Next Article in Journal
Evolutionary Game and Simulation Analysis of Power Plant and Government Behavior Strategies in the Coupled Power Generation Industry of Agricultural and Forestry Biomass and Coal
Next Article in Special Issue
Photosensitivity of Nanostructured Schottky Barriers Based on GaP for Solar Energy Applications
Previous Article in Journal
Developing an Appropriate Energy Trading Algorithm and Techno-Economic Analysis between Peer-to-Peer within a Partly Independent Microgrid
Previous Article in Special Issue
Online Multiphase Flow Measurement of Crude Oil Properties Using Nuclear (Proton) Magnetic Resonance Automated Measurement Complex for Energy Safety at Smart Oil Deposits
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase-Homogeneous LiFePO4 Powders with Crystallites Protected by Ferric-Graphite-Graphene Composite

1
Department of Electrochemical Production Technology, St. Petersburg State Institute of Technology, Moskovski Ave. 26, 190013 St. Petersburg, Russia
2
Division of Solid State Physics, Ioffe Institute, Politekhnicheskaya Str. 26, 194021 St. Petersburg, Russia
3
Institute of Chemistry, Saratov State University, Astrakhanskaya Str. 83, 410012 Saratov, Russia
4
Research Park, RC XRD, St. Petersburg State University, Universitetskaya nab. 7–9, 199034 St. Petersburg, Russia
5
RnD Center TFTE, Politekhnicheskaya Str. 26, 194021 St. Petersburg, Russia
6
Department “Chemistry and Chemical Technology”, Petrochemical Institute, Omsk State Technical University, Mira Ave. 11, 644050 Omsk, Russia
7
Department of Electronics, St. Petersburg State Electrotechnical Univeristy, ul. Professora Popova 5, 197022 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Energies 2023, 16(3), 1551; https://doi.org/10.3390/en16031551
Submission received: 27 December 2022 / Revised: 26 January 2023 / Accepted: 1 February 2023 / Published: 3 February 2023
(This article belongs to the Special Issue Nuclear and New Energy Technology)

Abstract

:
Phase-homogeneous LiFePO4 powders have been synthesized. The content of impurity crystalline phases was less than 0.1%, according to synchrotron diffractometry (SXRD) data. Anisotropic crystallite sizes L ¯ V h k l were determined by XRD. A low resistance covering layer of mechanically strong ferric-graphite-graphene composite with impregnated ferric (Fe3+) particles < 10 nm in size increases the cycleability compared to industrial cathodes. In accordance with the corrosion model, the destruction of the Fe3+-containing protective layer of crystallites predominates at the first stage, and at the second stage Fe escapes into the electrolyte and to the anode. The crystallite size decreases due to amorphization that starts from the surface. The rate capability, Q(t), has been studied as a function of L ¯ V h k l , of the correlation coefficients r i k between crystallite sizes, of the Li diffusion coefficient, D, and of the electrical relaxation time, τel. For the test cathode with a thickness of 8 μm, the values of D = 0.12 nm2/s, τel = 8 s were obtained. To predict the dependence Q(t), it is theoretically studied in ranges closest to experimental values: D = 0.5 ÷ 0.03 nm2/s, τel = 8/1 s, average sizes along [010] L ¯ 1 = 90/30 nm, averaged r ¯ = 0/1.

1. Introduction

The values of the battery capacity, Q0, the rate capability, Q(t), and the maximum number of discharge-charge cycles are important target battery parameters. As shown in [1], the phase homogeneity of LiFePO4 suffers from the impurity phases which appear as synthesis residues at low temperatures (<550 °C) and ferric (Fe3+) compounds at high temperatures in the presence of residual oxygen. The first-principal modelling of Li-Fe-P-O2 phase diagrams [2] supported that.
Several attempts have been made to develop a technology for producing highly efficient and phase-homogeneous electrode powders using: various raw materials and processing methods [3,4], variations in the composition and proportions of the loaded raw materials [5,6,7], growth duration and additional multi-stage post-growth annealing [8,9], extra pure initial chemicals of so-called “battery quality” [10], one-pot synthesis [11,12], etc. Coating of crystallites with various functional layers was also found useful [13]: metal oxides, glasses, etc. and the most popular carbon-based materials [14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31] were tried as the coating substances. The latter include various carbon phases [14], nanocarbons of different shapes (spheres, tubes and pores) [15], graphene [16,17,18]. Sucrose [19,20,21], glucose [22,23], adipic acid [24,25,26], polyvinyl alcohol [27,28], polymeric additives [29,30], and ferrocene [31] were used as catalysts for graphitization.
Our approach is based on (i) one-pot LiFePO4 liquid-phase synthesis from high purity lithium and iron acetates taken as the starting materials, (ii) multi-stage thermal processing including acetic acid evaporation at its boiling temperature and protective coating formation from adipic acid and polyvinyl alcohol, (iii) elimination of potential impurity sources, (iv) stage-by-stage control of the effect of the regime parameters on crystallite sizes and coating quality.
X-ray diffractometry (XRD) is commonly used to characterize the presence and size of crystallites in electrode powders. Its synchrotron version (SXRD) [32,33,34] is the most sensitive to the impurities in crystallites. Various surface-sensitive methods are used to analyze the LiFePO4 particle surface [19,35,36,37,38]: Auger electron spectroscopy (AES), electron diffraction (EDS) [39], X-ray photoelectron spectroscopies (XPS) [40,41,42], inductively coupled plasma (ICP) [39].
We used both synchrotron and laboratory X-ray sources SXRD and XRD, respectively, to increase the sensitivity towards the presence of impurity phases and to determine anisotropic size distributions of LiFePO4 crystallites, as well as to study degradation during cycling of test cathodes.
Mössbauer spectroscopy (MS) makes it possible to detect iron-containing compounds and study their properties [43,44,45,46,47,48,49]. In particular, dependence of hyperfine spectral parameters on the charge of Fe ions [45,46,47,48] have been used to study reversible delithiation processes in LiFePO4 crystallites. Line broadening associated with an increase in the disorder around Fe2+ towards the crystallite surface was observed; oxidation suppression by surface carbon was also found [45]. In this work, the MS results were used for determination of the proportion of the Fe3+ ferric impurity compounds in LiFePO4 (which is itself a Fe2+ compound).
To reveal the role of ferric compounds, the following results are important: (1) up to 5% ferric states are typically present at regular lattice sites with reduced symmetry [50], which is not detected by XRD; (2) lowering the content of X-ray inactive ferric compounds from 17% to 9% improves the electrochemical parameters, even compared with those materials where the content was zero [51]; (3) presence of significant amounts of Fe3+ cations at broad interfaces improves electronic and ionic conductivity [52]; (4) a significant number of ferric states was detected on the surface of crystallites with MS and XPS [53]; (5) MS and ICP studies showed that an increase in the synthesis time led to an increase in the fraction of iron atoms with a long-range order compared to the fraction of iron atoms in the Li/Fe-PO4 amorphous phase [54]; (6) initial powders contained ferric compounds Li3Fe2(PO4)3 and α-Fe2O3, which became X-ray inactive after 6 h annealing at 700 °C [55]; (7) after low-temperature precipitation at 106 °C, the powder contained an amorphous ferric phase LiFePO4(OH), which transformed into crystalline LiFePO4 after annealing under reducing conditions; after annealing at 500 °C, the crystallite size increased from 17 nm to 35 nm [56]; (8) with differential scanning calorimetry (DSC) and derivative thermogravimetry (DTG), small particles of Li3Fe2(PO4)3 and presumably of Fe2O3 were detected on the surface of LiFePO4 crystallites after their oxidation below 470–475 °C [57].
We used Mössbauer spectroscopy, EDX and XRD studies to elucidate the non-trivial role of ferric compounds, in particular, in the protective layer formation and its degradation.
The combined use of transmission electron microscopy (TEM) and Raman spectroscopy (RS) is important in study of the various carbon phases in the cathode materials initial powders [58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80]. Using RS, it was found that some carbon materials mixtures have a unique combination of properties: mechanical characteristics, strength, chemical inertness and biocompatibility [61,62,63]. Among a large number of different grades of carbon black [65], the most famous is Ketjen black [70], which is widely used in electrode powder technologies [71,72,73,74,75,76] and is added especially when it is necessary to sharply increase the electronic conductivity [75,76].
The novelty and importance of Raman spectroscopies line intensity analysis lies in the study of the graphene phase’s order degree, the graphite phase mechanical strength, and the multilayer graphene phase conductivity compared to industrial LiFePO4 powders. Thus, the development of phase-homogeneous electrode powders may further increase battery capacity and cyclability, and optimize the rate capability.
In this work we synthesized LiFePO4 powders and coatings on the surface of their crystallites. The phase homogeneity was tested with SXRD, and its effect on the test cathodes cyclability was shown. Composition and properties of the layers covering the crystallites was studied using TEM and RS. MS was used to discover the role of ferric compounds, and XRD and EDX were used to study the mechanisms of degradation of the LiFePO4 test cathodes. LiFePO4 test cathodes were subjected to galvanostatic measurements. The hierarchy of relaxation times of electrochemical recharging of the cathodes was determined using the data on anisotropic size distribution of crystallites. We studied the dependence of Q(t) of the test cathodes on the crystallite parameters, the lithium diffusion coefficient D along [010] and the quality of their coating in terms of electric relaxation time τel. Q(t) calculations allowed us to assess the possibility of improving the technology.

2. Synthesis of LiFePO4 Powders and Surface Coatings, and Characterization Techniques

2.1. Synthesis of LiFePO4 Powder and Coatings

We used a modified version of the liquid-phase synthesis of LiFePO4 based on lithium and iron acetates as initial reagents [7,81]. They have good solubility in water and low thermal stability at the synthesis temperatures; acetic acid has a low boiling point which ensures its distillation during synthesis, and volatile components can be easily removed from the reaction zone during annealing [82,83,84,85,86,87,88,89]. Initial acetates were prepared from metallic iron and lithium carbonate by interaction with acetic acid (Snabtechmet), grade A.C.S. These materials are available and have a high degree of purity for large-scale production. Chemicals were used as received from the manufacturers; additional reduction of atmospheric impact was provided at the stages of pre-drying and annealing. Mechanochemical activation of the starting materials in a liquid medium promoted formation of an intermediate synthesis product, which was converted into LiFePO4 at a temperature lower by 100 °C than in the standard procedures. This created the prerequisites for obtaining LiFePO4 with a reduced content of impurity phases [87]. Mechanochemical activation was carried out in a saturated solution of ammonium dihydroorthophosphate in distilled water. Table 1 shows the following synthesis steps:
1. Lithium acetate is obtained by direct acetic acid action on lithium carbonate. The Fe2+ ion is easily oxidized by atmospheric oxygen to Fe3+, therefore, iron acetate is prepared by placing a calculated amount of iron in a flask containing an excess of acetic acid. The flask is stoppered and kept under vacuum until dissolved with stirring with a magnetic stirrer and heated at the end of the process.
2. Two organic additives, adipic acid (AA) [88,89,90,91] and polyvinyl alcohol (PA) [92,93,94], are used to obtain and condition the synthesized composite. AA crystallizes upon reaction mixture evaporation and decomposes upon heat treatment, resulting in a precipitate with a loosened structure. PA is adsorbed on the precipitate surface and contributes to the small crystals formed during the interaction of phosphoric acid and lithium and iron acetates.
3. Phosphoric acid is used as a source of phosphorus ions.
4. Pre-drying to evaporate CH3COOH and H2O in a stream of hot air, which prevented agglomeration of the starting materials.
5–6. Annealing for evaporation at different temperatures and crystallization. After drying in a stream of hot air, the sample was pelletized and placed in a sealed muffle furnace, which was constantly purged with especially pure nitrogen. The furnace was heated to 400 °C, the sample was held at this temperature for 1.5 h, then the heating was turned off and the furnace was expected to cool in a nitrogen flow to a temperature of 25 °C. The resulting intermediate was subjected to repeated grinding and tableting. This approach does not lead to the formation of the final product due to the low temperature, but it makes it possible to obtain a dense mixture of initial substances with a large interfacial surface and a fixed contact between the phases of the initial substances. Then the tableted sample is placed in a sealed muffle furnace, which is constantly purged with especially pure nitrogen; the temperature in the furnace rises to 670 °C. This approach is tested in a series of experiments near this temperature. The selected temperature regime provides the maximum LiFePO4 capacity of 65 mAh/g at 20 C rate. At all stages of preparation and synthesis, the above measures are taken, in particular, to exclude the transition of Fe2+ to Fe3+.
Preliminary technological experiments were also carried out, in particular using: iron acetate (CH3COO)2Fe, iron oxalate FeC2O4 × 2H2O, lithium carbonate Li2CO3, ammonium dihydroorthophosphate (NH4H2PO4) from VEKTON, grade A.C.S. Depending on the composition and the annealing modes, the following impurity crystalline phases were observed in the samples: Fe3O4 (ICDD 00-019-0629), Fe2O3 (ICDD 00-039-1346), FeCO3 (ICDD 00-029-0696), FePO4 (ICDD 00-050-1635). To compare the quality and target parameters of the developed sample N1, SPbTU, the following industrial LiFePO4 powders were used: N2, Phostech Lithium [95]; N3 OCELL Technologies N4, Golden Light Energy. The powders had specific capacities ranging from 145 mAh/g to 167 mAh/g at 0.1 C rate [7,81], given in the Table 2 at 20 C rate.
The novelty of our synthesis in comparison with [87,88,89] is the following set:
  • One-pot LiFePO4 liquid-phase synthesis using chemically pure lithium and iron acetates as starting materials.
  • Using raw materials of organic nature.
  • Using the tableting operation after the pyrolysis of organic matter, at 400 °C. This operation facilitates the course of the final synthesis of the topochemical reaction.
  • Drying, before heat treatment, was carried out in a stream of hot air, which prevented the agglomeration of the starting materials.
  • The synthesis intermediate is tableted, heat-treated at 400 °C, then re-milled and re-tableted before the final heat treatment at 670 °C. This made it possible to obtain a material of high phase purity.
  • Before repeated tableting, adipic acid was introduced, the pyrolysis of which in an inert medium (high-purity nitrogen) led to encapsulation of LiFePO4 in a carbon shell.
  • At all stages of the preparation, measures were taken to exclude the transition of Fe2+ to Fe3+, in particular, by isolation from atmospheric moisture.

2.2. Characterization Techniques

Note that conventional XRD is less sensitive than SXRD due to the lower intensity of laboratory radiation sources [32,33,34,96]. The SXRD experiments were done at the Structural Materials Science station of the Siberia-2 synchrotron radiation source of Kurchatov Institute Research Center [96]. Measurements were made at a wavelength of λ = 0.68886 Å in the transmission (Debye–Scherrer) mode with a Fujifilm Imaging Plate memory layer as a 2D detector, sample-to-detector distance of 200 mm, and exposure time of 15 min.
The anisotropic crystallite sizes L ¯ V h k l averaged over the length of the columns [97] were determined by XRD [98]. Bruker D8 Discover diffractometer was used in a parallel-beam linear-focus mode at 2θ = 15–125 deg, and MAUD software was utilized for profile fitting [94]. The primary beam was conditioned with a double-bounce channel-cut Ge220 monochromator to provide CuKa1 radiation with a wavelength of 1.54056 Å. The Cagliotti coefficients of the instrumental profile function were refined by fitting the data for a LaB6 powder specimen (NIST SRM 660c).
The powders were studied with a JEOL JEM 2100 high-resolution TEM at an accelerating voltage of 200 kV and crystal lattice resolution of 0.14 nm. The instrument was equipped with an INKA 250 Xray spectrometer. An image of the single-crystal gold lattice with the (111) interplanar spacing of 0.235 nm was used as reference for linear scale calibration.
Raman spectra were measured at room temperature in the “backscattering” geometry on a LabRam HR 800 spectrometer equipped with a confocal microscope. The measurements used the exciting light wavelengths of 532 nm and 633 nm, focused on the surface of the sample into a spot with a diameter of ~1 μm. In this case, the laser radiation power on the sample was maintained at a level of 2.0 mW. The use of a 600 pcs/mm diffraction grating made it possible to obtain a spectral resolution no worse than 2.5 cm−1.
The Mössbauer effect was measured at room temperature on 57Fe nuclei in the γ-ray transmission geometry through powders sputtered onto aluminum foil with a spot diameter of 20 mm. The movement of the 57Co(Rd) γ-radiation source in the spectrometer was carried out with a constant acceleration of the reference signal in the form of a triangle. Velocity calibration was performed using α-iron foil for two Doppler shift velocities of the gamma-ray source.

3. LiFePO4 Powders Phase Homogeneity Studies Using SXRD and Its Effect on the Test Cathodes Cyclability

3.1. Phase Homogeneity Using SXRD

The results are shown in Figure 1 and the quantitative composition of impurity crystalline phases are given in Table 2.

3.2. Cycling Test Cathode Cells

A test electrode was prepared of 80 wt.% powder, 10 wt.% acetylene black and 10 wt.% polyvinylidene fluoride (PVDF). It was applied as a sample homogenized suspension of acetylene black in a 5 wt.% solution of PVDF in N- methylpyrrolidone (analytical grade) on 1 cm2 aluminum plate, 0.4 mm thick, after which it was dried at 120 °C in air for 12 h [99]. To reduce the errors of the galvanostatic measurements of rate capability at small times, the cathode thickness was minimal, about 8 μm. The measurements included charge-discharge cycles with constant current loads: sequentially from 0.1 to 20 C rates in one cycle, then 15 cycles each, completing 10 cycles with a load of 1 rate. The current density of 1 C rate corresponded to 170 μA per 1 mg of sample and was analyzed at the potentials in the range of 2.6–4.3 V relative to the lithium reference electrode. Figure 2 shows the galvanostatic measurements results. Cycling predictions were obtained by making 150 charge-discharge cycles and by recalculating to the point of the capacitance reduction to 80 % of its initial value at a current of 1 C rate indicated in Table 2.
The following conclusions can be drawn from the results in Table 2:
  • No impurity crystallites were found in the samples N1 and N4. However, the cyclability of the latter was significantly worse than that of the developed sample N1, despite the trend towards an increase in the role of the [100] surface of larger crystallites in their cyclability. Consequently, the absence of impurity crystallites is a necessary but obviously not a sufficient condition.
  • Sample N1 had a high cyclability, but lower than that in the sample N3, which contained several impurity phases. In addition to lowering the resistance, these phases probably catch the degradation products, which slows down the formation of harmful impurity phases. It can be concluded that it is promising to search for such compositions or isovalent doping, including various mixed solid solutions, starting from our pure technology.

4. Crystallite Coating Composition and Properties Studied with TEM and RS

4.1. TEM Study

TEM images of LiFePO4 powders are shown in Figure 3. Several types of particles are observed in the powders:
-
Large 40–150 nm particles of LiFePO4 are observed in all TEM images; an example is shown in Figure 3a for the developed sample N1.
-
In almost all samples, nanocrystalline 5–10 nm Li3Fe2(PO4)3 particles are observed (Figure 3b,e,f) with lattice spacing of 0.428 nm, which corresponds to (200) or (−121) planes. It should be noted, however, that the (011) planes of LiFePO4 have a similar interplanar distance. The ferric compound Li3Fe2(PO4)3 on the surface of LiFePO4 crystallites appears as a result of insufficient oxygen content to complete the oxidation reaction.
-
Figure 3f shows crystallites with interplanar spacing of 0.220 nm that corresponds presumably to the (321) planes of Fe3P in the N3 sample, even though a similar distance can be found in the structures of other phases. According to [91,101,102], at T > 850 °C and in the presence of carbon, LiFePO4 is reduced to form Fe3P. As seen from Table 2, a significant amount of that phase is reliably detected in the sample N3 using SXRD; the plate-like shape of the particles was described in detail in [103].
-
Various structures of the carbon layers encapsulating the LiFePO4 particles can be seen in the samples. More ordered carbon shells are up to 5 nm thick and the amorphous shells are up to 20 nm thick. In some cases, particles without a carbon shell are observed. The properties of the carbon coatings will be discussed below in the RS section.
Figure 3. (ad)—TEM images of the developed N1 powder of 3 crystallites (a); (b) is an enlarged part of (a). Areas of the ferric-graphite-graphene composite coating layer are marked (b); FT snapshot obtained from this TEM image (c), a maximum close to the interplanar spacing of 0.428 nm (d). (e,f)—TEM images of industrial powders: N4 (e) and N3 (f) containing the impurity crystallite phase of Fe3P. Regions characteristic of multilayer graphene and the most probable Li3Fe2(PO4)3 are also marked.
Figure 3. (ad)—TEM images of the developed N1 powder of 3 crystallites (a); (b) is an enlarged part of (a). Areas of the ferric-graphite-graphene composite coating layer are marked (b); FT snapshot obtained from this TEM image (c), a maximum close to the interplanar spacing of 0.428 nm (d). (e,f)—TEM images of industrial powders: N4 (e) and N3 (f) containing the impurity crystallite phase of Fe3P. Regions characteristic of multilayer graphene and the most probable Li3Fe2(PO4)3 are also marked.
Energies 16 01551 g003aEnergies 16 01551 g003b
In none of the samples have particles of Fe2O3 been found. Perhaps they can be observed in other technologies, considering also some identification uncertainty with Li3Fe2(PO4)3, so the coating we observed was called a ferric-graphite-graphene composite.

4.2. Raman Spectroscopy Studies of Carbon Phases

Figure 4 shows the Raman spectra of the Sample N1 taken in the ranges from 800 to 3700 cm−1 [104].
The RS spectrum analysis is performed by the component separation method using a Lorentz line shape [105,106]. To improve reliability, two close excitation lines 532 nm and 633 nm are also used, since it is expected that the main conclusions from the results of comparing the line amplitudes should coincide in two measurement series. Table 3 includes all decomposition parameters useful for comparison between this and other studies, especially the second order linewidths.
According to [58,59,60], two lines at 1518 cm−1 and 1201 cm−1 have already been observed in disordered carbon black and diamond-like carbons. That could imply that the short-range vibrations of the sp3-coordinated carbons contribute to the disordered spectra. Unique sp2/sp3 nanohybrids as bulky nanodiamonds (NDs) and sp2 concentric onion-like carbons (OLC) [61,62] with outstanding mechanical performance, sufficient chemical inertness, excellent biocompatibility, high mechanical strength [61,62,63] are possible.
To interpret the lines in Figure 4, we use the sequence proposed in [64]. The spectra show lines arising from light scattering in spectral regions close to their position in the spectra of multilayer graphene or graphene-like layers: lines D (E2), G (E4), 2D (E5), D + G (E6), 2D′ (E7) [64,65,66]. Line G in the spectrum of graphene corresponds to nonresonant light scattering involving an optical phonon of E2g symmetry with a small wave vector. This phonon is caused by vibrations of carbon atoms in the layer plane. The appearance of the D line in the spectra is explained by resonant scattering involving electronic states from two nonequivalent K- and K′-points of the Brillouin zone and an optical phonon with a large wave vector. This process is forbidden by the quasi-momentum selection rule, but the condition for its conservation can be satisfied if the crystal lattice defect also participates in the scattering process. In structurally perfect graphene samples, line D should not be observed. The second order spectra in Figure 4b are markedly broadened. The nature of the 2D line, an overtone of the D line, is also associated with resonant light scattering involving electronic states. The quasi-momentum conservation conditions for such a process are always satisfied, so the 2D line will be present in the graphene spectrum even if it does not contain the D line. The combination D + G corresponds to a defect-induced double resonance “inter-valley” scattering process which is allowed through a defect-induced triple resonance process. The second order spectra lines are observed in graphene oxide, GO, [67,68], which is produced through graphite chemical oxidation and subsequent exfoliation via sonication, and in various modifications of carbon black (CB) powders [69].
It should be noted that the positions and widths of the observed second ordered three lines in the range 1900–3700 cm−1 correspond to those described in [70,71,72,77,78,79]. In particular, the papers [74,79] describe the dependence of the position of the 2D line, with a maximum of E5, on the magnitude of the deformation. At the same time, its width in Figure 4b is almost 10 times larger than that described in [64,65,66] and is comparable to its displacement at deformations of about 1%, which indicates the presence of a significant deformation disorder in the studied samples compared to multilayer graphene.
The carbon fraction diagnostics using Raman spectroscopy data is based on an analysis of the position, the width, and the intensity ratio of the observed lines. In our case, the integral intensity is obtained from the results of decomposition into components. The best known are the intensity ratios ID/IG and Isp2/Isp3 [64,65,66], in the used designations I2/I4 and (I2 + I4)/(I1 + I3), respectively. Obviously, the ratios of the total line intensities of the first and second order can also be useful, i.e., (I1 + …I4)/(I5 + …I7) [68]. Comparing the ratios of the line intensities, we can draw the following conclusions:
For the developed sample, the ID/IG (I2/I4) ratio was minimal compared to the industrial control samples. This indicates a smaller amount of disorder in the multilayer graphene subsystem and smaller sizes of clusters in their amorphous part [64,65,66].
The Isp2/Isp3 ((I2 + I4)/(I1 + I3)) ratio was also minimal in comparison with the control industrial samples. This means a larger role of graphite short-range sp3 bonds in the mechanically stronger amorphous part compared to the graphene multilayer lobes and the bridges between crystallites [58,59,60].
The ratio (I1 + …AI)/(I5 + …I7) was at a maximum suggesting a higher conductivity of the multilayer graphene due to a higher degree of screening of overtone photon-phonon interactions by carriers [64,65,66,68].

5. The Role of Ferric Compounds Studied with MS; Degradation Mechanisms of LiFePO4 Test Cathodes with XRD

5.1. Mössbauer Spectroscopy Studies

The obtained Mössbauer spectra are processed by the least squares method using the Lamb–Mössbauer factors [107]. Table 4 shows the MS calculations results obtained with a high Doppler shift rate of the gamma source. Figure 5 shows the Mössbauer spectra of sample N1, and Table 4 shows the calculation results. From Figure 5 and Table 4, it can be seen that the Mössbauer spectra consist of two doublets superimposed on each other, and no additional lines indicating the presence of another phase are observed. The values of the Fe2+ and Fe3+ absorption lines relative intensities are determined from the experimental spectra; the line half-widths are given in Table 4.
As can be seen in Figure 5, MS at room temperature (295 K) does not show any magnetic ordering lines traces, and the spectrum consists of quadrupole doublets. This means that the sample is in a paramagnetic state and no magnetic ordering traces or relaxation processes are observed. High intensity doublet lines are symmetrical. The spectra show a doublet with a small linewidth, maximum intensity, and hyperfine interaction (HFI) parameters: IS = 0.981(1) mm/s, QS = 2.926(2) mm/s. A doublet with similar HFI parameters IS = 1.23 mm/s and QS = 2.96 mm/s is observed for iron ions in the olivine structure, which corresponds to the high-spin iron Fe2+ in an octahedral M2 environment [48,80,108]. In the case of LiFePO4 with the olivine structure, a doublet with IS = 1.22 mm/s and QS = 2.80 mm/s, which is attributed to Fe2+ ions, is also observed [45,46]. However, the IS values obtained from experimental MS (Figure 5) are somewhat lower (0.981 mm/s). The small linewidth of the dominant doublet (0.281 mm/s) means that the Fe2+ ions occupy positions in the well-ordered LiFePO4 phase structure. The IS and QS values are close to those obtained using the density functional theory (DFT) calculations, taking into account both the spin polarization and the correlation of Fe 3d electrons [109]. This can be explained by the high-spin configuration of Fe2+ ions in a distorted octahedral environment formed by oxygen ions.
In addition to the dominant doublet on the MS (Figure 5), in the range of velocities from −0.25 to +0.7 mm/s, low intensity broad lines (G = 0.565 mm/s) were also observed. The nature of these lines formation is a subject of discussion in the literature [110,111,112,113]. This small contribution is often ascribed to a lithium-deficient phase in Li1−xFe2+1−xFe3+xPO4 or to partial reduction of LiFePO4 in an Ar/H2 atmosphere, leading to the formation of amorphous impurity phases such as FePO4 and/or Fe2P obtained by high-temperature annealing in a partially reducing atmosphere of Ar/H2 [111,112]. Based on the results of [113], it can be argued that the Fe2+ and No. 3 lines observed on the LiFePO4 MS (Figure 5) with the HFI parameters IS = 0.981 (1) and QS =2.928 (1) mm/s IS= 1.156 (0.04) mm/s and QS = 1.818 (0.028) mm/s, belong to LiFePO4 of the olivine type and FePO4 in the amorphous state, respectively. Comparison of the [114,115,116] results indicates the absence of any ferromagnetic or ferrimagnetic impurities, such as Fe2O3, in the samples, at least in the samples with carbon-coated crystallites. The similarity of valence and local coordination states of Fe2+ ions in glasses and in LiFePO4 crystals may be the reasons for the easy formation of LiFePO4 crystals during the crystallization of lithium iron phosphate glasses (see [117] and references therein). Therefore, the following mechanism of crystallization in glasses was considered (see [118] and references therein): first, LiFePO4 crystals are formed in glasses with a high content of Fe2+ ions, after which Li3Fe2(PO4)3 crystals appear in the remaining glass phase enriched with Fe3+ ions.
Thus, the following conclusions can be drawn:
In the initial equilibrium compositions of LiFePO4, the Fe3+ content in the samples is much higher than the value that the electrochemical decrease in the Li content could provide. In the developed sample, it was at least 6–8%. The amount of Fe3+ in the literature varies from 2% to 30% [18,51,54,55,56].
A ferric compound, most likely in the form of Li3Fe2(PO4)3, is found in TEM images (Figure 3) in the form of nanocrystallites on the surface of LiFePO4 particles in both sets of—those synthesized and industrial ones. This agrees with the results obtained in [52,53,57].
For high cyclability and low sample resistance, it is necessary to have an optimal amount of Fe3+ ferric compounds, which appear as by-products of LiFePO4 synthesis. As can be seen from Table 2, the Fe3+ content of about 2% will be insufficient. The values of 5–8% will be optimal for the capacity value [18,56,119] and cycling; significantly larger contents will be excessive [18,54,55,56] due to comparability of Fe3+ concentration with the total content of Fe ions.
These results, together with the results of the RS studies, demonstrate that the surface of the developed powder is a high-quality, low-resistance and mechanically strong ferric-graphite-graphene composite with inclusions of the Li3Fe2(PO4)3 ferric (Fe3+) compound with crystallite sizes < 10 nm, which increases cyclability compared to industrial cathodes.

5.2. Ageing Mechanisms of LiFePO4 and Test Cathode Study Using XRD and EDX

Degradation of LiFePO4 crystallites is only a part of the degradation of the cathode and of the battery as a whole [120,121,122,123]. The batteries were studied in the post-mortem state [120]; the possibility of using characterization techniques with spatial resolution from Å to mm-cm was discussed in [121]. The effect of stress factors (time, temperature and state-of-charge) on battery degradation during long-term testing up to 44 months was shown in [122]; the degradation mechanisms were classified into three levels—atomic, interface and electrode scale [123]. In [124] a review of manufacturer-provided characteristics of Li-ion batteries was made. A Radon–Nikodym based approach, where probability density is built first and then used to average observable dynamic characteristic was developed and applied to determination of relaxation rate distribution from experimental measurements.
Degradation of electrode powders can include deterioration of the conductive carbon network near the interfaces [125] and its amorphization [126], appearance of cracks in crystallites [125] and amorphization of their surface [127], impurity atoms introduction into the working crystallites [128]. To reduce the Fe diffusion into the electrolyte, nano-carbon coatings are used [129]. A pyrrole (PPy) coating suppresses Fe dissolution and allows for extended retention of the olivine structure [127]. Modification of carbon by using ZnO [130], and Poly(styrene sulfonic acid) membranes by polymerization of aniline improves the coating and reduces its resistance [131,132].
Separately, degradation of crystallites may result from chemical and mechanical attacks by stress–corrosion and erosion–corrosion [133]. In the case of acids present in the electrolyte, such as HF [134], impurities catalyze these attacks: iron-rich phases have a lower corrosion potential relative to LiFePO4, and phosphorus-rich impurity has a higher value [135]. Corrosion proceeds especially actively at the points of concentration of mechanical stresses, in places where cracks appear on the surface of crystallites [133]. Another example of such attacks is the formation of amorphous layers of LiFePO4(OH) on the surface of crystallites when powders are kept in a humid atmosphere, or due to the moisture and OH groups residual presence in batteries [136].
Thus, based on the literature analysis, we can conclude that, in general, the growth mechanism and LiFePO4 crystallites degradation is a complex chemical and electrochemical process. To describe the first irreversible phase of high-temperature degradation, their explanation is combined corrosion, stress– or erosion–induced. A significant part of the degradation in temperature ranges from growth to 100 °C can be described by the Avrami–Erofe’ev reversible mechanisms [137], Ostwald ripening reaction [138] and Ostwald’s rule of stages [139], provided that the cathode powder volume is preserved, for example, by excluding its components’ diffusion into the electrolyte.

5.3. Test Cathode Aging Study Using XRD and EDX

XRD Bruker D8 Discover diffractometer was used to determine also the unit cell volume, V (Table 5 and Figure S2).
Table 5 lists the results which correspond to the first aging phase characterized with a partial destruction (or stabilization) of the protective ferric-graphite-graphene layer on the surface of the crystallites. As can be seen from Figure S1, significant deviations from stoichiometry within the LiFePO4 olivine structure are possible when the equality 2 VLi ≈ FeLi is satisfied. With these deviations, the cell volume increases and, according to [54,140], for every two lithium vacancies 2 VLi, a FeLi defect arises (iron in lithium position).
Based on Table 5 and Figures S1 and S2 the following conclusions can be drawn:
According to EDX measurements, during the first aging phase, the Fe content in the N1-3 test cathodes decreases mainly due to its diffusion from the intercrystallite space into the electrolyte; an increase in the Fe content in the lowest quality sample N4 indicates the beginning of its crystallite destruction.
For the samples N1–4, a decrease in the size of crystallites is observed, and for the samples N2–4, a decrease in the volume of unit cells Vb-Va is observed, which is proportional to a decrease in the number of cycles. The latter also means a decrease in the FeLi and VLi defect concentrations, i.e., the crystallites approach the stoichiometric composition. However, an increase in deviation from stoichiometry is observed for the developed sample N1 in the first phase of degradation.
In the experiments with incompletely discharged test cathodes, an increase in crystallite size up to 60 nm is observed (significantly smaller than in the initial powder) as the concentration of the FePO4 phase increases (see Figure S2). In this case, the unit cell dimensions of the remaining crystallites decrease, i.e., they approach the stoichiometric composition. This means that smaller crystallites have a greater deviation from stoichiometry, by 5% VLi relative to the average value. Note that the XRD measurement of the discharged test cathodes is lengthy; therefore, the process of relaxation of partially discharged crystallites already ends as a result of the Li redistribution relaxation between crystallites in these samples [141,142].
Thus, the LiFePO4 cathode degradation occurs in two stages: at the first stage, the layer on the surface of the crystallites is destroyed; at the second stage, Fe escapes into the electrolyte and onto the anode with a decrease in the size of the crystallites due to thickening of the amorphized near-surface layer. In Figure 6, this two-stage process scheme is shown; it is close to that previously proposed for describing the Ostwald ripening reaction and Ostwald’s rule of LiFePO4 crystallization stages [137,138,139]. According to these works, at high temperatures, the growth part consists of Ostwald ripening stages from the metastable state of the feedstock with free energy ∆G > 0 up to the equilibrium state with zero energies. The presence of a local electrochemical potential must also be included in the height of the barriers.

6. The Galvanostatic Measurements of LiFePO4 Test Cathodes

The main target quantitative parameters of the electrodes are: rate capability Q(t) and capacity Q0, limit value at charging time t→∞. These parameters are actively used in the development of electrodes [143,144], batteries [145,146] and supercapacitors [147,148] to assess the quality of crystallites [149] and their carbon coatings [150,151] in studies of degradation of powders and batteries in general [120,121,122,123]. In addition to these parameters, three characteristic discharge/charge times associated with RC electrical (τel), diffusion relaxation (τd) and electrochemical reaction at the electrode/electrolyte interface (τc) are also important [152]. In turn, the first two consist of 3 components each, so we have 7 characteristic times in total. It was shown in [99] that the crystallite shape engineering task aiming to optimize the rate capability and increase the cathode capacity can be divided into two subtasks: 1. Achieving a large rate capability (and capacity) at big times or increasing the rate capability at small times. 2. Decreasing characteristic discharge/charge times to increase the rate capability at small times, which can be partially solved by improving the quality of their coating.
To develop the analytical dependence Q(t) and use it to describe the results of galvanostatic measurements of test cathodes, it is necessary to establish a hierarchy among these 7 relaxation times of electrochemical charge exchange. To do this, it is necessary to determine the value of the specific interfacial area in the electrodes [153], taking into account the anisotropic size distribution of crystallites.

6.1. Test Cathode Aging Study Using XRD and EDX

Let us define the term “specific interfacial area” as a s = S/V, where S is the total area of projections of the test cathode crystallites onto the (010) plane, and V is their total volume. It was shown in [98,99] that the combined use of the results of TEM and XRD measurements makes it possible to determine these parameters using the anisotropic size distribution of LIFePO4 powder crystallites, which is described by a 3-dimensional lognormal function:
f L ¯ = 1 L 1 L 2 L 3 2 π 3 det K exp 1 2   ln L ¯ ln L ¯ ¯ T Λ ¯ 1 ln L ¯ ln L ¯ ¯ ,
where L ¯ = L 1 L 2 L 3 —crystallite sizes, L ¯ ¯ = L ¯ 1 L ¯ 2 L ¯ 3 —their means, Λ ¯ = σ 1 2 r 12 σ 1 σ 2 r 13 σ 1 σ 3 r 21 σ 2 σ 1 σ 2 2 r 23 σ 2 σ 3 r 31 σ 3 σ 1 r 32 σ 3 σ 2 σ 3 2 —correlation moment matrix, and the product, C o v i k = r i k σ i σ k —covariances, r i k —correlation coefficients between the i-th and k-th anisotropic distributions with possible values from 0 to 1 [154], excluding negative values. Table 6 shows the parameters of the developed samples N1 and N2, described earlier in [99].
The calculation of a s is performed through the following steps:
-
using weight (0.015 g) of the initial amount in the LiFePO4 powder sample and its pycnometric density (3.6 g/cm3), the total volume of all crystallites in the cathode is calculated (V = 4.2 × 1018 nm3),
-
the average crystallite volume is calculated (Mathematica 12 notation):
v p r ¯ = T o t a l f ¯ v ¯ ,   3 ,
where v ¯ —3-dimensional N-bit matrix of particle volumes, each element of which for ellipsoid particles has a volume π 6   L i n * L j n * L k n . Index n runs over values from 1 to N, while L i n , L j n   and   L k n are the sizes of crystallites along the [010], [100] and [001] axes, respectively; f ¯ is the discretization of function (1) normalized to 1 in the form of a 3-dimensional N-bit matrix, each element of which means the probability of occurrence of the crystallites with the corresponding sizes. The Total operator means the matrix elements product and all products summation:
-
dividing V by v p r ¯ we obtain the number of particles Nct in the cathode and the S value by calculating an equation similar to (2). Instead of v ¯ it uses the matrix s ¯ —particle area projections onto the (010) plane, and the program line is as follows:
S = N c t   T o t a l f ¯ s ¯ ,   3 ,
-
as a result, for the developed powder, we obtain the number of crystallites in the cathode Nct = 4.1 × 1012; the total areas S = 8.3 × 1016 nm2 and the specific interfacial area a s 2 × 107 m−1 can be calculated. The results for 2 samples are shown in Table 6.
Since we have limited ourselves to powder improvement technology, electrochemical tests were carried out by fabricating thin test cathodes using a three-electrode cell [99]. In this case, only 3 out of 7 characteristic times will remain: diffusion relaxation τd along the crystallite [010] axis columns, RC electric τel associated with the coating of crystallites, and the response time to the electrode/electrolyte interface tc. To estimate the latter, we use the value of specific interfacial area a s obtained above, as well as the analyses given in [152,153,154], quantitative calculations and the following expression:
t c = F ε c e 1 t + 0 a s i 0 e x p α c F R T η ,
where Faraday’s constant F = 96,487 C·mol−1, porosity ε = 0.3, electrolyte concentration ce = 1000 mol·m−3, transference number t + 0   = 0.4, cathodic transfer coefficient α c = 0.5 taken from [155,156,157], and the values a s = 2 × 107 m−1, surface overpotential η = 0.1 V and reference exchange current density = 1 .5 A/m2 are obtained from the developed sample measurements. Substituting numerical values into (4), we obtain t c = 0.5 s. Thus, comparing the value of t c with those obtained in [155], we are convinced that in our sample with 8 μm cathode thickness, its value is indeed the minimum in the hierarchy of electrochemical charge exchange relaxation times, which allows simplifying the Q(t) analytical dependence, leaving only the parameters of f L ¯ and τd, τel.

6.2. Q(t) Dependence on Crystallite Parameters, Lithium Diffusion Coefficient D along [010] and the Quality of Their Coating (Electrical Relaxation Time τel)

To develop an analytical model Q(t), we make the following assumptions:
  • In [152,158], for some current source, for which Q(t) asymptotically approaches the limit value Q M at t→∞, and at t→0 it approaches the dependence Q M 2   τ t n , the following empirical equation was proposed:
    Q t = Q M 1 τ t n 1 e τ t n ,
    where τ is the time constant, and the exponent values n are defined in [152] for batteries and supercapacitors, as 0.5 and 1.0, respectively. That is, the exponent n is equal to the slope tangent of the dependence Q(t) in double logarithmic coordinates at small t. Equation (5) can be interpreted as follows: over time, Q(t) reaches its limit value Q M with probability [1 − P], where P is equal to the subtract in the square bracket of Equation (5) and has the meaning of the process probability not being implemented due to the limited rate [159].
  • According to [99], the crystallite is divided into columns with a cross-sectional area d x 3 * d x 2 along the Li diffusion direction—axis [010], along which the coordinate axis x 1 is directed. The crystallite rate capability q c r t ,   L 1 , L 2 , L 3 is determined by integrating over the plane (010) the rate capability q s e t of length M column:
    q c r t ,   L 1 , L 2 , L 3 = L 3 2 L 3 2 d x 3 L 2 2 L 2 2 q s e t d x 2 ,
    where L 1 , L 2   and   L 3 are the crystallite dimensions along the [010], [100] and [001] axes, respectively.
  • Sequential charge carriers flow in a crystallite column through a capacitor (the model of a dense electric double layer) and an element with distributed parameters (the model of Warburg element diffusion of the stage limiting the rate of the Faradaic process). The model can be considered similar to the electrical circuit in which the capacitor and the Warburg element are series-connected [42,160].
  • The chain Figure 7 corresponds to the probabilistic equation of the sequence of events [154]:
    q s e t = q M 1 Ρ C 1 Ρ W = q M 1 τ e l t 1 1 e τ e l t 1 1 τ d t 0.5 1 e τ d t 0.5 ,  
    in which the rate capability q s e t is not realized with probability Ρ C   and   Ρ W in time t. The dependence τ d =   M 2   π 2   D is used, where D = Const, which describes the process of diffusion (desorption) from a finite size M with associated boundary conditions [161].
  • Next, similarly to Equations (2) and (3), the desired dependences are calculated:
    Q t , D ,   τ e l = T o t a l f ¯ q c r ¯ ,   3 ,
  • Figure 8 shows the fitting of the dependence of the calculated sum results (7) on the discharge time with reference to the experimental normalization value of the rate capability at t n r = 80 s (10 C rate), which is intermediate between the dominant contributions.
Figure 7. Cathode equivalent circuit of a crystallite column consisting of a capacitor C1 and Warburg element W1 series connection.
Figure 7. Cathode equivalent circuit of a crystallite column consisting of a capacitor C1 and Warburg element W1 series connection.
Energies 16 01551 g007
Figure 8. (a) Theoretical Q t , D ,   τ e l dependence, Equation (8), on the discharge time for the developed powder with the steps of fitting the most optimal black curve to the experimental points. (b) The same optimal curve and experimental points on large scales along the axes and large values of D and τ e l to demonstrate tilt angles. Straight lines with slopes n corresponding to the Warburg element and capacitor are shown, with the intersection at the point t = 6 s close to the obtained value τ e l = 8 s.
Figure 8. (a) Theoretical Q t , D ,   τ e l dependence, Equation (8), on the discharge time for the developed powder with the steps of fitting the most optimal black curve to the experimental points. (b) The same optimal curve and experimental points on large scales along the axes and large values of D and τ e l to demonstrate tilt angles. Straight lines with slopes n corresponding to the Warburg element and capacitor are shown, with the intersection at the point t = 6 s close to the obtained value τ e l = 8 s.
Energies 16 01551 g008
The theoretical dependence Q(t) can use two relaxation times τ d and τ e l , which are the most important in the hierarchy of relaxation times. The procedure for fitting Q t , D ,   τ e l , expression (7), to the experimental Q(t) includes the use of the distribution parameters of anisotropic crystallite sizes, as well as the normalization value Q t n r at some t n r . For the obtained value τ e l = 8 s, the above estimate of the need to fulfill the inequality τ e l < t c = 0.5 s is performed with a large margin. As can be seen from Table 2 and Table 6, the crystallite average sizes along the [010] axis of the developed sample N 1 are reduced by 2–3 times relative to the rest. At the same time, it is significant that the crystallite average volume is even more reduced, in particular, by a factor of 30 compared to sample N4. This fact also indicates the high quality of the developed powder, since a decrease in the crystallite volume should obviously reduce the crystallite lifetime (cycling).

6.3. Q(t) Calculation in the Ranges of D, τ e l u r ¯ , Close to the Values of the Developed LiFePO4 Powder to Assess the Possibility of Improving Technologies

As can be seen from Figure 8b, a 25% decrease in D and a decrease in τ e l by a factor of 4 have practically no effect on the value of Q(t) in the practically important range of discharge rates up to 50 C rate; the latter corresponds to the rightmost experimental point. From Figure 8b, it is also seen (purple solid and dotted curves), that an obvious way to improve the powders would be technologies aimed at increasing D by a factor of three and reducing τel to extremely small values equal to t c = 0.5 s.
In [99 + SI] a significant number of different calculations are presented in wide ranges of anisotropic distribution parameters of crystallite sizes, which are in agreement with known experimental results. Below, the dependence of Q(t) on the values of the correlation coefficients of Equation (1) (the correlation between the crystallite linear sizes) will be described.
It should be noted that experimental determination of the correlation coefficients is quite reliable for crystallites with plate- and bar-like shapes, using SEM or TEM microscopy. The largest face of the crystallites is located in the object microscope plane and, in principle, no XRD measurement is required. However, there frequently are more complicated situations, with less anisotropy (<5-fold). Here, only complementary XRD and TEM measurements are possible [98]. The essence of the technique is that the difference between the column averaged L ¯ i   XRD   and the volume averaged L ¯ i   TEM can be related to the average sizes of real measurements   L ¯ i for log-normal distributions. In this case, the TEM measurement results produce a correlation cloud between the longitudinal and transverse sizes of crystallites and the corresponding marginal distribution functions [159]. Figure 9a shows such a cloud obtained by digitizing TEM images of the developed sample. Luckily, they are described by a lognormal function (normal in the ln coordinate). This allows us to divide them into 3 components along the crystallographic axes and then, using a fitting procedure, find the correlators r i k which are also shown in Table 6. The procedure is described in [99] and implies using the correlation between the longitudinal and transverse particle sizes (see Figure 9a) r b s as a trial to obtain r i k . The program fits the row-by-row sum of the 3-D matrix f ¯ to the corresponding 3-D marginal size distributions along the crystallographic axes. Figure 9a shows the ordering of points, which is due to the discreteness of the particle size digitization, as well as the coincidence of some particles in sizes.
Figure 9b shows a crystallite density distribution contour map. To take the correlations into account, the cloud can be discretized into sections and correlations calculated for each section. An example for 3 sections is shown in Figure 9b. However, to reduce the error in determining r i k , it is necessary to increase the number of digitized particles, which will significantly increase the processing time, considering the overlap of the particles in TEM images [98]. On the other hand, the difference between the r i k values the studied samples are relatively small, no more than 3 tenths; therefore, to determine the dependence of Q(t) on this parameter, we use the r ¯ average value in a wider range from 0 to 1. The calculation results shown in Figure 10 are carried out using the average value r ¯ indicated in Table 6.
As can be seen from Figure 10b, the dependence of Q(t) on r ¯ in the range of short discharge times (high power) is nonmonotonic for the developed sample. Note that it satisfies the inequality   L ¯ 1 > L ¯ 2 , and   L ¯ 1 ~   L ¯ 3 . To increase Q(t) in the region of small t, when these relations are satisfied, it is necessary to increase r ¯ , in other words, to reduce the cloud width in Figure 9b. As can be seen in the inset to Figure 10a, this can be expected to be quite realistic, since the r ¯ achieved value is on the threshold of its sharp rise. However, when developing a powder with particle sizes   L ¯ 1 < L ¯ 2 ,   L ¯ 3 , the Q(t) dependence decreases as r ¯ decreases; i.e., it is necessary to strive for an increase in the cloud width.
To discuss the mechanisms of the discovered Q(t) dependence on r ¯ , it is necessary to pay attention to the nonmonotonic change in the rbs value with increasing particle sizes observed in Figure 9b. The inset to Figure 10a shows a scheme according to which Li leaves small particles with rbs = 0.70 at small discharge t (high currents), and then, as the time t increases, Li additionally leaves medium-sized particles with rbs = 0.52, and finally, Li leaves large particles with rbs = 0.79. Since Q(t) is strongly dependent on the particle size, all these factors together may possibly lead to the observed dependence of Q(t) on r ¯ .
Thus, as shown in Figure 8b, to improve the developed powder technology, it is necessary to increase the lithium diffusion coefficient D in LiFePO4 crystallites by up to three times. In this case, it is necessary to improve their coatings quality from reducing the electrical relaxation time τel to the electrochemical reaction duration at the electrode/electrolyte interface τ c = 0.5 s. For an additional increase in Q(t) (and power) in the region of short recharge times, it is necessary to optimize the correlation coefficients between the crystallites anisotropic sizes. They can be controlled by the cloud width of the correlations between the transverse and longitudinal dimensions of TEM particle images (see Figure 9b).

7. Conclusions

  • One-pot synthesis has been developed for LiFePO4 powders with impurity phase content of less than 0.1%, with 2–3 times smaller crystallites along the [010] axis, with 2–3 times greater cycling compared to the industrial samples, and with the particles covered by a mechanically strong, low-resistance ferric-graphite-graphene composite protective layer with inclusions of ferric (Fe3+) compound particles 5–10 nm in size. The ordered carbon shell thickness reaches 5 nm, and the amorphous shell is up to 20 nm.
  • To detect impurity crystallite phases, SXRD was used, since conventional XRD is less sensitive due to the lower intensity of laboratory sources compared to the synchrotron.
  • Control of adipic acid and polyvinyl alcohol concentrations and use of multistage annealing modes makes it possible to control the coating quality. The composite layer improves cyclability compared to industrial cathodes.
  • The role of ferric Fe3+ compounds:
    -
    the content of ferric Fe3+ compounds is much higher (at least 6–8%) than the value expected from the electrochemical decrease in the Li content. The amount of Fe3+ reported in the literature varies from 2% to 30% (Table 7).
    -
    in a controlled way, Fe3+ compounds can be formed on the surface when the volatile components are not completely removed during LiFePO4 synthesis from an intermediate low-temperature amorphous phase.
    -
    To obtain highly cyclable and low resistant samples, it is necessary to have some optimal amount of the Fe3+ ferric compounds, which appear as LiFePO4 synthesis by-products. EDX studies of the tested cathode show that the total number of Fe atoms is reduced compared to the original samples. We have not detected Fe2O3, but it was observed in other technologies.
    -
    According to the corrosion degradation model, an increase in the cycle number leads to a decrease in the ferric Fe3+ compounds content on the surface of crystallites. These compounds play a certain sacrificial role [162,163], disappearing as the cathode resource is exhausted, and impurity phases can play the role of a catalyst for this breakdown. However, some of them, such as iron phosphide, weaken the catalytic activity. The degradation occurs in two stages: at the first stage, the layer on the crystallite surface is destroyed, and at the second stage, Fe escapes into the electrolyte and onto the anode with a decrease in the crystallite size due to increasing amorphization of the near-surface layer of crystallites.
  • Galvanostatic studies of the N1 sample test cathodes were carried out in 3 stages with an assessment of the possibility to further improve the technology.
    5.1
    To develop a theoretical dependence Q(t) that takes into account the 3D lognormal crystallite size distribution f L ¯ , the response time of the electrode/electrolyte interface t c is estimated using the specific interfacial area in the electrodes a s = S/V, where S is the total the projected area of the test cathode crystallites on the (010) plane, and V is their total volume. The value tc = 0.5 s is obtained.
    5.2
    Comparing the t c value with those obtained in [155], it appears to be a minimal one in the hierarchy of relaxation times of electrochemical charge exchange. This makes it possible to simplify the theoretical equation for the Q(t) dependence on the f L ¯ parameters. Fitting the theoretical dependence to the experimental data gives the value of the Li diffusion coefficient, D = 0.12 nm2/s. The value of τ e l = 8   s satisfies the inequality τ e l > t c = 0.5 s
    5.3
    Q(t) calculations in the ranges of diffusion coefficients D, electrical relaxation times τ e l , and correlation coefficients r ¯ close to the values characteristic of the developed LiFePO4 powder show that a decrease in D by 25% and a decrease in τ e l by a factor of 4 has practically no effect on the Q(t) value in the practically important range of discharge rates up to 50 C. Improving the powder technology should be aimed at increasing D three times and reducing τ e l to extremely small values closer to t c = 0.5 s. For an additional increase in Q(t) (and power) in the short recharge time region, it is necessary to optimize the values of the correlation coefficients between the anisotropic crystallite sizes.
  • Table 7 summarizes the obtained parameters and compares them with the known ones, taking into account comments on them.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en16031551/s1, Figure S1: The dependence of the LiFePO4 orthorhombic olivine structure unit cell volume on the values of Li deficiency and Fe excess; Figure S2: The results of XRD measurements of N1 test cathode depending on its discharge degree; Section S1: SEM and EDX studies of samples before and after test cathodes 100-fold cycling made from developed and commercial powders.

Author Contributions

D.A., powder synthesis, methodology; A.B., conceptualization, writing, software; A.K., MS investigation, analysis; A.N., SEM investigation, analysis; E.E., data curation, original draft preparation; A.U., electrochemistry investigation, analysis; I.K., XRD investigation, analysis; V.L., RS investigation, analysis; M.T., TEM investigation, analysis; E.T., data analysis, validation. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

Sincere thanks to V.G. Malyshkin for discussing the aging mechanisms. XRD studies were performed in Resource Center of St. Petersburg State University, SXRD studies—in Synchrotron radiation source of Kurchatov Institute Research Center, SEM studies—in Federal Joint Research Center “Material science and characterization in advanced technology”, TEM studies—in Omsk Regional Center of Collective Usage SB RAS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yamada, A.; Chung, S.C.; Hinokuma, K.J. Optimized LiFePO4 for Lithium Battery Cathodes. Electrochem. Soc. 2001, 148, A224–A229. [Google Scholar] [CrossRef]
  2. Ong, S.P.; Wang, L.; Kang, B.; Ceder, G. Li–Fe–P–O2 Phase Diagram from First Priciples Calculations. Chem. Mater. 2008, 20, 1798–1807. [Google Scholar] [CrossRef]
  3. Bezerra, C.A.G.; Davoglio, R.A.; Biaggio, S.R.; Bocchi, N.; Rocha-Filho, R.C. High-purity LiFePO4 prepared by a rapid one-step microwave-assisted hydrothermal synthesis. J. Mater. Sci. 2021, 56, 10018–10029. [Google Scholar] [CrossRef]
  4. Henriksen, C.; Mathiesen, J.K.; Ravnsbæk, D.B. Improving capacity and rate capability of Li-ion cathode materials through ball milling and carbon coating—Best practice for research purposes. Solid State Ion. 2020, 344, 115–152. [Google Scholar] [CrossRef]
  5. Liu, Y.; Luo, G.Y.; Gu, Y.J.; Wu, F.Z.; Mai, Y.; Dai, X.Y. Study on the Preparation of LiFPO4 by Hydrothermal Method. IOP Conf. Ser. Mater. Sci. Eng. 2020, 761, 012004. [Google Scholar] [CrossRef]
  6. Nan, C.; Lu, J.; Li, L.; Li, L.; Peng, Q.; Li, Y. Size and shape control of LiFePO4 nanocrystals for better lithium ion battery cathode materials. Nano Res. 2013, 6, 469–477. [Google Scholar] [CrossRef]
  7. Kudryavtsev, E.N.; Sibiryakov, R.V.; Agafonov, D.V.; Naraev, V.N.; Bobyl, A.V. Modification of liquid-phase synthesis of lithium-iron phosphate, a cathode material for lithium-ion battery. Russ. J. Appl. Chem. 2012, 85, 879–882. [Google Scholar] [CrossRef]
  8. Janssen, Y.; Santhanagopalan, D.; Qian, D.; Chi, M.; Wang, X.; Hoffmann, C.S.; Meng, Y.; Khalifah, P.G. Reciprocal Salt Flux Growth of LiFePO4 Single Crystals with Controlled Defect Concentrations. Chem. Mater. 2013, 25, 4574–4584. [Google Scholar] [CrossRef]
  9. Abakumov, A.M.; Fedotov, S.S.; Antipov, E.V.; Tarascon, J.-M. Solid state chemistry for developing better metal-ion batteries. Nat. Commun. 2020, 11, 4976. [Google Scholar] [CrossRef]
  10. Ming, X.-L.; Wang, R.; Li, T.; Wu, X.; Yuan, L.-J.; Zhao, Y. Preparation of Micro-Nano-Structured FePO4·2H2O for LiFePO4 Cathode Materials by the Turbulent Flow Cycle Method. ACS Omega 2021, 6, 18957–18963. [Google Scholar] [CrossRef]
  11. Jittmonkong, K.; Roddecha, S.; Sriariyanun, M. One-pot Synthesis of LiFePO4 Nano-particles Entrapped in Mesoporous Melamine-Formaldehyde Matrix as the Promising Cathode Materials for the Next Generation Lithium Ion Batteries. Mater. Today Proc. 2019, 17, 1284–1292. [Google Scholar] [CrossRef]
  12. Zhang, B.; Wang, S.; Liu, L.; Li, Y.; Yang, J. One-Pot Synthesis of LiFePO4/N-Doped C Composite Cathodes for Li-ion Batteries. Materials 2022, 15, 4738. [Google Scholar] [CrossRef]
  13. Kaurz, G.; Gates, B.D. Review–Surface Coatings for Cathodes in Lithium Ion Batteries: From Crystal Structures to Electrochemical Performance. J. Electrochem. Soc. 2022, 169, 043504. [Google Scholar] [CrossRef]
  14. Khodabakhshi, S.; Fulvio, P.F.; Andreoli, E. Carbon black reborn: Structure and chemistry for renewable energy harnessing. Carbon 2020, 162, 604–649. [Google Scholar] [CrossRef]
  15. Ren, W.; Wang, K.; Yang, J.; Tan, R.; Hu, J.; Guo, H.; Duan, Y.; Zheng, J.; Lin, Y.; Pan, F. Soft-contact conductive carbon enabling depolarization of LiFePO4 cathodes to enhance both capacity and rate performances of lithium ion batteries. J. Power Sources 2016, 331, 232–239. [Google Scholar] [CrossRef]
  16. Wang, X.; Feng, Z.; Huang, J.; Deng, W.; Li, X.; Zhang, H.; Wen, Z. Graphene-decorated carbon-coated LiFePO4 nanospheres as a high-performance cathode material for lithium-ion batteries. Carbon 2018, 127, 149–157. [Google Scholar] [CrossRef]
  17. Li, Y.; Qi, F.; Guo, H.; Guo, Z.; Li, M.; Wu, W. Characteristic investigation of an electrochemical-thermal coupled model for a LiFePO4/Graphene hybrid cathode lithium-ion battery. Case Stud. Therm. Eng. 2019, 13, 100387. [Google Scholar] [CrossRef]
  18. Zhang, B.; Xu, Y.; Wang, J.; Ma, X.; Hou, W.; Xue, X. Electrochemical performance of LiFePO4/graphene composites at low temperature affected by preparation technology. Electrochim. Acta 2021, 368, 137575. [Google Scholar] [CrossRef]
  19. Xi, Y.; Lu, Y. Toward Uniform In Situ Carbon Coating on Nano-LiFePO4 via a Solid-State Reaction. Ind. Eng. Chem. Res. 2020, 59, 13549–13555. [Google Scholar] [CrossRef]
  20. Zhang, H.; Wang, L.; Chen, Y.; Wen, X. Regenerated LiFePO4/C for scrapped lithium iron phosphate powder batteries by pre-oxidation and reduction method. Ionics 2022, 28, 2125–2133. [Google Scholar] [CrossRef]
  21. Ravet, N.; Chouinard, Y.; Magnan, J.F.; Besner, S.; Gauthier, M.; Armand, M. Electroactivity of natural and synthetic triphylite. J. Power Sources 2001, 97–98, 503–507. [Google Scholar] [CrossRef]
  22. Huynh, L.T.N.; Nguyen, H.H.A.; Tran, T.T.D.; Nguyen, T.T.T.; Nguyen, T.M.A.; La, T.H.; Tran, V.M.; Le, M.L.P. Electrode composite LiFePO4@ Carbon: Structure and electrochemical performances. J. Nanomater. 2019, 2019, 2464920. [Google Scholar] [CrossRef]
  23. Altin, E.; Altundag, S.; Gultek, E.; Altin, S. Li1+xFePO4 (x = 0–0.5) production from Fe3+ sources by glass-ceramic technique with different carbon sources and investigation of structural, thermal and electrochemical performance. J. Non-Cryst. Solids 2022, 586, 121546. [Google Scholar] [CrossRef]
  24. Lim, H.H.; Jang, I.C.; Lee, S.B.; Karthikeyan, K.; Aravindan, V.; Lee, Y.S. The important role of adipic acid on the synthesis of nanocrystalline lithium iron phosphate with high rate performance. J. Alloys Compd. 2010, 495, 181–184. [Google Scholar] [CrossRef]
  25. Xu, H.; Jing, M.; Huang, Z.; Li, J.; Wang, T.; Yuan, W.; Ju, B.; Shen, X. Cross-Linked Polypropylene Oxide Solid Electrolyte Film with Enhanced Mechanical, Thermal, and Electrochemical Properties via Additive Modification. ACS Appl. Polym. Mater. 2021, 3, 6539–6547. [Google Scholar] [CrossRef]
  26. Ramkumar, V.; Gardas, R.L. Structural Arrangements of Guanidinium-Based Dicarboxylic Acid Ionic Liquids and Insights into Carbon Dioxide Uptake through Structural Voids. Cryst. Growth Des. 2022, 22, 3646–3655. [Google Scholar] [CrossRef]
  27. Guan, L.; Liu, M.; Yu, F.; Qiu, T.; Zhou, T.; Lin, X. A LiFePO4 regeneration method based on PVAc alcoholysis reaction. Renew. Energy 2021, 175, 559–567. [Google Scholar] [CrossRef]
  28. Wang, Y.; Liu, Z.; Zhou, S. An effective method for preparing uniform carbon coated nano-sized LiFePO4 particles. Electrochim. Acta 2011, 58, 359–363. [Google Scholar] [CrossRef]
  29. Ravet, N.; Gauthier, M.; Zaghib, K.; Goodenough, J.B.; Mauger, A.; Gendron, F.; Julien, C.M. Mechanism of the Fe3+ Reduction at Low Temperature for LiFePO4 Synthesis from a Polymeric Additive. Chem. Mater. 2007, 19, 2595–2602. [Google Scholar] [CrossRef]
  30. Liao, Y.; Li, G.; Xu, N.; Chen, T.; Wang, X.; Li, W. Synergistic effect of electrolyte additives on the improvement in interfacial stability between ionic liquid based gel electrolyte and LiFePO4 cathode. Solid State Ion. 2019, 329, 31–39. [Google Scholar] [CrossRef]
  31. Doeff, M.M.; Wilcox, J.D.; Kostecki, R.; Lau, G. Optimization of carbon coatings on LiFePO4. J. Power Sources 2006, 163, 180–184. [Google Scholar] [CrossRef]
  32. Kim, S.; Mathew, V.; Kang, J.; Gim, J.; Song, J.; Jo, J.; Kim, J. High rate capability of LiFePO cathodes doped with an excess amount of Ti. Ceram. Int. 2016, 42, 7230–7236. [Google Scholar] [CrossRef]
  33. Chang, H.-H.; Chang, C.-C.; Wu, H.-C.; Guo, Z.-Z.; Yang, M.-H.; Chiang, Y.-P.; Sheu, H.-S.; Wu, N.-L. Kinetic study on low-temperature synthesis of LiFePO4 via solid-state reaction. J. Power Sources 2006, 158, 550–556. [Google Scholar] [CrossRef]
  34. Lin, F.; Liu, Y.; Yu, X.; Cheng, L.; Singer, A.; Shpyrko, O.G.; Xin, H.L.; Tamura, N.; Tian, C.; Weng, T.-C.; et al. Synchrotron X-ray Analytical Techniques for Studying Materials Electrochemistry in Rechargeable Batteries. Chem. Rev. 2017, 117, 13123–13186. [Google Scholar] [CrossRef] [PubMed]
  35. Sun, X.; Xu, Y. Fe excess in hydrothermally synthesized LiFePO4. Mater. Lett. 2012, 84, 139–142. [Google Scholar] [CrossRef]
  36. Zhang, Y.; Alarco, J.A.; Nerkar, J.Y.; Best, A.S.; Snook, G.A.; Talbot, P.C.; Cowie, B.C.C. Observation of Preferential Cation Doping on the Surface of LiFePO4 Particles and Its Effect on Properties. ACS Appl. Energy Mater. 2020, 3, 9158–9167. [Google Scholar] [CrossRef]
  37. Kapaev, R.R.; Novikova, S.A.; Kulova, T.L.; Skundin, A.M.; Yaroslavtsev, A.B. Synthesis of LiFePO4 nanoplatelets as cathode materials for Li-ion batteries. Nanotechnol. Russ. 2016, 11, 757–760. [Google Scholar] [CrossRef]
  38. Kim, J.-K.; Jeong, S.M. Physico-electrochemical properties of carbon coated LiFePO4 nanoparticles prepared by different preparation method. Appl. Surf. Sci. 2020, 505, 144630. [Google Scholar] [CrossRef]
  39. Watts, J.F.; Wolstenholme, J. An Introduction to Surface Analysis by XPS and AES; John Wiley & Sons Ltd: Hoboken, NJ, USA, 2020; pp. 1–288. [Google Scholar]
  40. Grissa, R.; Abramova, A.; Tambio, S.-J.; Lecuyer, M.; Deschamps, M.; Fernandez, V.; Greneche, J.-M.; Guyomard, D.; Lestriez, B.; Moreau, P. Thermomechanical Polymer Binder Reactivity with Positive Active Materials for Li Metal Polymer and Li-Ion Batteries: An XPS and XPS Imaging Study. ACS Appl. Mater. Interfaces 2019, 11, 18368–18376. [Google Scholar] [CrossRef]
  41. Kumar, Y.A.; Kim, H.-J. Effect of Time on a Hierarchical Corn Skeleton-Like Composite of CoO@ZnO as Capacitive Electrode Material for High Specific Performance Supercapacitors. Energies 2018, 11, 3285. [Google Scholar] [CrossRef] [Green Version]
  42. Yoon, J.-H.; Kumar, Y.A.; Sambasivam, S.; Hira, S.A.; Krishna, T.N.V.; Zeb, K.; Uddin, W.; Kumar, K.D.; Obaidat, I.M.; Kim, S.; et al. Highly efficient copper-cobalt sulfide nano-reeds array with simplistic fabrication strategy for battery-type supercapacitors. J. Energy Storage 2020, 32, 101988. [Google Scholar] [CrossRef]
  43. Eds Long, G.J.; Grandjean, F. Mossbauer Spectroscopy Applied to Magnetism and Materials Science; Springer Science and Business Media: New York, NY, USA, 1996; Volume 1, pp. 1–479. [Google Scholar]
  44. Wu, Y.; Holze, R. Surface Science in Batteries. In Surface and Interface Science: Applications of Surface Science I, 1st ed.; Wandelt, K., Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Hoboken, NJ, USA, 2020; pp. 381–427. [Google Scholar] [CrossRef]
  45. Andersson, A.S.; Kalska, B.; Häggström, L.; Thomas, J.O. Lithium extraction/insertion in LiFePO4: An X-ray diffraction and Mössbauer spectroscopy study. Solid State Ion. 2000, 130, 41–52. [Google Scholar] [CrossRef]
  46. Jalkanen, K.; Lindén, J.; Karppinen, M. Local structures in mixed LixFe1–yMyPO4 (M=Co, Ni) electrode materials. J. Solid State Chem. 2015, 230, 404–410. [Google Scholar] [CrossRef]
  47. Khalifi, M.E.; Lippens, P.-E. First-Principles Investigation of the 57Fe Mössbauer Parameters of LiFePO4 and FePO. J. Phys. Chem. C 2016, 120, 28375–28389. [Google Scholar] [CrossRef]
  48. Yaroslavtsev, S.; Novikova, S.; Rusakov, V.; Vostrov, N.; Kulova, T.; Skundin, A.; Yaroslavtsev, A. LiFe1-XMgXPO4/C as cathode materials for lithium-ion batteries. Solid State Ion. 2018, 317, 149–155. [Google Scholar] [CrossRef]
  49. Hosono, E.; Wang, Y.; Kida, N.; Enomoto, M.; Kojima, N.; Okubo, M.; Matsuda, H.; Saito, Y.; Kudo, T.; Honma, I.; et al. Synthesis of Triaxial LiFePO4 Nanowire with a VGCF Core Column and a Carbon Shell through the Electrospinning. ACS Appl. Mater. Interfaces 2010, 2, 212–218. [Google Scholar] [CrossRef] [PubMed]
  50. Bini, M.; Ferrari, S.; Capsoni, D.; Mustarelli, P.; Spina, G.; Giallo, F.D.; Lantieri, M.; Leonelli, C.; Rizzutie, A.; Massarotti, V. Pair distribution function analysis and Mo¨ssbauer study of defects in microwave-hydrothermal LiFePO4. RSC Adv. 2012, 2, 250–258. [Google Scholar] [CrossRef]
  51. Bazzi, K.; Mandal, B.P.; Nazri, M.; Naik, V.M.; Garg, V.K.; Oliveira, A.C.; Vaishnava, P.P.; Nazri, G.A.; Naik, R. Effect of surfactants on the electrochemical behavior of LiFePO4 cathode material for lithium ion batteries. J. Power Sources 2014, 265, 67–74. [Google Scholar] [CrossRef]
  52. Yaroslavtsev, S.A.; Vostrov, N.I.; Novikova, S.A.; Kulova, T.L.; Yaroslavtsev, A.B.; Rusakov, V.S. Study of Delithiation Process Features in LixFe0.8M0.2PO4 (M = Mg, Mn, Co, Ni) by Mössbauer Spectroscopy. J. Phys. Chem. C 2020, 124, 13026–13035. [Google Scholar] [CrossRef]
  53. Reklaitis, J.; Davidonis, R.; Dindune, A.; Valdniece, D.; Jasulaitiene, V.; Baltrunas, D. Characterization of LiFePO4/C composite and its thermal stability by Mossbauer and XPS spectroscopy. Phys. Status Solidi B 2016, 253, 2283–2288. [Google Scholar] [CrossRef]
  54. Jensen, K.M.Ø.; Christensen, M.; Gunnlaugsson, H.P.; Lock, N.; Bøjesen, E.D.; Proffen, T.; Iversen, B.B. Defects in Hydrothermally Synthesized LiFePO4 and LiFe1-xMnxPO4 Cathode Materials. Chem. Mater. 2013, 25, 2282–2290. [Google Scholar] [CrossRef]
  55. Lazarević, Z.Ž.; Križan, G.; Križan, J.; Milutinović, A.; Ivanovski, V.N.; Mitrić, M.; Gilić, M.; Umićević, A.; Kuryliszyn-Kudelska, I.; Romčević, N.Ž. Characterization of LiFePO4 samples obtained by pulse combustion under various conditions of synthesis. J. Appl. Phys. 2019, 126, 085109. [Google Scholar] [CrossRef]
  56. Sun, S.; An, Q.; Tian, Z.; Zhao, X.; Shen, X. Low-Temperature Synthesis of LiFePO4 Nanoplates/C Composite for Lithium Ion Batteries. Energy Fuels 2020, 34, 11597–11605. [Google Scholar] [CrossRef]
  57. Jie, Y.; Yang, S.; Hu, F.; Li, Y.; Ye, L.; Zhao, D.; Jin, W.; Chang, C.; Lai, Y.; Chen, Y. Gas evolution characterization and phase transformation during thermal treatment of cathode plates from spent LiFePO4 batteries. Thermochim. Acta 2020, 684, 178483. [Google Scholar] [CrossRef]
  58. Wilcox, J.D.; Doeff, M.M.; Marcinek, M.; Kostecki, R. Factors influencing the quality of carbon coatings on LiFePO4. J. Electrochem. Soc. 2007, 154, A389–A395. [Google Scholar] [CrossRef]
  59. Rouzaud, J.N.; Oberlin, A.; Beny-Bassez, C. Carbon films: Structure and microtexture (optical and electron microscopy, Raman spectroscopy). Thin Solid Film. 1983, 105, 75–96. [Google Scholar] [CrossRef]
  60. Bonhomme, F.; Lassegues, J.C.; Servant, L. Raman spectroelectrochemistry of a carbon supercapacitor. J. Electrochem. Soc. 2001, 148, E450–E458. [Google Scholar] [CrossRef]
  61. Duan, X.; Tian, W.; Zhang, H.; Sun, H.; Ao, Z.; Shao, Z.; Wang, S. sp2/sp3 Framework from Diamond Nanocrystals: A Key Bridge of Carbonaceous Structure to Carbocatalysis. ACS Catal. 2019, 9, 7494–7519. [Google Scholar] [CrossRef]
  62. Osswald, S.; Yushin, G.; Mochalin, V.; Kucheyev, S.O.; Gogotsi, Y. Control of sp2/sp3 Carbon Ratio and Surface Chemistry of Nanodiamond Powders by Selective Oxidation in Air. J. Am. Chem. Soc. 2006, 128, 11635–11642. [Google Scholar] [CrossRef]
  63. Vejpravová, J. Mixed sp2–sp3 Nanocarbon Materials: A Status Quo Review. Nanomaterials 2021, 11, 2469. [Google Scholar] [CrossRef]
  64. Ferrari, A.C.; Basko, D.M. Raman spectroscopy as a versatile tool for studying the properties of graphene. Nat. Nanotechnol. 2013, 8, 235–246. [Google Scholar] [CrossRef]
  65. Wu, J.-B.; Lin, M.-L.; Cong, X.; Liu, H.-N.; Tan, P.-H. Raman spectroscopy of graphene-based materials and its applications in related devices. Chem. Soc. Rev. 2018, 47, 1822–1873. [Google Scholar] [CrossRef]
  66. Wu, J.-B.; Lin, M.-L.; Tan, P.-H. Raman Spectroscopy of Monolayer and Multilayer Graphenes Chapter 1. In Raman Spectroscopy of Two-Dimensional Materials; Tan, P.-H., Ed.; Springer Series in Materials Science 276; Springer Nature Singapore Pte Ltd.: Singapore, 2019. [Google Scholar]
  67. Bokobza, L.; Bruneel, J.-L.; Couzi, M. Raman spectra of carbon-based materials (from graphite to carbon black) and of some silicone composites. Carbon 2015, 1, 77–94. [Google Scholar] [CrossRef]
  68. Casiraghi, C. Doping dependence of the Raman peaks intensity of graphene close to the Dirac point. Phys. Rev. B 2009, 80, 233407. [Google Scholar] [CrossRef]
  69. Sun, L.; Deng, Q.; Fang, B.; Li, Y.; Deng, L.; Yang, B.; Ren, X.; Zhang, P. Carbon-coated LiFePO4 synthesized by a simple solvothermal method. CrystEngComm 2016, 18, 7537–7543. [Google Scholar] [CrossRef]
  70. Public Spectra. Raman Spectrum of Ketjen Black. CAS Registry Number 1333-86-4.
  71. Vivo-Vilches, J.F.; Celzard, A.; Fierro, V.; Devin-Ziegler, I.; Brosse, N.; Dufour, A.; Etienne, M. Lignin-based carbon nanofibers as electrodes for vanadium redox couple electrochemistry. Nanomaterials 2019, 9, 106. [Google Scholar] [CrossRef]
  72. Vadahanambia, S.; Chunb, H.-H.; Jung, K.H.; Park, H. Nitrogen doped holey carbon nano-sheets as anodes in sodium ion battery. RSC Adv. 2016, 6, 38112–38116. [Google Scholar] [CrossRef]
  73. Tan, H.; Xu, L.; Geng, H.; Rui, X.; Li, C.; Huang, S. Nanostructured Li3V2(PO4)3 Cathodes. Small 2018, 14, 1800567. [Google Scholar] [CrossRef]
  74. Feng, C.; Chen, Y.; Liu, D.; Zhang, P. Conductivity and electrochemical performance of LiFePO4 slurry in the lithium slurry battery. IOP Conf. Ser. Mater. Sci. Eng. 2017, 207, 012076. [Google Scholar] [CrossRef] [Green Version]
  75. Lion Specialty Chemicals Co. KETJENBLACK Highly Electro-Conductive Carbon Black; Lion: Tokyo, Japan, 2020. [Google Scholar]
  76. Folaranmi, G.; Bechelany, M.; Sistat, P.; Cretin, M.; Zaviska, F. Activated Carbon Blended with Reduced Graphene Oxide Nanoflakes for Capacitive Deionization. Nanomaterials 2021, 11, 1090. [Google Scholar] [CrossRef]
  77. Knight, D.S.; White, W.B. Characterization of diamond films by Raman spectroscopy. J. Mater. Res. 1989, 4, 385–393. [Google Scholar] [CrossRef]
  78. Cong, C.; Wang, Y.; Yu, T. Raman Spectroscopy Study of Two-Dimensional Materials Under Strain. Chapter 6. In Raman Spectroscopy of Two-Dimensional Materials; Tan, P.-H., Ed.; Springer Series in Materials Science: Berlin/Heidelberg, Germany, 2019; Volume 276, pp. 111–129. [Google Scholar] [CrossRef]
  79. Yoon, D.; Son, Y.-W.; Cheong, H. Strain-dependent splitting of the double-resonance Raman scattering band in graphene. Phys. Rev. Lett. 2011, 106, 155502. [Google Scholar] [CrossRef] [PubMed]
  80. Davydov, V.V.; Grebenikova, N.M.; Smirnov, K.Y. Method of Monitoring the State of Flowing Media with Low Transparency That Contain Large Inclusions. Meas. Tech. 2019, 62, 519–526. [Google Scholar] [CrossRef]
  81. Kamzin, A.S.; Bobyl, A.V.; Ershenko, E.M.; Terukov, E.I.; Agafonov, D.V.; Kudryavtsev, E.N. Structure and Electrochemical Characteristics of LiFePO4 Cathode Materials for Rechargeable LiIon Batteries. Phys. Solid State 2013, 55, 1385–1394. [Google Scholar] [CrossRef]
  82. Doeff, M.; Hu, M.Y.; McLarnon, F.; Kostecki, R. Effect of surface carbon structure on the electrochemical performance of LiFePO4. Electrochem. Solid State Lett. 2003, 6, A207–A209. [Google Scholar] [CrossRef]
  83. Higuchi, M.; Katayama, K.; Azuma, Y.; Yukawa, M.; Suhara, M. Synthesis of LiFePO4 cathode material by microwave processing. J. Power Sources 2003, 119–121, 258–261. [Google Scholar] [CrossRef]
  84. Yang, J.; Xu, J.J. Nonaqueous sol-gel synthesis of high-performance LiFePO4. Electrochem. Solid State Lett. 2004, 7, A515–A518. [Google Scholar] [CrossRef]
  85. Deb, A.; Bergmann, U.; Cramer, S.P.; Cairns, E.J. Structural investigations of LiFePO4 electrodes and in situ studies by Fe X-ray absorption spectroscopy. Electrochim. Acta 2005, 50, 5200–5207. [Google Scholar] [CrossRef]
  86. Fu, Q.; Sarapulova, A.; Trouillet, V.; Zhu, L.; Fauth, F.; Mangold, S.; Welter, E.; Indris, S.; Knapp, M.; Dsoke, S.; et al. In Operando Synchrotron Diffraction and in Operando X-ray Absorption Spectroscopy Investigations of Orthorhombic V2O5 Nanowires as Cathode Materials for Mg-Ion Batteries. J. Am. Chem. Soc. 2019, 141, 2305–2315. [Google Scholar] [CrossRef]
  87. Kim, D.-H.; Kim, J. Synthesis of lithium manganese phosphate nanoparticle and its properties. J. Phys. Chem. Solids 2007, 68, 734–737. [Google Scholar] [CrossRef]
  88. Lee, S.B.; Jang, I.C.; Lim, H.H.; Aravindan, V.; Kim, H.S.; Lee, Y.S. Preparation and electrochemical characterization of LiFePO4 nanoparticles with high rate capability by a sol–gel method. J. Alloys Compd. 2010, 491, 668–672. [Google Scholar] [CrossRef]
  89. Son, C.G.; Chang, D.R.; Kim, H.S.; Lee, Y.S. Synthesis and Electrochemical Properties of Nanocrystalline LiFePO4 Obtained by Different Methods. J. Electrochem. Sci. Technol. 2011, 2, 103–109. [Google Scholar] [CrossRef]
  90. Son, C.G.; Yang, H.M.; Lee, G.W.; Cho, A.R.; Aravindan, V.; Kim, H.S.; Kim, W.S.; Lee, Y.S. Manipulation of adipic acid application on the electrochemical properties of LiFePO4 at high rate performance. J. Alloys Compd. 2011, 509, 1279–1284. [Google Scholar] [CrossRef]
  91. Li, M.; Mu, B. Effect of different dimensional carbon materials on the properties and application of phase change materials: A review. Appl. Energy 2019, 242, 695–715. [Google Scholar] [CrossRef]
  92. Jugović, D.; Mitrić, M.; Cvjetićanin, N.; Jančar, B.; Mentus, S.; Uskoković, D. Synthesis and characterization of LiFePO4/C composite obtained by sonochemical method. Solid State Ion. 2008, 179, 415–419. [Google Scholar] [CrossRef]
  93. Zhao, B.; Jiang, Y.; Zhang, H.; Tao, H.; Zhong, M.; Jiao, Z. Morphology and electrical properties of carbon coated LiFePO4 cathode materials. J. Power Sources 2009, 189, 462–466. [Google Scholar] [CrossRef]
  94. Chen, H.; Chen, Y.; Gong, W.; Xiang, K.; Sun, B.; Liu, J. Preparation and electrochemical performance of LiFePO4/C composite with network connections of nano-carbon wires. Mater. Lett. 2011, 65, 559–561. [Google Scholar] [CrossRef]
  95. Gauthier, M. Phosphate materials for lithium batteries and energy storage. Procedia Eng. 2012, 46, 234–254. [Google Scholar] [CrossRef] [Green Version]
  96. Chernyshov, A.A.; Veligzhanin, A.A.; Zubavichus, Y.V. Structural materials science end-station at the Kurchatov synchrotron radiation source: Recent instrumentation upgrades and experimental results. Nucl. Instr. Meth. Phys. Res. A 2009, 603, 95–98. [Google Scholar] [CrossRef]
  97. Ectors, D.; Goetz-Neunhoeffer, F.; Neubauer, J. A generalized geometric approach to anisotropic peak broadening due to domain morphology. J. Appl. Cryst. 2015, 48, 189–194. [Google Scholar] [CrossRef]
  98. Bobyl, A.; Kasatkin, I. Anisotropic crystallite size distributions in LiFePO4 powders. RSC Adv. 2021, 11, 1379–1385. [Google Scholar] [CrossRef] [PubMed]
  99. Bobyl, A.; Nam, S.-C.; Song, J.-H.; Ivanishchev, A.; Ushakov, A. Rate Capability of LiFePO4 Cathodes and the Shape Engineering of Their Anisotropic Crystallites. J. Electrochem. Sci. Technol. 2022, 13, 438–452. [Google Scholar] [CrossRef]
  100. Julien, C.M.; Mauger, A.; Zaghib, K. Surface effects on electrochemical properties of nano-sized LiFePO4. J. Mater. Chem. 2011, 21, 9955–9968. [Google Scholar] [CrossRef]
  101. Herle, P.S.; Ellis, B.; Coombs, N.; Nazar, L.F. Nano-network electronic conduction in iron and nickel olivine phosphates. Nat. Mater. 2004, 3, 147–152. [Google Scholar] [CrossRef] [PubMed]
  102. Kosova, N.V.; Devyatkina, E.T. Lithium Iron Phosphate Synthesis Using Mechanical Activation. Chem. Sustain. Dev. 2012, 20, 61–68. [Google Scholar]
  103. Ershenko, E.; Bobyl, A.; Boiko, M.; Zubavichus, Y.; Runov, V.; Trenikhin, M.; Sharkov, M. Fe3P impurity phase in high-quality LiFePO4: X-ray diffraction and neutron-graphical studies. Ionics 2017, 23, 2293–2300. [Google Scholar] [CrossRef]
  104. Babaev, A.A.; Zobov, M.E.; Kornilov, D.Y.; Tkachev, S.V.; Terukov, E.I.; Levitskii, V.S. Optical and electrical properties of graphene oxide. Opt. Spectrosc. 2018, 125, 1014–1018. [Google Scholar] [CrossRef]
  105. Mallet-Ladeira, P.; Puech, P.; Toulouse, C.; Cazayous, M.; Ratel-Ramond, N.; Weisbecker, P.; Vignoles, G.L.; Monthioux, M.A. Raman study to obtain crystallite size of carbon materials: A better alternative to the Tuinstra–Koenig law. Carbon 2014, 80, 629–639. [Google Scholar] [CrossRef]
  106. Puech, P.; Kandara, M.; Paredes, G.; Moulin, L.; Weiss-Hortala, E.; Kundu, A.; Ratel-Ramond, N.; Plewa, J.-M.; Pellenq, R.; Monthioux, M. Analyzing the Raman spectra of graphenic carbon materials from kerogens to nanotubes: What type of information can be extracted from defect bands? C 2019, 5, 69. [Google Scholar] [CrossRef] [Green Version]
  107. Aldon, L.; Perea, A.; Womes, M.; Ionica-Bousquet, C.M.; Jumas, J.C. Determination of the Lamb-Mössbauer factors of LiFePO4 and FePO4 for electrochemical in situ and operando measurements in Li-ion batteries. J. Solid State Chem. 2010, 183, 218–222. [Google Scholar] [CrossRef]
  108. Li, Z.; Shinno, I. Next nearest neighbor effects in triphylite and related phosphate minerals. Mineral. J. 1997, 19, 99–107. [Google Scholar] [CrossRef]
  109. Lippens, P.-E.; Khalifi, M.E.; Chamas, M.; Perea, A.; Sougrati, M.-T.; Ionica-Bousquet, C.; Aldon, L.; Olivier-Fourcade, J.; Jumas, J.-C. How Mössbauer spectroscopy can improve Li-ion batteries. Hyperfine Interact 2012, 206, 35–46. [Google Scholar] [CrossRef]
  110. Xia, X.; Wang, Z.; Chen, L. Regeneration and characterization of air-oxidized LiFePO4. Electrochem. Commun. 2008, 10, 1442–1444. [Google Scholar] [CrossRef]
  111. Yu, D.Y.W.; Donoue, K.; Kadohata, T.; Murata, T.; Matsuta, S.; Fujitani, S. Impurities in LiFePO4 and their influence on material characteristics. J. Electrochem. Soc. 2008, 155, A526–A530. [Google Scholar] [CrossRef]
  112. Dhindsa, K.S.; Mandal, B.P.; Bazzi, K.; Lin, M.W.; Nazri, M.; Nazri, G.A.; Naik, V.M.; Garg, V.K.; Oliveira, A.C.; Vaishnava, P.; et al. Enhanced electrochemical performance of graphene modified LiFePO4 cathode material for lithium ion batteries. Solid State Ion. 2013, 253, 94–100. [Google Scholar] [CrossRef]
  113. Shiratsuchi, T.; Okada, S.; Yamaki, J.; Yamashit, S.; Nishid, T. Cathode performance of olivine-type LiFePO4 synthesized by chemical lithiation. J. Power Sources 2007, 173, 979–984. [Google Scholar] [CrossRef]
  114. Maccario, M.; Croguennec, L.; Wattiaux, A.; Suard, E.; Le Cras, F.; Delmas, C. C-containing LiFePO4 materials-Part I: Mechano-chemical synthesis and structural characterization. Solid State Ion. 2008, 179, 2020–2026. [Google Scholar] [CrossRef]
  115. Salah, A.; Zaghib, K.; Mauger, A.; Gendron, F.; Julien, C. Magnetic studies of the carbothermal effect on LiFePO4. Phys. Status Solidi A Appl. Mater. Sci. 2006, 203, R1–R3. [Google Scholar] [CrossRef]
  116. Salah, A.A.; Mauger, A.; Zaghib, K.; Goodenough, J.B.; Ravet, N.; Gauthier, M.; Gendron, F.; Julien, C.M. Reduction Fe3+ of impurities in LiFePO4 from pyrolysis of organic precursor used for carbon deposition. J. Electrochem. Soc. 2006, 153, A1692–A1701. [Google Scholar] [CrossRef]
  117. Hirose, K.; Honma, T.; Doi, Y.; Hinatsu, Y.; Komatsu, T. Mössbauer analysis of Fe ion state in lithium iron phosphate glasses and their glass-ceramics with olivine-type LiFePO4 crystals. Solid State Commun. 2008, 146, 273–277. [Google Scholar] [CrossRef]
  118. Wang, Z.; Su, S.; Yu, C.; Chen, Y.; Xia, D. Synthesises, characterizations and electrochemical properties of spherical-like LiFePO4 by hydrothermal method. J. Power Sources 2008, 184, 633–636. [Google Scholar] [CrossRef]
  119. Bazzi, K.; Nazri, M.; Naik, V.; Garg, V.; Oliveira, A.; Vaishnava, P.; Nazri, G.; Naik, R. Enhancement of electrochemical behavior of nanostructured LiFePO4/Carbon cathode material with excess Li. J. Power Sources 2016, 306, 17–23. [Google Scholar] [CrossRef]
  120. Waldmann, T.; Iturrondobeitia, A.; Kasper, M.; Ghanbari, N.; Aguesse, F.; Bekaert, E.; Daniel, L.; Genies, S.; Gordon, I.J.; Loble, M.W.; et al. Post-mortem analysis of aged lithium-ion batteries: Disassembly methodology and physico-chemical analysis techniques. J. Electrochem. Soc. 2016, 163, A2149–A2164. [Google Scholar] [CrossRef]
  121. Pender, J.P.; Jha, G.; Youn, D.H.; Ziegler, J.M.; Andoni, I.; Choi, E.J.; Heller, A.; Dunn, B.S.; Weiss, P.S.; Penner, R.M.; et al. Electrode degradation in lithium-ion batteries. ACS Nano 2020, 14, 1243–1295. [Google Scholar] [CrossRef]
  122. Sui, X.; Swierczynski, M.; Teodorescu, R.; Stroe, D.-I. The Degradation Behavior of LiFePO4/C Batteries during Long-Term Calendar Aging. Energies 2021, 14, 1732. [Google Scholar] [CrossRef]
  123. Wang, L.; Qiu, J.; Wang, X.; Chen, L.; Cao, G.; Wang, J.; Zhang, H.; He, X. Insights for understanding multiscale degradation of LiFePO4 cathodes. eScience 2022, 2, 125–137. [Google Scholar] [CrossRef]
  124. Bobyl, A.V.; Zabrodskii, A.G.; Malyshkin, V.G.; Novikova, O.V.; Terukov, E.I.; Agafonov, D.V. Degradation of Li-ion batteries. Application of Generalized Radon--Nikodym Approach to Direct Estimation of Degradation Rate Distribution. Izvestiya RAN Energetika 2018, 2, 46–58. Available online: https://www.researchgate.net/publication/309175377 (accessed on 23 November 2016).
  125. Scipioni, R.; Jørgensen, P.S.; Stroe, D.I.; Younesi, R.; Simonsen, S.B.; Norby, P.; Hjelm, J.; Jensen, S.H. Complementary analyses of aging in a commercial LiFePO4/graphite 26650 cell. Electrochim. Acta 2018, 284, 454–468. [Google Scholar] [CrossRef]
  126. Scipioni, R.; Jørgensen, P.S.; Ngo, D.-T.; Simonsen, S.B.; Liu, Z.; Yakal-Kremski, K.J.; Wang, H.; Hjelm, J.; Norby, P.; Barnett, S.A.; et al. Electron microscopy investigations of changes in morphology and conductivity of LiFePO4/C electrodes. J. Power Sources 2016, 307, 259–269. [Google Scholar] [CrossRef]
  127. Lia, X.; Jiang, F.; Qu, K.; Wang, Y.; Pan, Y.; Wang, M.; Liu, Y.; Xu, H.; Chen, J.; Huang, Y.; et al. First Atomic-Scale Insight on Degradation in Lithium Iron Phosphate Cathodes by Transmission Electron Microscopy. J. Phys. Chem. Lett. 2020, 11, 4608–4617. [Google Scholar] [CrossRef]
  128. Zhao, N.; Li, Y.; Zhao, X.; Zhi, X.; Liang, G. Effect of particle size and purity on the low temperature electrochemical performance of LiFePO4/C cathode material. J. Alloys Compd. 2016, 683, 123–132. [Google Scholar] [CrossRef]
  129. Wang, J.; Yang, J.; Tang, Y.; Li, R.; Liang, G.; Sham, T.-K.; Sun, X. Surface aging at olivine LiFePO4: A direct visual observation of iron dissolution and the protection role of nano-carbon coating. J. Mater. Chem. A 2013, 1, 1579–1586. [Google Scholar] [CrossRef]
  130. Chen, X.; Li, Y.; Wang, J. Enhanced Electrochemical Performance of LiFePO4 Originating from the Synergistic Effect of ZnO and C Co-Modification. Nanomaterials 2021, 11, 12. [Google Scholar] [CrossRef]
  131. Tan, S.; Tieu, J.H.; Bélanger, D. Chemical Polymerization of Aniline on a Poly(styrene sulfonic acid) Membrane:  Controlling the Polymerization Site Using Different Oxidants. J. Phys. Chem. B 2005, 109, 14085–14092. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, Y.; Wang, Y.; Hosono, E.; Wang, K.; Zhou, H. The Design of a LiFePO4/Carbon Nanocomposite with a Core–Shell Structure and Its Synthesis by an In Situ Polymerization Restriction Method. Angew. Chem. Int. Ed. 2008, 47, 7461–7465. [Google Scholar] [CrossRef]
  133. Callister, W.D.; Rethwisch, D.G. Materials Science and Engineering, 9th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014; 990p. [Google Scholar]
  134. Blyr, A.; Sigala, C.; Amatucci, G.; Guyomard, D.; Chabre, Y.; Tarascon, J.-M. Self-Discharge of LiMn2O4/C Li-Ion Cells in Their Discharged State: Understanding by Means of Three-Electrode Measurements. J. Electrochem. Soc. 1998, 145, 194–209. [Google Scholar] [CrossRef]
  135. Wang, J.; Tang, Y.; Yang, J.; Li, R.; Liang, G.; Sun, X. Nature at LiFePO4 aging process: Roles of impurity phases. J. Power Sources 2013, 238, 454–463. [Google Scholar] [CrossRef]
  136. Cuisinier, M.; Dupré, N.; Moreau, P.; Guyomard, D. NMR Monitoring of Electrode/Electrolyte Interphase in the Case of Air-exposed and Carbon Coated LiFePO4. J. Power Sources 2013, 243, 682–690. [Google Scholar] [CrossRef]
  137. Chen, J.; Bai, J.; Chen, H.; Graetz, J. In Situ Hydrothermal Synthesis of LiFePO4 Studied by Synchrotron X-ray Diffraction. J. Phys. Chem. Lett. 2011, 2, 1874–1878. [Google Scholar] [CrossRef]
  138. Zhu, J.; Fiore, J.; Li, D.; Kinsinger, N.M.; Wang, Q.; DiMasi, E.; Guo, J.; Kisailus, D. Solvothermal Synthesis, Development, and Performance of LiFePO4 Nanostructures. Cryst. Growth Des. 2013, 13, 4659–4666. [Google Scholar] [CrossRef]
  139. Chung, S.-Y.; Kim, Y.-M.; Kim, J.-G.; Kim, Y.-J. Multiphase transformation and Ostwald’s rule of stages during crystallization of a metal phosphate. Nat. Phys. 2009, 5, 68–73. [Google Scholar] [CrossRef]
  140. Malik, R.; Burch, D.; Bazant, M.; Ceder, G. Particle Size Dependence of the Ionic Diffusivity. Nano Lett. 2010, 10, 4123–4127. [Google Scholar] [CrossRef]
  141. Orikasa, Y.; Maeda, T.; Koyama, Y.; Murayama, H.; Fukuda, K.; Tanida, H.; Arai, H.; Matsubara, E.; Uchimoto, Y.; Ogumi, Z. Transient Phase Change in Two Phase Reaction between LiFePO4 and FePO4 under Battery Operation. Chem. Mater. 2013, 25, 1032–1039. [Google Scholar] [CrossRef]
  142. Raj, H.; Rani, S.; Sil, A. Two-Phase Composition (LiFePO4/FePO4) and Phase Transformation Dependence on Charging Current: In Situ and Ex Situ Studies. Energy Fuels 2020, 34, 14874–14881. [Google Scholar] [CrossRef]
  143. Ahsan, Z.; Ding, B.; Cai, Z.; Wen, C.; Yang, W.; Ma, Y.; Zhang, S.; Song, G.; Javed, M.S. Recent Progress in Capacity Enhancement of LiFePO4 Cathode for Li-Ion Batteries. J. Electrochem. Energy Convers. Storage 2021, 18, 010801. [Google Scholar] [CrossRef]
  144. Tolganbek, N.; Yerkinbekova, Y.; Kalybekkyzy, S.; Bakenov, Z.; Mentbayeva, A. Current state of high voltage olivine structured LiMPO4 cathode materials for energy storage applications: A review. J. Alloys Compd. 2021, 882, 160774. [Google Scholar] [CrossRef]
  145. Masias, A.; Marcicki, J.; Paxton, W.A. Opportunities and challenges of lithium ion batteries in automotive applications. ACS Energy Lett. 2021, 6, 621–630. [Google Scholar] [CrossRef]
  146. Mahmud, S.; Rahman, M.; Kamruzzaman, M.; Ali, M.O.; Emon, M.S.A.; Khatun, H.; Ali, M.R. Recent advances in lithium-ion battery materials for improved electrochemical performance: A review. Results Eng. 2022, 15, 100472. [Google Scholar] [CrossRef]
  147. Moniruzzaman, M.; Kumar, Y.A.; Pallavolu, M.R.; Arbi, H.M.; Alzahmi, S.; Obaidat, I.M. Two-Dimensional Core-Shell Structure of Cobalt-Doped@MnO2 Nanosheets Grown on Nickel Foam as a Binder-Free Battery-Type Electrode for Supercapacitor Application. Nanomaterials 2022, 12, 3187. [Google Scholar] [CrossRef]
  148. Kumar, Y.A.; Kim, H.-J. Enhanced electrochemical performance of nanoplate nickel cobaltite (NiCo2O4) supercapacitor applications. RSC Adv. 2019, 9, 1115–1122. [Google Scholar] [CrossRef] [Green Version]
  149. Yang, Z.; Dai, Y.; Wang, S.; Yub, J. How to make lithium iron phosphate better: A review exploring classical modification approaches in-depth and proposing future optimization methods. J. Mater. Chem. A 2016, 4, 18210–18222. [Google Scholar] [CrossRef]
  150. Wen, L.; Guan, Z.; Wang, L.; Hu, S.; Lv, D.; Liu, X.; Duan, T.; Liang, G. Effect of Carbon-Coating on Internal Resistance and Performance of Lithium Iron Phosphate Batteries. J. Electrochem. Soc. 2022, 169, 050536. [Google Scholar] [CrossRef]
  151. Ramasubramanian, B.; Sundarrajan, S.; Chellappan, V.; Reddy, M.V.; Ramakrishna, S.; Zaghib, K. Recent Development in Carbon-LiFePO4 Cathodes for Lithium-Ion Batteries: A Mini Review. Batteries 2022, 8, 133. [Google Scholar] [CrossRef]
  152. Tian, R.; Park, S.; King, P.; Cunningham, G.; Coelho, J.; Nicolosi, V.; Coleman, J. Quantifying the factors limiting rate performance in battery electrodes. Nat. Commun. 2019, 10, 1933. [Google Scholar] [CrossRef] [PubMed]
  153. Newman, J.; Balsara, N.P. Electrochemical Systems, 4th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2021; 608p. [Google Scholar]
  154. Rice, J. Mathematical Statistics and Data Analysis; Brooks/Cole Cengage Learning: Belmont, CA, USA, 2007; p. 138. [Google Scholar]
  155. Jiang, F.; Peng, P. Elucidating the performance limitations of lithium-ion batteries due to species and charge transport through five characteristic parameters. Sci. Rep. 2016, 6, 32639. [Google Scholar] [CrossRef]
  156. Fang, W.; Ramadass, P.; Zhang, Z.J. Study of internal short in a Li-ion cell-II. Numerical investigation using a 3D electrochemical-thermal model. J. Power Sources 2014, 248, 1090–1098. [Google Scholar] [CrossRef]
  157. Davydov, V.V.; Myazin, N.S.; Velichko, E.N. Characteristics of spectrum registration of condensed medium by the method of nuclear-magnetic resonance in a weak field. Tech. Phys. Lett. 2017, 43, 607–610. [Google Scholar] [CrossRef]
  158. Higgins, T.M.; Coleman, J.N. Avoiding resistance limitations in high-performance transparent supercapacitor electrodes based on large-area, high-conductivity PEDOT: PSS films. ACS Appl. Mater. Interfaces 2015, 7, 16495–16506. [Google Scholar] [CrossRef]
  159. Triola, M.F. Elementary Statistics Technology Update, 11th ed.; Pearson Education: London, UK, 2012; 888p. [Google Scholar]
  160. Lvovich, V.F. Distributed Impedance Models. Impedance Spectroscopy: Applications to Electrochemical and Dielectric Phenomena; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2012; 368p. [Google Scholar]
  161. Beckman, I. Higher Mathematics: The Mathematical Apparatus of Diffusion; Yurayt Publishing House: Moscow, Russia, 2020; 459p. [Google Scholar]
  162. Zhang, Y.; Zhang, W.; Shen, S.; Yan, X.; Wu, R.; Wu, A.; Lastoskie, C.; Zhang, J. Sacrificial Template Strategy toward a Hollow LiNi1/3Co1/3Mn1/3O2 Nanosphere Cathode for Advanced Lithium-Ion Batteries. ACS Omega 2017, 2, 7593–7599. [Google Scholar] [CrossRef]
  163. Wang, D.; Zhang, Z.; Hong, B.; Lai, Y. Self-sacrificial organic lithium salt enhanced initial columbic efficiency for safer and greener lithium-ion batteries. Chem. Commun. 2019, 55, 10737–10739. [Google Scholar] [CrossRef]
  164. Yoo, J.W.; Zhang, K.; Patil, V.; Lee, J.T.; Jung, D.-W.; Pu, L.S.; Oh, W.; Yoon, W.-S.; Park, J.H.; Yi, G.-R. Porous supraparticles of LiFePO4 nanorods with carbon for high rate Li-ion batteries. Mater. Express 2018, 8, 316–324. [Google Scholar] [CrossRef]
  165. Churikov, A.V.; Ivanishchev, A.V.; Ushakov, A.V.; Romanova, V.O. Diffusion aspects of lithium intercalation as applied to the development of electrode materials for lithium-ion batteries. J. Solid State Electrochem. 2014, 18, 1425–1441. [Google Scholar] [CrossRef]
  166. Zhang, W.-J. Structure and performance of LiFePO4 cathode materials: A review. J. Power Sources 2011, 196, 2962–2970. [Google Scholar] [CrossRef]
Figure 1. SXRD measurements of the developed and industrial LiFePO4 powders, N1 and N2–4, respectively, for 2 Theta intervals 9.0–11.7 (a) and 17.0–19.1 (b). Arrows near the abscissa indicate the expected peak positions.
Figure 1. SXRD measurements of the developed and industrial LiFePO4 powders, N1 and N2–4, respectively, for 2 Theta intervals 9.0–11.7 (a) and 17.0–19.1 (b). Arrows near the abscissa indicate the expected peak positions.
Energies 16 01551 g001
Figure 2. (a) Galvanostatic charge-discharge curves, (b) variation of capacity with the cycle number. The values of the charge and discharge currents (C rate -units) are on the diagrams.
Figure 2. (a) Galvanostatic charge-discharge curves, (b) variation of capacity with the cycle number. The values of the charge and discharge currents (C rate -units) are on the diagrams.
Energies 16 01551 g002
Figure 4. Sample N1 Raman spectra at 532 nm in the range 800–2100 cm−1 (a) and second ordered RS in the range 1900–3700 cm−1 (b). Decomposition into the Lorentz components with energies in their maxima E1–E7, given in Table 3, is shown.
Figure 4. Sample N1 Raman spectra at 532 nm in the range 800–2100 cm−1 (a) and second ordered RS in the range 1900–3700 cm−1 (b). Decomposition into the Lorentz components with energies in their maxima E1–E7, given in Table 3, is shown.
Energies 16 01551 g004
Figure 5. Experimental MS of the developed LiFePO4 (N1) powder. Upper and lower parts—spectra are recorded with high and low speed of Doppler shift of the gamma-ray source, respectively. The best fitting results of the model spectrum are shown as a solid line.
Figure 5. Experimental MS of the developed LiFePO4 (N1) powder. Upper and lower parts—spectra are recorded with high and low speed of Doppler shift of the gamma-ray source, respectively. The best fitting results of the model spectrum are shown as a solid line.
Energies 16 01551 g005
Figure 6. Scheme of two-stage LiFePO4 cathode degradation. Here, the free energy in the metastable state is ∆G at the battery operating temperatures. With an increased number of cycles, the protective ferric-graphite-graphene layer is destroyed with the corrosion activation energy ∆Gpa; a decrease in this activation energy by ∆Gimpa occurs in the presence of impurities that catalyze the corrosion processes. The second stage is degradation with activation energy of ∆Gcra of unprotected crystallite destruction into an amorphous phase.
Figure 6. Scheme of two-stage LiFePO4 cathode degradation. Here, the free energy in the metastable state is ∆G at the battery operating temperatures. With an increased number of cycles, the protective ferric-graphite-graphene layer is destroyed with the corrosion activation energy ∆Gpa; a decrease in this activation energy by ∆Gimpa occurs in the presence of impurities that catalyze the corrosion processes. The second stage is degradation with activation energy of ∆Gcra of unprotected crystallite destruction into an amorphous phase.
Energies 16 01551 g006
Figure 9. (a) Correlation cloud of transverse and longitudinal sizes Ls and Lb of LiFePO4 for 4326 particles and the corresponding marginal functions of their distributions fitted with Lognormal functions. (b) Particle-density isolines map divided into parts by vertical straight lines with small, intermediate and large particles 1/3 of their total; red dotted lines—smoothed average values. The correlation coefficients of these parts are indicated; the inset shows an enlarged part near the particle density maximum.
Figure 9. (a) Correlation cloud of transverse and longitudinal sizes Ls and Lb of LiFePO4 for 4326 particles and the corresponding marginal functions of their distributions fitted with Lognormal functions. (b) Particle-density isolines map divided into parts by vertical straight lines with small, intermediate and large particles 1/3 of their total; red dotted lines—smoothed average values. The correlation coefficients of these parts are indicated; the inset shows an enlarged part near the particle density maximum.
Energies 16 01551 g009
Figure 10. (a) Q(t) dependences on the mean values of the correlation coefficients r ¯ In the inset, Q(t) dependence on r ¯ at t = 36 s (marked by a vertical dotted line); the red cross marks the sample N1 r ¯ value. A scheme of r ¯ changes during measurements from small to large t is shown. (b) Q(t) dependences on r ¯ normalized to Q0.01 at t = 36 s for the indicated average crystallite sizes; the thick curve and the red cross are for N1 sample.
Figure 10. (a) Q(t) dependences on the mean values of the correlation coefficients r ¯ In the inset, Q(t) dependence on r ¯ at t = 36 s (marked by a vertical dotted line); the red cross marks the sample N1 r ¯ value. A scheme of r ¯ changes during measurements from small to large t is shown. (b) Q(t) dependences on r ¯ normalized to Q0.01 at t = 36 s for the indicated average crystallite sizes; the thick curve and the red cross are for N1 sample.
Energies 16 01551 g010
Table 1. Sequence of LiFePO4 liquid-phase synthesis.
Table 1. Sequence of LiFePO4 liquid-phase synthesis.
StepsReagents
1Preparation of acetatesFe + LiCO3 +CH3COOH → Fe(CH3COO)2, LiCH3COO, H2O
2Organic AdditivesAA, PA
3Phosphoric acidH3PO4
4Pre-drying at 100 °CEvaporation of CH3COOH, H2O
5Annealing 1.5 h at 400 °C in an Ar atmosphereEvaporation of CH3COOH, H2O, CO2, ((CH2))4CO
6Annealing 1.5 h at 670 °C in an Ar atmosphereCrystallization of LiFePO4
Table 2. SXRD composition of impurity crystallites in powders [96], XRD anisotropic sizes of LiFePO4 crystallites L ¯ V h k l averaged over the length of their columns [97] (from [98]). MS fractions of Fe2+ and Fe3+ compounds (description below). Forecast cycling of a test cathode after 100 times cycling at 1 C rate. The errors of L ¯ V h k l reported in parentheses characterize the reproducibility.
Table 2. SXRD composition of impurity crystallites in powders [96], XRD anisotropic sizes of LiFePO4 crystallites L ¯ V h k l averaged over the length of their columns [97] (from [98]). MS fractions of Fe2+ and Fe3+ compounds (description below). Forecast cycling of a test cathode after 100 times cycling at 1 C rate. The errors of L ¯ V h k l reported in parentheses characterize the reproducibility.
Impurity Phase, % Q,
mAh/g
L ¯ V 100 ,   nm L ¯ V 010 , nm L ¯ V 001 , nm Forecast CyclingMössbauer
20 C RateFe2+, %Fe3+, %
N1not detected6366 (5)82 (5)89 (7)3500964
N2Li3PO4, 1.01 (2)58145 (26)131 (13)185 (17)1000955
N3Li3PO4, 2.39 (3)
Fe2P, 2.34 (3)
Fe3P, 2.02 (3)
57141 (5)146 (15)165 (7)5000928
N4not detected40230 (20)261 (8)242 (30)800982
Table 3. The Raman spectra parameters: E1, I1 and W1—position of the maximum, area and width of one Lorentz approximation for the 1st peak and so on for the others shown in Figure 4 for the excitation wavelengths of 532 nm and 633 nm.
Table 3. The Raman spectra parameters: E1, I1 and W1—position of the maximum, area and width of one Lorentz approximation for the 1st peak and so on for the others shown in Figure 4 for the excitation wavelengths of 532 nm and 633 nm.
E1,
I1,
W1
E2,
I2,
W2
E3,
I3,
W3
E4,
I4,
W4
E5,
I5,
W5
E6,
I6,
W6
E7,
I7,
W7
ID/IG (I2/I4)Isp2/Isp3 ((I2 + I4)/(I1 + I3))(I1 + …AI)/
(I5 + …I7)
Excitation 532 nm
N11203
64,354
130
1346
464,462
191
1513
109,518
128
1596
183,271
64
2681
84,869
391
2915
114,677
376
3176
7613
112
2.533.733.96
N41187
45,591
94
1341
733,854
190
1527
130,196
119
1599
244,016
58
2668
146,229
337
2919
231,601
344
3180
17,221
113
3.015.562.92
N21190
22,138
110
1342
288,779
188
1522
53,844
120
1600
98,169
59
2651
37,209
307
2909
88,666
380
3185
5240
117
2.945.093.53
Excitation 633 nm
N11201
63,280
131
1346
475,058
195
1514
106,974
126
1596
182,032
64
2716
65,290
328
2949
38,599
234
3178
9359
128
2.613.867.3
N41193
11,389
112
1332
133,742
180
1528
21,170
123
1602
37,709
56
2630
16,408
293
2883
20,492
318
3178
692
78
3.555.265.42
N21198
7329
117
1331
69,319
179
1516
11,644
127
1601
19,955
57
2594
9304
340
2870
9744
334
3178
350
130
3.474.705.58
Table 4. Parameters of hyperfine interactions obtained by mathematical processing of Mössbauer spectra.
Table 4. Parameters of hyperfine interactions obtained by mathematical processing of Mössbauer spectra.
Line MarkingIS, mm/sQS, mm/sG, mm/sInt, (%)Charge State Fe
Fe2+0.981 ± 0.0012.928 ± 0.0010.281 ± 0.00194.8 ± 0.2Fe2+
11.156 ± 0.0401.818 ± 0.0280.345 ± 0.0251.4. ± 0.3Fe2+
20.172 ± 0.0800.830 ± 0.0800.345 ± 0.0252.6 ± 0.4Fe3+
30.404 ± 0.4000.452 ± 0.0600.345 ± 0.0251.2. ± 0.3Fe3+
Table 5. Results of degradation of the test LiFePO4 cathodes after 100-fold cycling at 1C discharge rate; the extrapolated prediction of their cycling was given above in Table 2. ΔQ—capacity reduction, Fe atomic % obtained using EDX (Section S1), cell volumes before and after cycling, Vb-Va—their changes.
Table 5. Results of degradation of the test LiFePO4 cathodes after 100-fold cycling at 1C discharge rate; the extrapolated prediction of their cycling was given above in Table 2. ΔQ—capacity reduction, Fe atomic % obtained using EDX (Section S1), cell volumes before and after cycling, Vb-Va—their changes.
ΔQ,
mAh/g
EDX, Fe, % L ¯ V X R D ,   nm V = a × b × c, nm3Vb-Va, nm3
BeforeAfterBeforeAfterBeforeAfterBefore–After
N10.713.75.47190 (20)160 (50)291.183291.37−0.09
N24.06.282.98220 (50)210 (50)291.376290.0691.303
N30.47.452.92119 (11)164 (17)290.723290.6330.09
N46.211.1712.5800 (200)150 (20)291.184290.880.304
Table 6. Parameters of N1 and N2 samples. Column 1—parameters of crystallites, average size L ¯ 1 (nm) and variance σ 1 of Lognormal distribution along the [010] axis, etc. and columns 2, 3—parameters for [100], [001] axes. Columns 4–6 are the correlation coefficients, while 7 are their average values. Column 8 shows the total area S of the cross sections of crystallites on the (010) plane, 9 shows the diffusion coefficients, 10 is the electrical relaxation time (see Section 6.2 below).
Table 6. Parameters of N1 and N2 samples. Column 1—parameters of crystallites, average size L ¯ 1 (nm) and variance σ 1 of Lognormal distribution along the [010] axis, etc. and columns 2, 3—parameters for [100], [001] axes. Columns 4–6 are the correlation coefficients, while 7 are their average values. Column 8 shows the total area S of the cross sections of crystallites on the (010) plane, 9 shows the diffusion coefficients, 10 is the electrical relaxation time (see Section 6.2 below).
12345678910
L ¯ 1 ,   σ 1 L ¯ 2 ,   σ 2 L ¯ 3 , σ 3 r12r13r23 r ¯ as, m2D, nm2/sτel, s
N160, 0.4149, 0.4072, 0.380.870.640.730.752.0 × 1070.16 (0.4)8
N292, 0.43108, 0.41160, 0.350.720.560.530.603.1 × 1070.3–2.1 (0.4)20
Table 7. Parameters of LiFePO4 powders: developed, studied industrial and described in the references, as well as comments on the comparison procedure.
Table 7. Parameters of LiFePO4 powders: developed, studied industrial and described in the references, as well as comments on the comparison procedure.
DevelopedIndustrialFrom References
Growth T, °C Two steps
400 °C, 670 °C
unknown[1] 400–800; [3] 700; [5,18] 650; [6] 550; [8] 810; [19,22] 650–700; [24] 670; [33] 550–800; [37] 550; [49,55] 700; [51] 600.
Technologyone-pot liquid-phaseunknown[143] More than 10 types, main: Solvothermal, Hydrothermal, Stripping synthesis, Sol-gel.
Protected layerferric-graphite-graphenemostly ferric-graphite[143] More than 30 types, main: [14] different carbon; [15] nanocarbons; [16,17,18] graphene; [19,20,21] sucrose; [22,23] glucose; [24,25,26] adipic acid; [27,28] polyvinyl alcohol; [29,30] polymeric additive; [31] ferrocene.
Q, mAh/g0.1 C 15110 C 82128–163
72–80
[151] commercial 121–160, best MWCNT, essentially mixed; [143] capacity growth of commercial powder from 160 to 208 mAh/g, more than theoretical 170 mAh/g (*)
Particle sizes, nm L ¯ V 100 66
L ¯ V 010 82
L ¯ V 001 89
141–230
131–261
165–242
[1] 300–7000; [22] 240–3000; [90] 180–300; [118] 500–30,200; [136] 120; [138] 95–280; [141] 60–1000; [144] 40–500; [149] 20–140; [151] 90–300; [164] 30–158. (**)
Cell volumes, 0A [22] 290.63–290.94; [36] 291.08–292.07; [88] 289.7–291.2; [90] 291.3–292.3; [114] 291.33–291.63; [118] 289.8–291.9.
Cycling3500800–1000,
5000 with ferric impurity
[121] 50–600; [143] 50–1000 with different capacity retentions 92–100% (**)
BET, m2/g12.59.4–13.4[1] 1–20; [6] 32–66; [16] 49.3–59.4; [19] 49.6; [121] 15.5; [132] 50; [136] 19–25; [164] 35.
D, nm2/s0.120.25–0.45[3] 2; [6] 2–3; [11,16] 0.01–1; [12] 4.9–7.2; [16] 3.4–8.8, 1.8–1000; [22] 109; [36] 1.2–8.2; [51] 900–2400; [119] 7400–42,000; [128] 0.83–27.3; [149] 1; [151] 0.67–11.7; [165] 1–100; [166] 0.01–10. (***)
τel, s83–30[152] 3–200
tc, s0.50.5[155] 0.1 to >100
Fe3+, %42–8[36] 0.28–0.39; [51] 9–17; [52] 5.16; [54] 2–12; [56] 5–26; [111] 7; [117] 7–25; [119] 5.4–17; [140] 1.13–17
(*) The identification of commercial trends also requires an analysis of market feasibility information, or its public accumulation. An example would be NREL chart www.nrel.gov/pv/cell-efficiency.html (accessed on 14 November 2022) about laboratory solar cells and separately about industrial modules. (**) Difficult to compare because there are no universal certification requirements for measurements. (***) According to [165] using geometrical area of electrode, BET area, particle spherical surface etc. might distort the D values. The used dimension nm2/s is more descriptive in relation to particle sizes [99].
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

Agafonov, D.; Bobyl, A.; Kamzin, A.; Nashchekin, A.; Ershenko, E.; Ushakov, A.; Kasatkin, I.; Levitskii, V.; Trenikhin, M.; Terukov, E. Phase-Homogeneous LiFePO4 Powders with Crystallites Protected by Ferric-Graphite-Graphene Composite. Energies 2023, 16, 1551. https://doi.org/10.3390/en16031551

AMA Style

Agafonov D, Bobyl A, Kamzin A, Nashchekin A, Ershenko E, Ushakov A, Kasatkin I, Levitskii V, Trenikhin M, Terukov E. Phase-Homogeneous LiFePO4 Powders with Crystallites Protected by Ferric-Graphite-Graphene Composite. Energies. 2023; 16(3):1551. https://doi.org/10.3390/en16031551

Chicago/Turabian Style

Agafonov, Dmitry, Aleksandr Bobyl, Aleksandr Kamzin, Alexey Nashchekin, Evgeniy Ershenko, Arseniy Ushakov, Igor Kasatkin, Vladimir Levitskii, Mikhail Trenikhin, and Evgeniy Terukov. 2023. "Phase-Homogeneous LiFePO4 Powders with Crystallites Protected by Ferric-Graphite-Graphene Composite" Energies 16, no. 3: 1551. https://doi.org/10.3390/en16031551

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