different germanium loadings.
According to CHN analysis, the carbon content in Ge3N4-rGO-20 and GeP-rGO-20 samples were equal to 76.2 and 74.2 wt.%, respectively (Table 1), corresponding to the ≈24 and ≈26 wt.% content of germanium nitride and phosphide, respectively. Thus, the approximate specific capacity values per gram of Ge3N4 and GeP were equal to 2595 and 1790 mAh g−1 at a 100 mAg−1 scan rate, respectively. The obtained values were close to the theoretical capacity for Ge3N4 (2465 mAh g−1) and GeP (1850 mAh g−1) charge capacities, indicating highly efficient Ge3N4-rGO-20 and GeP-rGO-20 electrode materials and the reversibility of Li3N matrix formation. In this argument, we ignored the contribution of the graphene oxide to the capacity, which, at most, would amount to 280 mA per g of active germanium compound. Comparison of a CV plot for a germanium nitride anode with higher and lower loadings of active phase (Figure 3b and Figure S9b, respectively) with those for germanium oxide allowed us to conclude with regards to the reversible Li3N formation resulting in higher specific capacity of germanium nitride electrodes.
The rate capabilities of germanium dioxide, nitride, and phosphide composite anodes with different loading of reduced graphene oxide at 100 to 3000 mA g−1 currents are presented in Figure 4d–f, respectively. All materials demonstrated good high-rate performance. An increase in charge–discharge current by a factor of 30 to 3000 mA g−1 resulted in the preservation of 33, 36, and 46% of the specific charge capacity of Ge3N4-rGO-20, Ge3N4-rGO-50, and Ge3N4-rGO-80 materials, respectively (Figure 4e). The GeP-based materials retained around 40% of the initial capacity at a high rate of 3000 mA g−1 (Figure 4f). Similar values, equal to 32, 50, and 48%, were obtained for initial germanium dioxide–rGO composites with 20, 50, and 80% loading of the inorganic phase, respectively. To the best of our knowledge, the obtained specific capacity value of 830 mAh g−1 at 3000 mA g−1 for Ge3N4-rGO-80 is the highest attained for electrodes based on germanium nitride, while the experimental values for germanium oxide and phosphide are similar to the best-published values [35].
We tested the GeO2-rGO-80, Ge3N4-rGO-80, and GeP-rGO-80 materials for cycling stability at a high rate of 1000 mA g−1 (Figure S11). The specific capacity values were equal to 360, 390, and 300 mAh g−1 after 500 cycles for GeO2-rGO-80, Ge3N4-rGO-80, and GeP-rGO-80, respectively.
Conversion anodes are characterized by high-voltage hysteresis and, thus, poorer energy efficiency compared to alloying and conversion-alloying anodes [75,76,77,78,79]. Since the electrodes exhibited considerable conversion reaction contribution, the voltage hysteresis was computed. The calculation protocol is provided in the Supporting Information. The average voltage hystereses calculated at the 30th cycle at 100 mA g−1 were equal to 0.70, 0.46, and 0.51 V for GeO2-rGO-80, Ge3N4-rGO-80, and GeP-rGO-80, respectively. The germanium nitride and phosphide anodes exhibited lower voltage hystereses compared to the GeO2-rGO anode and to other conversion electrodes, and their voltage hystereses were similar to tin sulfide electrodes, which exhibited a voltage hysteresis of 0.48 V [79]. The comparison of the voltage hysteresis, Δ V ¯ of several conversion LIB anodes, was reported by Zhong et al. [80]. The best voltage hystereses of LIB anodes, namely, Fe3O4 and MnO, showed ( Δ V ¯ ) > 0.55 V, which was higher than the values obtained for the germanium nitride LIB anode. This was not surprising. The Ge-based anodes studied in this research were better classified as conversion–alloying anodes rather than conversion reaction electrodes.

3. Materials and Methods

3.1. Synthetic Procedures

3.1.1. Synthesis of Highly Soluble Germanium Oxide (HSGO)

Synthesis was conducted according to [71]. Briefly, 0.5208 g of GeCl4 (2.43 mmol) was added to 20 mL of 5 wt.% H2O2. The precipitate was dispersed in an ultrasonic bath for 2 min. The dispersion was neutralized with 2 mL of NH4OH (12.5 wt.% solution) until pH 9. After an additional 5 min of stirring, 20 mL of alcohol was added for full precipitation. The precipitate was separated by filtration on a glass filter and washed three times with 3 mL of 50 wt.% EtOH, once with 3 mL EtOH, and once with 3 mL diethyl ether. A total of 0.3892 g of APG was obtained after drying (92.2% yield based on Ge). Then, 0.389 g of APG was heated in a furnace at 300 °C for 15 min to produce 0.231 g of HSGO powder (98.7% yield based on Ge).

3.1.2. Preparation of Graphene Oxide (GO) Dispersion

GO was synthesized from exfoliated graphite by a modified Hummers method, and a detailed protocol is described in our previous articles [54,58,64].

3.1.3. Synthesis of GeO2-rGO

A total of 7.5 g of aqueous GO dispersion (1.2 wt.%) was mixed with 17.5 mL of DIW with sonication. Then, 350 mg of HSGO and 1 mL of NH4OH 25 wt.% aqueous solution were introduced. The precipitation of GeO2 onto the GO surface was accomplished by the addition of 100 mL of ethanol under vigorous stirring. The coated GO was centrifuged and washed two times with ethanol and once with diethyl ether. The material was dried under vacuum at room temperature and then heated in Ar at 300 °C for 3 h. Samples with different GeO2 loading were prepared and assigned as GeO2-rGO-80, GeO2-rGO-50, and GeO2-rGO-20.

3.1.4. Synthesis of Ge3N4-rGO

GeO2-rGO powders with different Ge loading were introduced to the hot tube furnace under dry NH3 flow at 700 °C for 10 min. Dark powder was obtained and assigned as Ge3N4-rGO-80, Ge3N4-rGO-50, and Ge3N4-rGO-20.

3.1.5. Synthesis of GeP-rGO

The GeO2-rGO materials were mixed with red phosphorous and sealed in a quartz tube. The tube was heated at 590 °C for 2 h. Dark gray powder was collected and assigned as GeP-rGO-80, GeP-rGO-50, and GeP-rGO-20 depending on the initial GeO2 loading.

3.2. Material Characterization

Details of the material characterization techniques are provided in the Supplementary Information.

3.3. Electrochemical Measurements

The active electrode material was mixed with Super P (Alfa Aesar) and polyvinylidene difluoride (PVDF) (Sigma-Aldrich) in a weight ratio of 70:15:15 with N-methyl pyrrolidone (NMP) as the medium to form a slurry. The slurry was then coated on a roughened copper foil as a current collector using a doctor blade. The electrode was then dried at 80 °C and pressed in a roll press. The electrodes were cut into 16 mm diameter discs and further dried at 110 °C for 4 h in a vacuum before being introduced into an argon-filled glovebox. The average loading of active material on the Cu foil was 1.5–2 mg cm−2. The electrodes were assembled with Li metal counter electrodes in a 2016-type coin cell. A total of 1 M LiPF6 in EC:DEC was used as an electrolyte. Then, 5% v/v fluorinated ethylene carbonate (FEC, Alfa Aesar) was added to the electrolyte to improve solid electrolyte interface (SEI) stability. The coin cells were assembled in an Ar-filled glovebox with less than 0.1 ppm oxygen. The charge/discharge tests were performed with a Neware battery tester within a voltage window of 0.01–3.0 V vs. Li/Li+ at 100–3000 mA g−1 rate. Cyclic voltammetry (0.005–3.0 V, 0.1 mV s−1) was performed with a VersaSTAT3 potentiostat.

4. Conclusions

As far as we know, this is the first report on the Ge3N4-rGO and GeP-rGO composites. The highly soluble germanium oxide (HSGO) was used as a versatile precursor of germanium-based electrode materials. HSGO was prepared in mild conditions with high yield from the germanium coordination peroxo compound, namely, ammonium peroxogermanate (APG), (NH4)6[Ge6(μ-OO)6(μ-O)6(OH)6]·6H2O. In the first stage, HSGO was used as a precursor for the deposition of uniform coating consisting of germanium dioxide nanoparticles on reduced graphene oxide (GeO2-rGO). The loading of GeO2 phase was regulated by the ratio of HSGO and graphene oxide starting materials since full conversion was demonstrated. The highly reactive germanium dioxide on graphene oxide support allows for the conversion of the oxide to nitride and phosphide phases. It was shown that the morphology of Ge3N4 and GeP crystalline phases depended on the loading of germanium dioxide in the GeO2-rGO composite. For low loadings of 20 wt.% germanium compound, only a very thin film of active material was obtained, whereas at higher loading, coatings with bimodal size distribution were demonstrated.
The homogeneous distribution of germanium compounds (exclusively) over the surface of the graphene oxide was demonstrated in all the electrode coatings. However, in the germanium phosphide and nitride electrodes, the particle size distribution was bimodal, comprised of approximately 2 nm thick homogenous coverage of the rGO surface and larger nanoparticles of around 200 nm particles densely decorating the surface. The ratio between these two phases was a function of the germanium oxide loading; only thin-film coating was formed from 20 wt.% GeO2 loading, while the fraction of the large particles was considerably higher for the materials prepared from the 80 wt.% GeO2 loading. The charge capacity was accordingly higher for the high-loading materials, but alloying pulverization was much lower for the active materials with smaller particles. Excellent stability of Ge3N4-rGO-20 and GeP-rGO-20 composites was demonstrated by 150 cycles at 100 mA g−1 with charge capacity loss of less than 17% and 23%, respectively, in the last 150 cycles. The higher stability of the low Ge-loaded anodes was attributed to the superior adhesion of the uniform nanoparticle coverage of the rGO and its larger specific contact area rather than to the higher carbon content, since GeO2 electrodes did not exhibit superior stability at low germanium loading.
Germanium is an expensive resource, and therefore, even for niche applications, its theoretical charge capacity should be approached. The Ge3N4-rGO-20 and GeP-rGO-20 demonstrate that even at a 100 mA g−1 rate, the conversion reaction was close to quantitative, and the almost theoretical charge capacity of the germanium compounds was approached: 2595 vs. 2465 mAh g−1 for Ge3N4 and 1790 vs. 1850 mAh g−1 for GeP. The germanium oxide was less efficiently exploited due to its lower conversion reversibility compared to GeP and Ge3N4 electrodes, following previous reports.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms24076860/s1.

Author Contributions

Conceptualization, A.G.M., P.V.P. and O.L.; methodology, A.G.M. and P.V.P.; synthesis, D.A.G., A.A.M., T.M.F., V.C., E.A.M. and T.A.T.; characterization, A.A.M., A.G.M., D.A.G., E.A.M. and T.A.T.; writing—original draft preparation, A.A.M., A.G.M. and P.V.P.; writing—review and editing, O.L.; visualization, A.A.M., A.G.M., T.M.F. and E.A.M.; supervision, P.V.P.; project administration, A.G.M., P.V.P. and O.L.; funding acquisition, A.G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation, grant number 23-23-00583, https://rscf.ru/en/project/23-23-00583/ (accessed on 3 April 2023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and Supplementary Materials.

Acknowledgments

This work was conducted using the experimental facilities of the JRC PMR IGIC RAS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Emsley, J. Nature’s Building Blocks: Everything You Need to Know about the Elements; Oxford University Press: Oxford, UK, 2011. [Google Scholar]
  2. Lange, T.; Njoroge, W.; Weis, H.; Beckers, M.; Wuttig, M. Physical Properties of Thin GeO2 Films Produced by Reactive DC Magnetron Sputtering. Thin Solid Films 2000, 365, 82–89. [Google Scholar] [CrossRef]
  3. Xu, M.-F.; Shi, X.-B.; Jin, Z.-M.; Zu, F.-S.; Liu, Y.; Zhang, L.; Wang, Z.-K.; Liao, L.-S. Aqueous Solution-Processed GeO2: An Anode Interfacial Layer for High Performance and Air-Stable Organic Solar Cells. ACS Appl. Mater. Interfaces 2013, 5, 10866–10873. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, Z.K.; Li, M.; Yuan, D.X.; Shi, X.B.; Ma, H.; Liao, L.S. Improved Hole Interfacial Layer for Planar Perovskite Solar Cells with Efficiency Exceeding 15%. ACS Appl. Mater. Interfaces 2015, 7, 9645–9651. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, Z.; Guo, J.; Zhang, Y.; Liu, J.; Ponraj, J.S.; Dhanabalan, S.C.; Zhai, T.; Liu, X.; Song, Y.; Zhang, H. 2D GeP-Based Photonic Device for near-Infrared and Mid-Infrared Ultrafast Photonics. Nanophotonics 2020, 9, 3645–3654. [Google Scholar] [CrossRef]
  6. Hayakawa, R.; Yoshida, M.; Ide, K.; Yamashita, Y.; Yoshikawa, H.; Kobayashi, K.; Kunugi, S.; Uehara, T.; Fujimura, N. Structural Analysis and Electrical Properties of Pure Ge3N4 Dielectric Layers Formed by an Atmospheric-Pressure Nitrogen Plasma. J. Appl. Phys. 2011, 110, 064103. [Google Scholar] [CrossRef]
  7. Yuan, L.; Han, C.; Yang, M.-Q.; Xu, Y.-J. Photocatalytic Water Splitting for Solar Hydrogen Generation: Fundamentals and Recent Advancements. Int. Rev. Phys. Chem. 2016, 35, 1–36. [Google Scholar] [CrossRef]
  8. Jishiashvili, D.; Kapaklis, V.; Devaux, X.; Politis, C.; Kutelia, E.; Makhatadze, N.; Gobronidze, V.; Shiolashvili, Z. Germanium Nitride Nanowires Produced by Thermal Annealing in Hydrazine Vapor. Adv. Sci. Lett. 2009, 2, 40–44. [Google Scholar] [CrossRef]
  9. Zhang, S.L.; Wang, W.; Zhang, E.H.; Xiao, W. Half-Metallic Ferromagnetism in Transition-Metal Doped Germanium Nitride: A First-Principles Study. Phys. Lett. A 2010, 374, 3234–3237. [Google Scholar] [CrossRef] [Green Version]
  10. Yang, M.; Wang, S.J.; Peng, G.W.; Wu, R.Q.; Feng, Y.P. Ab Initio Study on Intrinsic Defect Properties of Germanium Nitride Considered for Gate Dielectric. Appl. Phys. Lett. 2007, 91, 132906. [Google Scholar] [CrossRef]
  11. Maeda, K.; Saitoh, N.; Inoue, Y.; Domen, K. Dependence of Activity and Stability of Germanium Nitride Powder for Photocatalytic Overall Water Splitting on Structural Properties. Chem. Mater. 2007, 19, 4092–4097. [Google Scholar] [CrossRef]
  12. Liu, Y.; Zhang, S.; Zhu, T. Germanium-Based Electrode Materials for Lithium-Ion Batteries. ChemElectroChem 2014, 1, 706–713. [Google Scholar] [CrossRef]
  13. Bogart, T.D.; Chockla, A.M.; Korgel, B.A. High Capacity Lithium Ion Battery Anodes of Silicon and Germanium. Curr. Opin. Chem. Eng. 2013, 2, 286–293. [Google Scholar] [CrossRef]
  14. Li, X.; Yang, Z.; Fu, Y.; Qiao, L.; Li, D.; Yue, H.; He, D. Germanium Anode with Excellent Lithium Storage Performance in a Germanium/Lithium–Cobalt Oxide Lithium-Ion Battery. ACS Nano 2015, 9, 1858–1867. [Google Scholar] [CrossRef]
  15. Sierra, L.; Gibaja, C.; Torres, I.; Salagre, E.; Avilés Moreno, J.R.; Michel, E.G.; Ocón, P.; Zamora, F. Alpha-Germanium Nanolayers for High-Performance Li-Ion Batteries. Nanomaterials 2022, 12, 3760. [Google Scholar] [CrossRef]
  16. Loaiza, L.C.; Monconduit, L.; Seznec, V. Si and Ge-Based Anode Materials for Li-, Na-, and K-Ion Batteries: A Perspective from Structure to Electrochemical Mechanism. Small 2020, 16, 1905260. [Google Scholar] [CrossRef]
  17. Liang, S.; Cheng, Y.J.; Zhu, J.; Xia, Y.; Müller-Buschbaum, P. A Chronicle Review of Nonsilicon (Sn, Sb, Ge)-Based Lithium/Sodium-Ion Battery Alloying Anodes. Small Methods 2020, 4, 200218. [Google Scholar] [CrossRef]
  18. Liu, X.; Wu, X.Y.; Chang, B.; Wang, K.X. Recent Progress on Germanium-Based Anodes for Lithium Ion Batteries: Efficient Lithiation Strategies and Mechanisms. Energy Storage Mater. 2020, 30, 146–169. [Google Scholar] [CrossRef]
  19. Kennedy, T.; Mullane, E.; Geaney, H.; Osiak, M.; O’Dwyer, C.; Ryan, K.M. High-Performance Germanium Nanowire-Based Lithium-Ion Battery Anodes Extending over 1000 Cycles through in Situ Formation of a Continuous Porous Network. Nano Lett. 2014, 14, 716–723. [Google Scholar] [CrossRef] [Green Version]
  20. Chan, C.K.; Zhang, X.F.; Cui, Y. High Capacity Li Ion Battery Anodes Using Ge Nanowires. Nano Lett. 2008, 8, 307–309. [Google Scholar] [CrossRef]
  21. Lee, G.-H.; Kwon, S.J.; Park, K.-S.; Kang, J.-G.; Park, J.-G.; Lee, S.; Kim, J.-C.; Shim, H.-W.; Kim, D.-W. Germanium Microflower-on-Nanostem as a High-Performance Lithium Ion Battery Electrode. Sci. Rep. 2014, 4, 6883. [Google Scholar] [CrossRef] [Green Version]
  22. Li, D.; Feng, C.; Liu, H.K.; Guo, Z. Hollow Carbon Spheres with Encapsulated Germanium as an Anode Material for Lithium Ion Batteries. J. Mater. Chem. A 2015, 3, 978–981. [Google Scholar] [CrossRef] [Green Version]
  23. Ngo, D.T.; Kalubarme, R.S.; Le, H.T.T.; Park, C.-N.; Park, C.-J. Conducting Additive-Free Amorphous GeO2/C Composite as a High Capacity and Long-Term Stability Anode for Lithium Ion Batteries. Nanoscale 2015, 7, 2552–2560. [Google Scholar] [CrossRef] [PubMed]
  24. Koo, J.H.; Paek, S.M. Microwave-Assisted Synthesis of Ge/GeO2-Reduced Graphene Oxide Nanocomposite with Enhanced Discharge Capacity for Lithium-Ion Batteries. Nanomaterials 2021, 11, 319. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, J.; Song, K.; Zhu, C.; Chen, C.C.; Van Aken, P.A.; Maier, J.; Yu, Y. Ge/C Nanowires as High-Capacity and Long-Life Anode Materials for Li-Ion Batteries. ACS Nano 2014, 8, 7051–7059. [Google Scholar] [CrossRef] [PubMed]
  26. Jahel, A.; Darwiche, A.; Matei Ghimbeu, C.; Vix-Guterl, C.; Monconduit, L. High Cycleability Nano-GeO2/Mesoporous Carbon Composite as Enhanced Energy Storage Anode Material in Li-Ion Batteries. J. Power Source 2014, 269, 755–759. [Google Scholar] [CrossRef]
  27. McNulty, D.; Geaney, H.; Buckley, D.; O’Dwyer, C. High Capacity Binder-Free Nanocrystalline GeO2 Inverse Opal Anodes for Li-Ion Batteries with Long Cycle Life and Stable Cell Voltage. Nano Energy 2018, 43, 11–21. [Google Scholar] [CrossRef]
  28. Yan, S.; Song, H.; Lin, S.; Wu, H.; Shi, Y.; Yao, J. GeO2 Encapsulated Ge Nanostructure with Enhanced Lithium-Storage Properties. Adv. Funct. Mater. 2019, 29, 1807946. [Google Scholar] [CrossRef]
  29. Son, Y.; Park, M.; Son, Y.; Lee, J.-S.; Jang, J.-H.; Kim, Y.; Cho, J. Quantum Confinement and Its Related Effects on the Critical Size of GeO2 Nanoparticles Anodes for Lithium Batteries. Nano Lett. 2014, 14, 1005–1010. [Google Scholar] [CrossRef]
  30. Li, W.; Li, X.; Yu, J.; Liao, J.; Zhao, B.; Huang, L.; Abdelhafiz, A.; Zhang, H.; Wang, J.H.; Guo, Z.; et al. A Self-Healing Layered GeP Anode for High-Performance Li-Ion Batteries Enabled by Low Formation Energy. Nano Energy 2019, 61, 594–603. [Google Scholar] [CrossRef]
  31. Kulova, T.L.; Skundin, A.M.; Gavrilin, I.M.; Kudryashova, Y.O.; Martynova, I.K.; Novikova, S.A. Binder-Free Ge-Co-P Anode Material for Lithium-Ion and Sodium-Ion Batteries. Batteries 2022, 8, 98. [Google Scholar] [CrossRef]
  32. Li, W.; Li, H.; Lu, Z.; Gan, L.; Ke, L.; Zhai, T.; Zhou, H. Layered Phosphorus-like GeP5: A Promising Anode Candidate with High Initial Coulombic Efficiency and Large Capacity for Lithium Ion Batteries. Energy Environ. Sci. 2015, 8, 3629–3636. [Google Scholar] [CrossRef]
  33. Li, X.; Li, W.; Shen, P.; Yang, L.; Li, Y.; Shi, Z.; Zhang, H. Layered GeP-Black P(Ge2P3): An Advanced Binary-Phase Anode for Li/Na-Storage. J. Ceram. Int. 2019, 45, 15711–15714. [Google Scholar] [CrossRef]
  34. Yan, Y.; Ruan, J.; Xu, H.; Xu, Y.; Pang, Y.; Yang, J.; Zheng, S. Fast and Stable Batteries with High Capacity Enabled by Germanium-Phosphorus Binary Nanoparticles Embedded in a Porous Carbon Matrix via Metallothermic Reduction. ACS Appl. Mater. Interfaces 2020, 12, 21579–21585. [Google Scholar] [CrossRef]
  35. Shen, H.; Huang, Y.; Chang, Y.; Hao, R.; Ma, Z.; Wu, K.; Du, P.; Guo, B.; Lyu, Y.; Wang, P.; et al. Narrowing Working Voltage Window to Improve Layered GeP Anode Cycling Performance for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 17466–17473. [Google Scholar] [CrossRef]
  36. Kulova, T.; Gryzlov, D.; Skundin, A.; Gavrilin, I.; Kudryashova, Y.; Pokryshkin, N. Anode Material Synthesized from Red Phosphorus and Germanium Nanowires for Lithium-Ion and Sodium-Ion Batteries. Int. J. Electrochem. Sci 2021, 16, 211229. [Google Scholar] [CrossRef]
  37. Nam, K.H.; Sung, G.K.; Choi, J.H.; Youn, J.S.; Jeon, K.J.; Park, C.M. New High-Energy-Density GeTe-Based Anodes for Li-Ion Batteries. J. Mater. Chem. A 2019, 7, 3278–3288. [Google Scholar] [CrossRef]
  38. Wei, Y.; Huang, L.; He, J.; Guo, Y.; Qin, R.; Li, H.; Zhai, T. Healable Structure Triggered by Thermal/Electrochemical Force in Layered GeSe2 for High Performance Li-Ion Batteries. Adv. Energy Mater. 2018, 8, 1703635. [Google Scholar] [CrossRef]
  39. Feng, D.; Liu, Q.; Li, Z.; Zeng, T. Confining Nano-GeS2 in Cross-Linked Porous Carbon Networks for High-Performance and Flexible Li-Ion Battery Anodes. ACS Appl. Energy Mater. 2021, 4, 6096–6105. [Google Scholar] [CrossRef]
  40. Zeng, T.; Feng, D.; Liu, Q.; Hao, S.; Zhou, R. Boosting Cyclability Performance of GeP Anode via In-Situ Generation of Free Expansion Volume. J. Alloys Compd. 2021, 883, 160857. [Google Scholar] [CrossRef]
  41. Tseng, K.W.; Huang, S.B.; Chang, W.C.; Tuan, H.Y. Synthesis of Mesoporous Germanium Phosphide Microspheres for High-Performance Lithium-Ion and Sodium-Ion Battery Anodes. Chem. Mater. 2018, 30, 4440–4447. [Google Scholar] [CrossRef]
  42. Kulova, T.; Gryzlov, D.; Skundin, A.; Gavrilin, I.; Martynova, I.; Kudryashova, Y. Causes of Germanium Phosphide Degradation under Prolonged Cycling. EIS Study. Int. J. Electrochem. Sci. 2022, 17, 220224. [Google Scholar] [CrossRef]
  43. Pereira, N.; Balasubramanian, M.; Dupont, L.; McBreen, J.; Klein, L.C.; Amatucci, G.G. The Electrochemistry of Germanium Nitride with Lithium. J. Electrochem. Soc. 2003, 150, A1118. [Google Scholar] [CrossRef]
  44. Nakhutsrishvili, I.G.; Dzhishiashvili, D.A.; Miminoshvili, E.B.; Mushkudiani, M.A. Preparation of Germanium Oxynitride Films in Ammonia. Inorg. Mater. 2000, 36, 1340–1341. [Google Scholar] [CrossRef]
  45. Kim, C.; Hwang, G.; Jung, J.W.; Cho, S.H.; Cheong, J.Y.; Shin, S.; Park, S.; Kim, I.D. Fast, Scalable Synthesis of Micronized Ge3N4@C with a High Tap Density for Excellent Lithium Storage. Adv. Funct. Mater. 2017, 27, 1605975. [Google Scholar] [CrossRef]
  46. Ramana, C.V.; Troitskaia, I.B.; Gromilov, S.A.; Atuchin, V.V. Electrical Properties of Germanium Oxide with α-Quartz Structure Prepared by Chemical Precipitation. Ceram. Int. 2012, 38, 5251–5255. [Google Scholar] [CrossRef]
  47. Qiu, H.; Zeng, L.; Lan, T.; Ding, X.; Wei, M. In Situ Synthesis of GeO2/Reduced Graphene Oxide Composite on Ni Foam Substrate as a Binder-Free Anode for High-Capacity Lithium-Ion Batteries. J. Mater. Chem. A 2015, 3, 1619–1623. [Google Scholar] [CrossRef]
  48. Qin, W.; Chen, T.; Hu, B.; Sun, Z.; Pan, L. GeO2 Decorated Reduced Graphene Oxide as Anode Material of Sodium Ion Battery. Electrochim. Acta 2015, 173, 193–199. [Google Scholar] [CrossRef]
  49. Ma, D.-L.; Yuan, S.; Huang, X.-L.; Cao, Z.-Y. Synthesis of Ultrathin GeO2-Reduced Graphene Oxide (RGO) Sheets for a High-Capacity Lithium-Ion Battery Anode. Energy Technol. 2014, 2, 342–347. [Google Scholar] [CrossRef]
  50. Medvedev, A.G.; Mikhaylov, A.A.; Grishanov, D.A.; Yu, D.Y.W.; Gun, J.; Sladkevich, S.; Lev, O.; Prikhodchenko, P.V. GeO2 Thin Film Deposition on Graphene Oxide by the Hydrogen Peroxide Route: Evaluation for Lithium-Ion Battery Anode. ACS Appl. Mater. Interfaces 2017, 9, 9152–9160. [Google Scholar] [CrossRef]
  51. Lv, D.; Gordin, M.L.; Yi, R.; Xu, T.; Song, J.; Jiang, Y.B.; Choi, D.; Wang, D. GeOx/Reduced Graphene Oxide Composite as an Anode for Li-Ion Batteries: Enhanced Capacity via Reversible Utilization of Li2O along with Improved Rate Performance. Adv. Funct. Mater. 2014, 24, 1059–1066. [Google Scholar] [CrossRef]
  52. Qin, J.; Cao, M. Multidimensional Germanium-Based Materials as Anodes for Lithium-Ion Batteries. Chem.-Asian J. 2016, 11, 1169–1182. [Google Scholar] [CrossRef]
  53. Sladkevich, S.; Mikhaylov, A.A.; Prikhodchenko, P.V.; Tripol’skaya, T.A.; Lev, O. Antimony Tin Oxide (ATO) Nanoparticle Formation from H2O2 Solutions: A New Generic Film Coating from Basic Solutions. Inorg. Chem. 2010, 49, 9110–9112. [Google Scholar] [CrossRef]
  54. Sladkevich, S.; Gun, J.; Prikhodchenko, P.V.; Gutkin, V.; Mikhaylov, A.A.; Medvedev, A.G.; Tripol’skaya, T.A.; Lev, O. The Formation of a Peroxoantimonate Thin Film Coating on Graphene Oxide (GO) and the Influence of the GO on Its Transformation to Antimony Oxides and Elemental Antimony. Carbon 2012, 50, 5463–5471. [Google Scholar] [CrossRef]
  55. Grishanov, D.A.; Mikhaylov, A.A.; Medvedev, A.G.; Gun, J.; Prikhodchenko, P.V.; Xu, Z.J.; Nagasubramanian, A.; Srinivasan, M.; Lev, O. Graphene Oxide-Supported β-Tin Telluride Composite for Sodium- and Lithium-Ion Battery Anodes. Energy Technol. 2018, 6, 127–133. [Google Scholar] [CrossRef]
  56. Mikhaylov, A.A.; Medvedev, A.G.; Tripol’skaya, T.A.; Popov, V.S.; Mokrushin, A.S.; Krut’ko, D.P.; Prikhodchenko, P.V.; Lev, O. H2O2 Induced Formation of Graded Composition Sodium-Doped Tin Dioxide and Template-Free Synthesis of Yolk–Shell SnO2 Particles and Their Sensing Application. Dalton Transact. 2017, 46, 16171–16179. [Google Scholar] [CrossRef]
  57. Mikhaylov, A.A.; Medvedev, A.G.; Grishanov, D.A.; Sladkevich, S.; Gun, J.; Prikhodchenko, P.V.; Xu, Z.J.; Nagasubramanian, A.; Srinivasan, M.; Lev, O. Vanadium Oxide Thin Film Formation on Graphene Oxide by Microexplosive Decomposition of Ammonium Peroxovanadate and Its Application as a Sodium Ion Battery Anode. Langmuir 2018, 34, 2741–2747. [Google Scholar] [CrossRef]
  58. Prikhodchenko, P.V.; Gun, J.; Sladkevich, S.; Mikhaylov, A.A.; Lev, O.; Tay, Y.Y.; Batabyal, S.K.; Yu, D.Y.W. Conversion of Hydroperoxoantimonate Coated Graphenes to Sb2S3 Graphene for a Superior Lithium Battery Anode. Chem. Mater. 2012, 24, 4750–4757. [Google Scholar] [CrossRef]
  59. Lakshmi, V.; Chen, Y.; Mikhaylov, A.A.; Medvedev, A.G.; Sultana, I.; Rahman, M.M.; Lev, O.; Prikhodchenko, P.V.; Glushenkov, A.M. Nanocrystalline SnS2 Coated onto Reduced Graphene Oxide: Demonstrating the Feasibility of a Non-Graphitic Anode with Sulfide Chemistry for Potassium-Ion Batteries. Chem. Commun. 2017, 53, 8272–8275. [Google Scholar] [CrossRef] [Green Version]
  60. Lakshmi, V.; Mikhaylov, A.A.; Medvedev, A.G.; Zhang, C.; Ramireddy, T.; Rahman, M.M.; Cizek, P.; Golberg, D.; Chen, Y.; Lev, O.; et al. Probing Electrochemical Reactivity in an Sb2S3-Containing Potassium-Ion Battery Anode: Observation of an Increased Capacity. J. Mater. Chem. A 2020, 8, 11424–11434. [Google Scholar] [CrossRef]
  61. Prikhodchenko, P.V.; Yu, D.Y.W.; Batabyal, S.K.; Uvarov, V.; Gun, J.; Sladkevich, S.; Mikhaylov, A.A.; Medvedev, A.G.; Lev, O. Nanocrystalline Tin Disulfide Coating of Reduced Graphene Oxide Produced by the Peroxostannate Deposition Route for Sodium Ion Battery Anodes. J. Mater. Chem. A 2014, 2, 8431–8437. [Google Scholar] [CrossRef]
  62. Mikhaylov, A.A.; Medvedev, A.G.; Mason, C.W.; Nagasubramanian, A.; Madhavi, S.; Batabyal, S.K.; Zhang, Q.; Gun, J.; Prikhodchenko, P.V.; Lev, O. Graphene Oxide Supported Sodium Stannate Lithium Ion Battery Anodes by the Peroxide Route: Low Temperature and No Waste Processing. J. Mater. Chem. A 2015, 3, 20681–20689. [Google Scholar] [CrossRef]
  63. Yu, D.Y.W.; Batabyal, S.K.; Gun, J.; Sladkevich, S.; Mikhaylov, A.A.; Medvedev, A.G.; Novotortsev, V.M.; Lev, O.; Prikhodchenko, P.V. Antimony and Antimony Oxide@graphene Oxide Obtained by the Peroxide Route as Anodes for Lithium-Ion Batteries. Main Group Metal Chem. 2015, 38, 43–50. [Google Scholar] [CrossRef]
  64. Sladkevich, S.; Gun, J.; Prikhodchenko, P.V.; Gutkin, V.; Mikhaylov, A.A.; Novotortsev, V.M.; Zhu, J.X.; Yang, D.; Hng, H.H.; Tay, Y.Y.; et al. Peroxide Induced Tin Oxide Coating of Graphene Oxide at Room Temperature and Its Application for Lithium Ion Batteries. Nanotechnology 2012, 23, 485601. [Google Scholar] [CrossRef] [PubMed]
  65. Yu, D.Y.W.; Prikhodchenko, P.V.; Mason, C.W.; Batabyal, S.K.; Gun, J.; Sladkevich, S.; Medvedev, A.G.; Lev, O. High-Capacity Antimony Sulphide Nanoparticle-Decorated Graphene Composite as Anode for Sodium-Ion Batteries. Nat. Commun. 2013, 4, 2922. [Google Scholar] [CrossRef] [Green Version]
  66. Alpen, U.V.; Rabenau, A.; Talat, G.H. Ionic Conductivity in Li3N Single Crystals. Appl. Phys. Lett. 2008, 30, 621. [Google Scholar] [CrossRef]
  67. Nazri, G. Preparation, Structure and Ionic Conductivity of Lithium Phosphide. Solid State Ion. 1989, 34, 97–102. [Google Scholar] [CrossRef]
  68. Hayashi, A.; Iio, K.; Morimoto, H.; Minami, T.; Tatsumisago, M. Mechanochemical Synthesis of Amorphous Solid Electrolytes Using SiS2 and Various Lithium Compounds. Solid State Ion. 2004, 175, 637–640. [Google Scholar] [CrossRef]
  69. Miyamura, M.; Tomura, S.; Imai, A.; Inomata, S. Electrochemical Studies of Lithium Nitride Solid Electrolyte for Electrochromic Devices. Solid State Ion. 1981, 3–4, 149–152. [Google Scholar] [CrossRef]
  70. Nazri, G. Preparation, Characterization and Conductivity of Li3N, Li3P and Li3As. MRS Proc. 1988, 135, 117–130. [Google Scholar] [CrossRef]
  71. Grishanov, D.A.; Churakov, A.V.; Medvedev, A.G.; Mikhaylov, A.A.; Lev, O.; Prikhodchenko, P.V. Crystalline Ammonium Peroxogermanate as a Waste-Free, Fully Recyclable Versatile Precursor for Germanium Compounds. Inorg. Chem. 2019, 58, 1905–1911. [Google Scholar] [CrossRef]
  72. Kansuzyan, A.V.; Farafonova, S.D.; Saverina, E.A.; Krylova, I.V.; Balycheva, V.A.; Akyeva, A.Y.; Medvedev, A.G.; Nikolaevskaya, E.N.; Egorov, M.P.; Prikhodchenko, P.V.; et al. Highly Soluble Germanium Dioxide as a New Source of Germanium for Derivatization with Organic Compounds. Mend. Commun. 2022, 32, 25–27. [Google Scholar] [CrossRef]
  73. Uvarov, V.; Popov, I. An Estimation of the Correctness of XRD Results Obtained from the Analysis of Materials with Bimodal Crystallite Size Distribution. CrystEngComm 2015, 17, 8300–8306. [Google Scholar] [CrossRef]
  74. Wei, W.; Guo, L. One-Step In Situ Synthesis of GeO2/Graphene Composites Anode for High-Performance Li-Ion Batteries. Part. Part. Syst. Charact. 2013, 30, 658–661. [Google Scholar] [CrossRef]
  75. Yu, L.; Wang, L.P.; Liao, H.; Wang, J.; Feng, Z.; Lev, O.; Loo, J.S.C.; Sougrati, M.T.; Xu, Z.J. Understanding Fundamentals and Reaction Mechanisms of Electrode Materials for Na-Ion Batteries. Small 2018, 14, 1703338. [Google Scholar] [CrossRef]
  76. Li, L.; Zheng, Y.; Zhang, S.; Yang, J.; Shao, Z.; Guo, Z. Recent Progress on Sodium Ion Batteries: Potential High-Performance Anodes. Energy Environ. Sci. 2018, 11, 2310–2340. [Google Scholar] [CrossRef] [Green Version]
  77. Lee, Y.T.; Kuo, C.T.; Yew, T.R. Investigation on the Voltage Hysteresis of Mn3O4 for Lithium-Ion Battery Applications. ACS Appl. Mater. Interfaces 2021, 13, 570–579. [Google Scholar] [CrossRef]
  78. Wang, J.; Wang, L.; Eng, C.; Wang, J. Elucidating the Irreversible Mechanism and Voltage Hysteresis in Conversion Reaction for High-Energy Sodium–Metal Sulfide Batteries. Adv. Energy Mater. 2017, 7, 1602706. [Google Scholar] [CrossRef]
  79. Mikhaylov, A.A.; Medvedev, A.G.; Buldashov, I.A.; Fazliev, T.M.; Mel’nik, E.A.; Tripol’skaya, T.A.; Sladkevich, S.; Nikolaev, V.A.; Lev, O.; Prikhodchenko, P.V. Green Synthesis of Zinc Sulfide-Reduced Graphene Oxide Composite and Its Application in Sodium-Ion Batteries. J. Alloys Compd. 2022, 910, 164769. [Google Scholar] [CrossRef]
  80. Zhong, K.; Xia, X.; Zhang, B.; Li, H.; Wang, Z.; Chen, L. MnO Powder as Anode Active Materials for Lithium Ion Batteries. J. Power Source 2010, 195, 3300–3308. [Google Scholar] [CrossRef]
Figure 1. Powder X-ray diffractograms of GeO2-rGO-20 (a), GeO2-rGO-50 (b), GeO2-rGO-80 (c), Ge3N4-rGO-20 (d), Ge3N4-rGO-50 (e), Ge3N4-rGO-80 (f), GeP-rGO-20 (g), GeP-rGO-50 (h), and GeP-rGO-80 (i) powders. Vertical bars below represent the standard diffraction data for corresponding crystalline germanium compounds.
Figure 1. Powder X-ray diffractograms of GeO2-rGO-20 (a), GeO2-rGO-50 (b), GeO2-rGO-80 (c), Ge3N4-rGO-20 (d), Ge3N4-rGO-50 (e), Ge3N4-rGO-80 (f), GeP-rGO-20 (g), GeP-rGO-50 (h), and GeP-rGO-80 (i) powders. Vertical bars below represent the standard diffraction data for corresponding crystalline germanium compounds.
Ijms 24 06860 g001
Figure 2. SEM images of GeO2-rGO-80 (a,b), Ge3N4-rGO-80 (c,d), and GeP-rGO-80 (e,f).
Figure 2. SEM images of GeO2-rGO-80 (a,b), Ge3N4-rGO-80 (c,d), and GeP-rGO-80 (e,f).
Ijms 24 06860 g002
Figure 3. Cycling voltammetry of GeO2-rGO-80 (a), Ge3N4-rGO-80 (b), and GeP-rGO-80 (c) at 0.1 mV s−1. First and second charge−discharge curves for GeO2-rGO-80 (d), Ge3N4-rGO-80 (e), and GeP-rGO-80 (f) at 100 mA g−1.
Figure 3. Cycling voltammetry of GeO2-rGO-80 (a), Ge3N4-rGO-80 (b), and GeP-rGO-80 (c) at 0.1 mV s−1. First and second charge−discharge curves for GeO2-rGO-80 (d), Ge3N4-rGO-80 (e), and GeP-rGO-80 (f) at 100 mA g−1.
Ijms 24 06860 g003
Figure 4. Cycling voltammetry of GeO2-rGO-80 (a), Ge3N4-rGO-80 (b), and GeP-rGO-80 (c) at 0.1 mV s−1. First and second charge−discharge curves for GeO2-rGO-80 (d), Ge3N4-rGO-80 (e), and GeP-rGO-80 (f) at 100 mA g−1.
Figure 4. Cycling voltammetry of GeO2-rGO-80 (a), Ge3N4-rGO-80 (b), and GeP-rGO-80 (c) at 0.1 mV s−1. First and second charge−discharge curves for GeO2-rGO-80 (d), Ge3N4-rGO-80 (e), and GeP-rGO-80 (f) at 100 mA g−1.
Ijms 24 06860 g004
Table 1. Elemental analysis of GeO2-rGO, Ge3N4-rGO, and GeP-rGO materials.
Table 1. Elemental analysis of GeO2-rGO, Ge3N4-rGO, and GeP-rGO materials.
MaterialC, wt.% (by CHN)N, wt.% (by CHN)Ge:P, at:at (by EDX)
GeO2-rGO-8016.7
GeO2-rGO-5042.9
GeO2-rGO-2075.5
Ge3N4-rGO-8025.012.0
Ge3N4-rGO-5048.09.4
Ge3N4-rGO-2076.26.8
GeP-rGO-8013.5 1.1
GeP-rGO-5045.1 0.9
GeP-rGO-2074.3 1.2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mikhaylov, A.A.; Medvedev, A.G.; Grishanov, D.A.; Fazliev, T.M.; Chernyshev, V.; Mel’nik, E.A.; Tripol’skaya, T.A.; Lev, O.; Prikhodchenko, P.V. Electrochemical Behavior of Reduced Graphene Oxide Supported Germanium Oxide, Germanium Nitride, and Germanium Phosphide as Lithium-Ion Battery Anodes Obtained from Highly Soluble Germanium Oxide. Int. J. Mol. Sci. 2023, 24, 6860. https://doi.org/10.3390/ijms24076860

AMA Style

Mikhaylov AA, Medvedev AG, Grishanov DA, Fazliev TM, Chernyshev V, Mel’nik EA, Tripol’skaya TA, Lev O, Prikhodchenko PV. Electrochemical Behavior of Reduced Graphene Oxide Supported Germanium Oxide, Germanium Nitride, and Germanium Phosphide as Lithium-Ion Battery Anodes Obtained from Highly Soluble Germanium Oxide. International Journal of Molecular Sciences. 2023; 24(7):6860. https://doi.org/10.3390/ijms24076860

Chicago/Turabian Style

Mikhaylov, Alexey A., Alexander G. Medvedev, Dmitry A. Grishanov, Timur M. Fazliev, Vasilii Chernyshev, Elena A. Mel’nik, Tatiana A. Tripol’skaya, Ovadia Lev, and Petr V. Prikhodchenko. 2023. "Electrochemical Behavior of Reduced Graphene Oxide Supported Germanium Oxide, Germanium Nitride, and Germanium Phosphide as Lithium-Ion Battery Anodes Obtained from Highly Soluble Germanium Oxide" International Journal of Molecular Sciences 24, no. 7: 6860. https://doi.org/10.3390/ijms24076860

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop