Next Article in Journal
Carbon Assessment of Greek Organic Red Wine with Life Cycle Assessment and Planetary Boundaries
Previous Article in Journal
Enhancing Mangrove Aboveground Biomass Estimation with UAV-LiDAR: A Novel Mutual Information-Based Feature Selection Approach
Previous Article in Special Issue
Optimized Production of High-Performance Activated Biochar from Sugarcane Bagasse via Systematic Pyrolysis and Chemical Activation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Environmentally Sustainable Anode Material for Lithium-Ion Batteries Derived from Cattle Bone Waste: A Full-Cell Analysis with a LiFePO4 Cathode

by
Muhammad Shajih Zafar
1,2,
Pejman Salimi
1,3,*,
Marco Ricci
1,
Jasim Zia
4 and
Remo Proietti Zaccaria
1,*
1
Italian Institute of Technology, via Morego 30, 16163 Genoa, Italy
2
Dipartimento di Informatica, Bioingegneria, Robotica e Ingegneria dei Sistemi (DIBRIS), University of Genoa, Via Opera Pia 13, 16145 Genoa, Italy
3
Dipartimento di Chimica e Chimica Industriale, University of Genoa, Via Dodecaneso, 31, 16146 Genoa, Italy
4
Department of Engineering for Innovation, Campus Ecotekne, University of Salento, Via Per Monteroni, 73100 Lecce, Italy
*
Authors to whom correspondence should be addressed.
Sustainability 2025, 17(7), 3005; https://doi.org/10.3390/su17073005
Submission received: 1 March 2025 / Revised: 21 March 2025 / Accepted: 25 March 2025 / Published: 28 March 2025
(This article belongs to the Special Issue Biomass Transformation and Sustainability)

Abstract

:
Modern society relies heavily on energy, driving global research into sustainable energy storage and conversion technologies. Concurrently, the increasing volume of waste generated by industrial and commercial activities emphasizes the need for effective waste management strategies. Carbonization emerges as a promising solution, converting waste into energy and valuable end products such as biochar. This study explores an approach for valorizing bone-based food waste, presenting innovative pathways for managing the escalating issue of food waste. We investigate carbon derived from cattle bone waste, carbonized at 800 °C (CBW8), to design sustainable full-cell lithium-ion batteries (FLIBs). FLIBs featuring CBW8 as the anode material and LiFePO4 as the cathode exhibit exceptional cycling life, even at high current rates. The cell demonstrates a high specific capacity of 165 mAh g−1 at 0.5 C, maintaining stable performance over 1800 cycles at various C-rates. This work not only advances the field of sustainable energy and waste management, but also opens new avenues for eco-friendly technological applications.

1. Introduction

In 2050, the worldwide demand for energy is projected to surge by approximately 50%, driven by factors such as population growth, urbanization, and the effects of climate change. According to the International Renewable Energy Agency (IRENA) report, achieving the 1.5 °C (e.g., limit of global temperature rise) scenario of global net zero CO2 emissions by 2050 requires a significant expansion of renewables across all sectors [1]. This entails a twelve-fold increase in renewable electricity capacity compared to 2020, necessitating an additional 1066 GW annual capacity production from 2023 onward. In this scenario, bioenergy—renewable energy derived from organic materials like biomass and capable of being converted into liquid fuels, heat, and electricity [2]—is projected to play a significantly larger role in the global energy mix, contributing an estimated 16% of total final energy consumption by 2050 [1]. Bioenergy’s share in the total energy mix is projected to rise significantly, with the demand expected to reach 56 exajoules (EJs) by 2050. This increase is crucial for decarbonizing sectors such as industry and transport, where the infrastructure of electrification is challenging. The bioenergy demand will be split between energy production (including power and heat generation) and as a feedstock for producing bio-based chemicals and materials [3,4,5].
The high potential of various types of biowastes for application in advanced energy storage systems has been well-documented [6,7,8]. For instance, carbon derived from sustainable resources has shown promise as a replacement for standard graphite in lithium-ion batteries (LIBs) [9,10]. Additionally, in supercapacitors, plant-based biowaste materials such as industrial hemp and honeydew peel can act as sustainable carbon precursors, inherently introducing heteroatom doping for enhanced electrochemical performance [11,12]. The unique properties of biochar, including its non-graphitic structure, porosity, and abundant presence of heteroatoms, make it an ideal candidate for electrodes in secondary batteries, particularly LIBs [13,14]. Life cycle assessment studies highlight that using waste materials for biochar production yields significant environmental benefits and aligns with circular economy principles, promoting resource efficiency and sustainability [15].
Animal bone is one of the most abundant biowaste by-products, with the global meat industry generating 130 billion kg of animal bones annually, of which the European Union contributes more than 10% [16]. Bones are composed of hydroxyapatite (HA) and collagen [17], which can be converted through carbonization into porous carbon materials with high numbers of defects [13]. These materials have potential applications in various fields, such as for inducing oxygen reduction reactions [18], supercapacitors [19,20], batteries [21,22], and water treatment [23,24].
In the specific sector of batteries, among the different chemistries nowadays commercially available, batteries based on lithium iron phosphate (LiFePO4: LFP) cathodes present the advantage of being more eco-friendly than other kinds of commercial batteries, as they eliminate the need for critical raw materials such as manganese, cobalt, and nickel. LFP-based batteries are widely favored by Chinese automakers due to their safety, technological independence, and abundant local raw materials [25]. These batteries not only offer a longer cycle life, lower costs, and reduced environmental impact but also excel in rapid charge and discharge capabilities. Additionally, LFP batteries provide high power density, high voltage, high energy density, low self-discharge rate, and minimal heating. These attributes make LFP-based batteries suitable for a wide range of applications [26], to the extent that the global market for batteries utilizing LFP cathodes was valued at USD 15.28 billion in 2023, with a growth projection reaching the remarkable amount of USD 124.42 billion by 2032 [27].
In the current study, cattle bone waste carbonized at 800 °C was used instead of graphite as the anode material and coupled with an LFP cathode to design full-cell lithium-ion batteries (FLIBs). The designed FLIBs exhibited promising electrochemical performance, with specific capacities of 155, 140, 95, and 65 mAh/gLFP during long cycling at 1, 2, 5, and 10 C, respectively. Unlike previous studies [22,28] that predominantly focused on half-cell utilization of (cattle bone) biomass, this research demonstrates the true potential of waste materials in real-world applications.

2. Materials and Methods

2.1. Materials

Cattle bone waste was taken from a meat market in Genoa, Italy. The mineral oil (μ: 41–17 cSt), castor oil (μ: 13.5 cSt), silicone oil (μ: 500 cSt) and polyvinylidene fluoride (PVDF)were purchased from Sigma Aldrich. Carboxymethyl cellulose (CMC) and carbon black (Super P) were purchased from GELON Lib Group Co., Ltd. Lithium iron phosphate (LFP: LiFePO4) were purchased from NEI corporation USA. Soybean oil (SALVADORI, μ: 43 cSt) was purchased from a local market. Milli-Q water was used for all the experiments.

2.2. Cattle Bone Waste (CBW) Biochar Preparation

To remove any trace amounts of meat and other organic impurities, the bone was immersed in boiling water for a short amount of time. After being cleaned, the bone was placed in a sealed plastic container and kept frozen at −18 °C for eventual use. The bone was divided into smaller pieces and placed in a tube furnace (three-zone split furnace: PSC 12/600H, Lenton, UK) for carbonization in an N2 atmosphere at the temperature of 800 °C with a heating rate of 5 °C/min and held for 1 h. After this treatment, the produced carbonized bone waste was cooled naturally down to room temperature. The powder yield of carbon is a.u. 74%. Finally, the carbonized bone waste was grounded and sieved to less than 100 micrometers and then washed with 1 M HCl and deionized water to remove possible contaminants, with a final drying overnight step run at 110 °C. In this work, we shall refer to the final cattle bone waste obtained by following the aforementioned procedure as CBW8. Figure 1 illustrates the complete process for producing biochar (CBW8) from cattle bone waste.

2.3. CBW8 Anode and LFP Cathode Electrode Preparation

The CBW8 electrode was fabricated using an environmentally friendly process involving CMC and deionized water. These materials were chosen as replacements for the toxic N-methyl-2-pyrrolidone (NMP) and biologically hazardous PVDF binder. To prepare the anode electrode, CBW8 powder (70 wt%) was mixed with 20 wt% conductive carbon black and 10 wt% CMC. After grinding, the mixture was dispersed in deionized water to form a homogeneous slurry. This slurry was then casted onto a copper current collector (10 µm thickness) using a doctor blade and dried on a hot plate for 3 h at 60 °C. The electrodes were subsequently cut into 12 mm diameter disks and further dried using a Buchi apparatus for 4 h at 80 °C. The achieved active mass loading of the electrodes was approximately 1–1.5 mg/cm2.
The cathode electrode was fabricated by mixing commercial LFP, carbon black (Super P), and PVDF with a weight ratio of 80:10:10 in NMP. Then, the cathode slurry was coated on the aluminum current collector. The active mass loading of the LFP electrode was about 1.8–2 mg cm−2 after drying.
For the full-cell fabrication, the CBW8 anode was prelithiated in the presence of a LP30 (1 M LiPF6 in EC: DMC) carbonate electrolyte using the lithium direct contact electrochemical method [8]. The procedure for anode prelithiation and capacity balancing of cathode and anode is reported in our previous work [29]. Finally, the CBW8 anode and the LFP cathode were both combined with a Celgard 2400 separator and LP30 electrolyte into a 2032-coin cell. The half-cell and full-cell analyses were conducted using the same electrolyte and separator for consistency.

2.4. Material Characterizations

A scanning electron microscope (SEM, JEOL JSM-6490LA) was used to investigate the morphology of CBW8. Firstly, the sample was coated with a gold 10 nm thick gold layer by a Cressington 208HR sputter coater (Cressington Scientific Instrument Ltd., UK). Furthermore, to study the thermal profiles of CBW8, thermogravimetric analyses (TGAs) using a Q500 analyzer (TA instrument, USA) were employed with a heating rate of 10 °C/min in a range of 30–800 °C and N2 atmosphere. A Fourier Transform Infrared (FTIR) spectrometer (Vertex 70v FT-IR, Bruker) with a single-reflection attenuated total reflection accessory (MIRacle ATR, PIKE Technologies) was then used to determine the functional groups in the CBW8. The spectra presented were taken in the average of 128 repetitive scans with a resolution of 4 cm−1 in the range of 4000–600 cm−1. A Renishaw micro-Raman spectrometer with an excitation wavelength of 514 nm, line of an Ar+ laser, a ×50 objective (numerical aperture: 0.75), and an incident power lower than 1 mW was instead employed for Raman analysis. A PANalytical Empyrean X-ray diffractometer with a 1.8 kW CuKα ceramic X-ray tube and a PIXcel3D 2 × 2 area detector with operating conditions of 45 kV and 40 mA was used to acquire the X-ray diffraction (XRD) response from CBW8. Furthermore, the specific surface area and pore size distribution of the CBW8 in the micro- and mesoporous range were conducted using Brunauer–Emmett–Teller (BET) and Barrette Joynere Halenda (BJH) analyses. For this purpose, under vacuum for 180 min, the sample was initially degassed at 150 °C to remove any water molecules potentially absorbed by the CBW8 material. Then, a gas sorption analysis based on nitrogen physisorption measurements was realized through an Autosorb-iQ (Quantachrome Instruments) analyser at the temperature of 77 K. By taking 20 uniformly spaced points in the relative pressure (P/P0) range from 0.05 to 0.3, the multipoint BET model was used to measure the specific surface areas of the CBW8. With the support of the BJH model, pore size distribution by considering 13 points was determined from the desorption isotherms (range 0.35 < P/P0 < 1). X-ray photoelectron spectroscopy (XPS) was then performed to analyze the chemical bonds in the CBW8 surface. The employed system was composed of a monochromatic X-ray source (set at 1486 eV), a hemispherical energy analyzer (Phoibos, HSA3500, also from Specs), and an electron spectrometer (Lab2, Specs, Berlin, Germany). The applied voltage, applied current, and pressure were chosen as 13 kV, 8 mA, and ≈1 × 10−9 millibar, respectively. For both wide and narrow scans, the large area lens mode was adopted. Furthermore, for the wide scan, an energy pass of 90 eV and an energy step of 1 eV were employed, while an energy pass of 30 eV and an energy step of 0.1 eV were instead set for the narrow high-resolution scan. Finally, the surface charge, having an energy of 7 eV and a filament current of 2.2 A, was neutralized using a flood gun.

2.5. Electrochemical Measurements

Room temperature cyclic voltammetry (CV) and galvanostatic charge and discharge analyses of the Li/CBW8 half-cell and CBW8/LFP full-cells were performed in the voltage ranges of 0.01–3 V and 0.8–3.9 V, respectively. Electrochemical impedance spectroscopy (EIS) was also performed in the frequency range from 10 kHz to 100 mHz at an open circuit potential with a 10 mV amplitude. BCS-805 BioLogic was employed for conducting all the electrochemical tests.

3. Results

3.1. Structural Characterization

SEM images of CBW8 were taken to assess the complete bone structure that resulted from the carbonization process (Figure 2a and Figure S1). Pure bone exhibits only a limited number of intrinsic pores prior to carbonization (Supporting Information, Figure S2). These pores are densely packed due to the presence of organic substances, such as collagen, proteins, lipids, and water, which occupy the pore spaces within the bone structure [19]. However, as a result of the release of the stored water and organic chemicals during the carbonization process, highly interconnected microchannels are created (Figure 2a). In this respect, the inner morphology of CBW8 is characterized also by the presence of pores in the micrometric range, as seen by the cross-section SEM images of Figure 2a. The elongated microchannels that run along the axis of most of the bone structure and associated pores are also observed. Additionally, the surface morphology of the microchannel walls also changes; it becomes rough and apparent that fiber-like structures are adorning the pore walls, which might be connected to the HA (hydroxyapatite) crystallite presence. SEM imaging of the CBW8 as a whole shows a structure with an intense network of interconnecting microchannels and micrometric pores. Figure S3 presents the SEM images of powder CBW8. For a better understanding of the pore size and distribution, BET analysis through the N2 adsorption/desorption method was adopted. The results, shown in Figure 2b, illustrate pores with different size distributions in a specific pore diameter range in CBW8. Therefore, it exhibits a specific surface area of 136.4 m2g−1, a mean pore volume of 0.319 cm3g−1, and a mean pore diameter of 3.839 nm in CBW8. Figure 2c depicts N2 adsorption/desorption isotherms at relative pressure (P/P0). The isotherm is categorized as Type IV, which is a feature of materials with mesopores (2–50 nm) [30]. Furthermore, according to the IUPAC, the narrow hysteresis loop shown in the isotherm is of the H3 type, which is linked to mixed-shaped pores (cylindrical, wedge, and slit-shaped) and is typical of N2 condensation in the mesopore area [31,32].
Thermogravimetric analysis (TGA) and first derivative TGA (DTGA) were used to examine the weight loss and thermal stability of the CBW8 during the heat treatment, as shown in Figure 3a and Figure S4. The analysis reveals an overall weight loss of 6.1%, confirming that CBW8 is thermally stable, with no significant organic residues remaining. This reduced weight loss reflects the material’s high purity. The primary weight loss of 2.7% up to 100 °C is attributed to the release of adsorbed moisture. Beyond this, up to 800 °C, the remaining weight loss is associated with the thermolysis of carbonates and partial dehydroxylation of HA [23].
A further important analysis that has been considered is about the verification of the chemical elements and organics located at the CBW8 surface. This study was run through X-ray photoelectron spectroscopy (XPS), with the results shown in Figure 3b–d and Figure S5. The spectra show the presence of oxygen (O), carbon (C), phosphorus (P), and calcium (Ca), with atomic percentages equal to 39.17%, 37.82%, 9.72%, and 13.28%, respectively (Figure S5 and Table S1). Furthermore, the peak area percentages of the four C1s subcomponents (C-C, C-O, C=O, and CaCO3) are 76.31%, 12.93%, 6.21%, and 4.55%, respectively (Table S2). The subcomponents of O1s include carbonates, organic oxygen, and P-O, with percentages of 84.33%, 6.82%, and 8.84%, respectively (Table S3). It should be noted that the high content of oxygen in the CBW8 material contributes to an increase in hydrophilicity in air [33]. These XPS results prove that at 800 °C, all the organic components are removed.
The next step was to gain insights about the presence of functional groups in the CBW8 material. For this reason, Fourier Transform Infrared Spectroscopy (FTIR) was performed, with the results shown in Figure 4a. The two broad peaks at 3281 cm−1 and 1642 cm−1 correspond to the stretching vibration of O-H, and they are related to the adsorbed moisture [34,35]. On the other hand, the dominating band at 1007 cm−1 and the band at 961 cm−1 are associated with the stretching vibration of the phosphate group, while the bands at 1456, 1414, and 872 cm−1 are related to the stretching vibrations of the carbonate group (CO32−) [36], the two components of HA (Cax(PO4, CO3)y(OH)). The peaks ascribed to the phosphate and carbonate groups (1456, 1414, 1007, 961, and 872 cm−1), provide evidence of HA existence. Another X-ray-based technique was employed to verify the degree of crystallinity of CBW8. Specifically, the X-ray diffraction (XRD) spectrum of CBW8 is shown in Figure 4b. According to the reference card (ICDD-01-074-9761), all detected peaks are ascribed to the characteristic peaks of standard HA [34]. The samples’ crystallinity, which results in the removal of collagen and other organic components without altering the HA structure and reveals an inorganic structure, is shown by the clear and strong peaks of the CBW8 spectra [37].
Finally, the level of graphitization of CBW8 was investigated through Raman spectroscopy, as shown in Figure 4c. Two sharp peaks at 1356 cm−1 (ID) and 1590 cm−1 (IG), respectively attributable to amorphous carbon (ID) (i.e., characteristic peak of sp3 defects) and graphite structure (IG) (sp2 bonded graphitic carbon), show the strong vibrations of the C-C of the carbonaceous material [38]. In particular, an IG/ID ratio of 1.18 is obtained, denoting a moderate level of graphitization [23,39] This influences both lithium-ion storage capacity and cycling stability. Moderate graphitization enhances conductivity and provides structural stability, facilitating reversible lithium-ion intercalation while maintaining long-term cycling performance [40]. This balance is crucial for maintaining both high capacity and stable cycling behavior over extended cycles.

3.2. Electrochemical Performance of CBW8 in Half-Cell and Full-Cell Configurations

The electrochemical activity of CBW8 was initially investigated in a lithium half-cell configuration (Figure 5). Afterwards, the potential application of CBW8 as anode in full-cell lithium-ion batteries (FLIBs) paired with an LFP cathode was explored. The half-cell results show that CBW8 could provide a robust capacity retention across various current densities, yielding capacities of 900, 720, 470, and 410 mAh g−1 at the current densities of 0.1, 0.2, 0.5, and 0.8 A g−1, respectively (see Figure 5a,b). In Figure 5b, a marginal capacity decrease is observed when the current density returns to 0.1 A g−1, indicating a good reversibility of the engineered electrode. Furthermore, the charge–discharge cycling also demonstrated an initial discharge capacity as high as 2300 mAh g−1, with an initial coulombic efficiency (ICE) of 37% at the current rate of 0.1 A g−1 (Figure 5b). This low ICE value is synonymous with the presence of a solid electrolyte interphase (SEI), an insulating film which is formed during the first reduction process, a typical characteristic of turbostratic porous carbon materials [41,42]. The half-cell was also tested for over 600 cycle at 0.5 A g−1, in order to evaluate its long cycling stability. The results demonstrate a remarkable average specific capacity of 417 mAh g−1 with 84% capacity retention throughout cycling (see Figure 5c). It is noted that the presence of oxygen groups in the electrode promotes rapid redox processes [43]. Moreover, the observed capacity and electrochemical stability of Li/CBW8 batteries can be attributed to the high surface area, porous architecture, and the presence of heteroatoms in the CBW8 electrode. On the other hand, the capacity fade of CBW8 in Li-metal half-cells is mainly due to irreversible lithium consumption, excessive SEI formation, and structural degradation induced by oxygen-rich functional groups. Furthermore, electrolyte decomposition and gas evolution exacerbate performance degradation.
To contextualize CBW8 performance, it is essential to compare it with conventional anodes such as graphite- and silicon-based materials. Graphite, the most widely used anode in LIBs, has a theoretical capacity of 372 mAh g−1 but is limited by sluggish kinetics and moderate rate capability [44]. Silicon-based anodes, while offering an exceptionally high theoretical capacity (~3590 mAh g−1), suffer from severe volumetric expansion and rapid capacity degradation [45]. In contrast, CBW8 not only delivers a comparable or slightly higher capacity than graphite but also exhibits remarkable cycling stability, making it a promising alternative. Additionally, its biomass-derived origin and straightforward synthesis process provide key advantages in terms of sustainability and cost-effectiveness, reinforcing its potential for practical energy storage applications.
Figure 5d describes the cyclic-voltammetry (CV) behavior of the half-cell at various scan rates. The well pronounced broad peaks are a characteristic of disordered carbon, a typical situation for lithium batteries [46]. During the lithiation process, a cathodic peak associated with the lithiation process is observed at about 0.1 V, while an anodic peak related to the delithiation process is observed at 0.2 V [47]. Furthermore, the figure also shows an increase in electrode current while the scan rate rises. This behavior is commonly attributed to the dynamic electrochemical processes occurring at the electrode–electrolyte interface. At higher scan rates, the charge/discharge processes are more limited by the available surface area, and the current response becomes more dependent on the electrochemical surface area (ECSA). The increase in current with higher scan rates is likely indicative of an increase in the effective electroactive surface area. This is often seen with materials exhibiting high porosity and complex microstructures, such as biomass-derived electrodes. As the scan rate increases, the rate of ion diffusion to the surface becomes faster, and the electrode’s surface reactivity plays a more significant role in determining the overall current response. For biomass anodes, this phenomenon may also be due to the complex porous structure of the material, which offers more accessible sites for ion interaction as the scan rate increases, thereby leading to a higher apparent surface area. However, this increase in surface area is not necessarily indicative of a true physical increase in the material’s surface but rather reflects the enhanced electrochemical response due to the increased interaction between the electrode surface and the electrolyte ions at higher scan rates.
One of the foremost challenges encountered by FLIBs is their capacity degradation occurring during storage and charge–discharge cycles. This degradation may manifest in various forms, including diminished coulombic efficiency (CE), reduced cell usable capacity, and increased cell impedance. Throughout the charge and discharge processes, the diffusion of lithium atoms in and out of the host electrode may determine diffusion-induced stresses and volume fluctuations within the electrode, which may result in SEI fractures [48]. In this work, prior to cell assembly, the CBW8 electrode underwent the lithiation process following the procedure outlined in the experimental section. This lithiation step facilitates the formation of an SEI layer on the anode surface, mitigating lithium-ion loss during the initial discharge of the full-cell. Another action that was taken to ensure a better cell performance was to establish a proper balance between the anode and cathode. Varying their ratio significantly impacts the cell voltage and delivered capacity. Precise balancing is imperative to avoid battery overcharge and anode plating, which can compromise safety and diminish the cycle life. In this research, optimal results, characterized by a high ICE of 80% and stable capacity, were achieved with an anode loading of 1.3 mg (active mass loading) coupled with an LFP cathode of 2 mg.
Figure 6a illustrates the CV profile for the first three cycles of the prelithiated CBW8/LFP full-cell at the scan rate of 0.1 mV s−1 within the potential range 0.8–3.9 V. During the positive voltage sweep (at ~3.6 V), lithium ions enter into the biochar electrode, while lithium-ion extraction occurs within the LFP cathode. Conversely, during the negative sweep (at ~3.3 V), lithium ions intercalate into the LFP cathode and de-intercalate from the CBW8 anode. Similarly, during the second and third cycles, the peaks at ~2.8/2.9 V during charging and ~2.5 V during discharging may stem from the two-step process of lithium-ion insertion/extraction into/from the electrodes [49]. The voltage differential of 0.3 V between the positive and negative peaks is indicative of electronic polarization. Figure 6b illustrates the extensive cycling tests of the full-cell, with varied C-rates ranging from 1 C to 10 C, then reverting to 1 C to assess the cell stability. The observed decrease in capacity at higher applied currents is attributed to increased polarization. The cell exhibits an initial discharge capacity (based on cathode active mass) of approximately 160 mAh g−1 at 1 C, closely approaching the theoretical capacity of LFP versus Li/Li+ (~170 mAh g−1). Upon increasing the C-rate to 2 C, a minimal capacity drop is observed. Furthermore, transitioning from 2 C to 5 C results in a capacity reduction of 25%, with a similarly marginal capacity change noted when shifting from 5 C to 10 C. Notably, the FLIBs demonstrate a good reversible capacity by reaching 140 mAh g−1 when returning at 1 C after 1750 cycles. Additionally, the cell exhibits an average CE of 99.62%.
The galvanostatic charge and discharge profiles (Figure 6c) encapsulate a composite representation of the anode and cathode profiles in full-cell configuration versus Li/Li+. The specific capacity decreases with increasing C-rate, reflecting the expected capacity fade at higher charge/discharge rates. At lower C-rates, the cell achieves capacities close to the theoretical value of LFP. The voltage plateau around 3.4 V during charging and 3.3 V during discharging corresponds to the Fe2+/Fe3+ redox reaction in the LFP cathode. Furthermore, Figure 6d, along with Figure S6, reports the cycling performance of the FLIBs in terms of gravimetric energy density and specific capacity of the CBW8/LFP cell at 0.5 C, respectively. Considering the active mass of the cathode only, the resultant gravimetric energy density of the cell stands at 480 Wh kg−1. Conversely, the cell gravimetric energy density is 295 Wh kg−1 when considering the weight of both electrodes (active mass). The electrochemical impedance spectroscopy (EIS) analysis of the CBW8/LFP full-cell has also been explored and documented in the Supporting Information (Figure S7). Notably, a negligible increase in internal resistance is observed after cycling, suggesting a properly balanced and performing cell. This is closely aligned with the results reported by Xu et al. [50]. They posited that the minimal capacity loss observed in cells with a higher capacity anode may be elucidated by two potential stress-related mechanisms: (i) with increased anode loading, the current density on the anode surface diminishes, potentially mitigating electrode particle cracking throughout cycling; and (ii) elevated anode loading diminishes anode utilization, thereby limiting electrode particle expansion and contraction during lithiation and delithiation, thereby alleviating mechanical degradation [50].
Importantly, the attained energy and reversibility of the CBW8/LFP full-cell under high current rates surpass what is reported for similar cells in the existing literature. In this respect, a comparison of the results obtained in this work with the main electrochemical properties of LFP-based LIBs recently reported in the literature is shown in Table 1.

4. Conclusions

In this study, we convincingly demonstrate the high potential of food waste as an economical, abundantly available, and scalable bio-based material for applications in the energy sector. The synergistic behavior of biowaste, derived from a simple carbonization process, exhibits inherent heteroatoms, high surface area, interlinked microchannels, and a hierarchical porous structure. These properties contribute to high storage capacity in lithium-ion batteries (LIBs). Our investigation reveals that full-cell LIBs using cattle bone waste (CBW) as the anode, coupled with LiFePO4 (LFP) as the cathode, maintain very stable capacity over prolonged cycling. This underscores the high potential of biowaste in Li intercalation/de-intercalation processes.
This work focuses on the practical application of this unique porous mechanism in real-world LIBs. This approach aligns with the United Nation’s Sustainable Development Goals (SDGs), advancing efforts toward sustainable development.

Supplementary Materials

The following Supporting Information can be downloaded at https://www.mdpi.com/article/10.3390/su17073005/s1, Figure S1. Outer and top surface SEM images of CBW8. Figure S2. SEM images of the cross-sectional view of the pristine bone were taken at different magnifications. Figure S3. SEM images of CBW8 powder. Figure S4. DTGA of CBW8. Figure S5. XPS of CBW8 (a) P2p and (b) Ca2p. Figure S6. Cycling performance of the pre-lithiated CBW8/LFP full-cell at 0.5 C (1C = 170 mA/gLFP) in the potential range o0.8–3.9 V. The analyses were performed at room temperature. Electrolyte: LP30. Figure S7. Nyquist plot CBW8/LFP full cell at fresh state, and after 100 cycles at 0.5 C. Table S1. The atomic percentage of CBW8 determined by XPS analysis. Table S2. Peak area (%) of the XPS C1s sub-components of CBW8. Table S3. Peak area (%) of the XPS O1s sub-components of CBW8. References [58,59,60,61] are cited in Supplementary Materials.

Author Contributions

M.S.Z. and P.S.: conceptualization, methodology, writing original draft, review and editing, data curation, investigation, and visualization. M.R. and J.Z.: methodology. P.S. and R.P.Z.: supervision, validation, resources, and review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors would like to thank Lara Marini for the TGA analysis, Riccardo Carzino for the XPS analysis, and Lea Pasquale for the BET analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. International Renewable Energy Agency (IRENA). Available online: https://www.irena.org/Digital-Report/World-Energy-Transitions-Outlook-2023 (accessed on 28 February 2025).
  2. International Energy Agency (Bioenergy). Available online: https://www.iea.org/energy-system/renewables/bioenergy (accessed on 28 February 2025).
  3. Cherubini, F. The Biorefinery Concept: Using Biomass Instead of Oil for Producing Energy and Chemicals. Energy Convers. Manag. 2010, 51, 1412–1421. [Google Scholar] [CrossRef]
  4. Papadis, E.; Tsatsaronis, G. Challenges in the Decarbonization of the Energy Sector. Energy 2020, 205, 118025. [Google Scholar] [CrossRef]
  5. Reijnders, L. Conditions for the Sustainability of Biomass Based Fuel Use. Energy Policy 2006, 34, 863–876. [Google Scholar] [CrossRef]
  6. Plavniece, A.; Volperts, A.; Dobele, G.; Zhurinsh, A.; Kaare, K.; Kruusenberg, I.; Kaprans, K.; Knoks, A.; Kleperis, J. Wood and Black Liquor-Based N-Doped Activated Carbon for Energy Application. Sustainability 2021, 13, 9237. [Google Scholar] [CrossRef]
  7. Mensah-Darkwa, K.; Zequine, C.; Kahol, P.K.; Gupta, R.K. Supercapacitor Energy Storage Device Using Biowastes: A Sustainable Approach to Green Energy. Sustainability 2019, 11, 414. [Google Scholar] [CrossRef]
  8. Salimi, P.; Venezia, E.; Taghavi, S.; Tieuli, S.; Carbone, L.; Prato, M.; Signoretto, M.; Qiu, J.; Proietti Zaccaria, R. Lithium-Metal Free Sulfur Battery Based on Waste Biomass Anode and Nano-Sized Li2S Cathode. Energy Environ. Mater. 2022, 7, e12567. [Google Scholar] [CrossRef]
  9. Salimi, P.; Askari, K.; Norouzi, O.; Kamali, S. Improving the Electrochemical Performance of Carbon Anodes Derived from Marine Biomass by Using Ionic-Liquid-Based Hybrid Electrolyte for LIBs. J. Electron. Mater. 2019, 48, 951–963. [Google Scholar] [CrossRef]
  10. Dai, Z.; Hou, H.; Liu, X.; Yao, Y.; Liao, Q.; Yu, C.; Li, D. Feasibility of Expired Waste Aspirin for Use as Lithium-Ion Battery Anode. Waste Biomass Valorization 2020, 11, 357–365. [Google Scholar] [CrossRef]
  11. Minakshi, M.; Mujeeb, A.; Whale, J.; Evans, R.; Aughterson, R.; Shinde, P.A.; Ariga, K.; Shrestha, L.K. Synthesis of Porous Carbon Honeycomb Structures Derived from Hemp for Hybrid Supercapacitors with Improved Electrochemistry. Chempluschem 2024, 89, e202400408. [Google Scholar] [CrossRef]
  12. Minakshi, M.; Samayamanthry, A.; Whale, J.; Aughterson, R.; Shinde, P.A.; Ariga, K.; Kumar Shrestha, L. Phosphorous—Containing Activated Carbon Derived From Natural Honeydew Peel Powers Aqueous Supercapacitors. Chem. Asian J. 2024, 19, e202400622. [Google Scholar] [CrossRef]
  13. Yao, Y.; Wu, F. Naturally Derived Nanostructured Materials from Biomass for Rechargeable Lithium/Sodium Batteries. Nano Energy 2015, 17, 91–103. [Google Scholar] [CrossRef]
  14. Norouzi, O.; Salimi, P.; Di Maria, F.; Pourhosseini, S.E.M.; Safari, F. Synthesis and Design of Engineered Biochars as Electrode Materials in Energy Storage Systems BT. In Production of Materials from Sustainable Biomass Resources; Fang, Z., Smith Richard, L.J., Tian, X.-F., Eds.; Springer: Singapore, 2019; pp. 233–265. ISBN 978-981-13-3768-0. [Google Scholar]
  15. Marzeddu, S.; Cappelli, A.; Ambrosio, A.; Décima, M.A.; Viotti, P.; Boni, M.R. A Life Cycle Assessment of an Energy-Biochar Chain Involving a Gasification Plant in Italy. Land 2021, 10, 1256. [Google Scholar] [CrossRef]
  16. Alexa Teigiserova, D.; Hamelin, L.; Thomsen, M. Towards Transparent Valorization of Food Surplus, Waste and Loss: Clarifying Definitions, Food Waste Hierarchy, and Role in the Circular Economy. Sci. Total Environ. 2020, 706, 136033. [Google Scholar] [CrossRef]
  17. Niu, J.; Shao, R.; Liu, M.; Zan, Y.; Dou, M.; Liu, J.; Zhang, Z.; Huang, Y.; Wang, F. Porous Carbons Derived from Collagen-Enriched Biomass: Tailored Design, Synthesis, and Application in Electrochemical Energy Storage and Conversion. Adv. Funct. Mater. 2019, 29, 1905095. [Google Scholar] [CrossRef]
  18. Wang, R.; Wang, K.; Wang, Z.; Song, H.; Wang, H.; Ji, S. Pig Bones Derived N-Doped Carbon with Multi-Level Pores as Electrocatalyst for Oxygen Reduction. J. Power Sources 2015, 297, 295–301. [Google Scholar] [CrossRef]
  19. Niu, J.; Shao, R.; Liu, M.; Liang, J.; Zhang, Z.; Dou, M.; Huang, Y.; Wang, F. Porous Carbon Electrodes with Battery-Capacitive Storage Features for High Performance Li-Ion Capacitors. Energy Storage Mater. 2018, 12, 145–152. [Google Scholar] [CrossRef]
  20. Shan, B.; Cui, Y.; Liu, W.; Zhang, Y.; Liu, S.; Wang, H.; Sun, L.; Wang, Z.; Wu, R. Fibrous Bio-Carbon Foams: A New Material for Lithium-Ion Hybrid Supercapacitors with Ultrahigh Integrated Energy/Power Density and Ultralong Cycle Life. ACS Sustain. Chem. Eng. 2018, 6, 14989–15000. [Google Scholar] [CrossRef]
  21. Wei, S.; Zhang, H.; Huang, Y.; Wang, W.; Xia, Y.; Yu, Z. Pig Bone Derived Hierarchical Porous Carbon and Its Enhanced Cycling Performance of Lithium–Sulfur Batteries. Energy Environ. Sci. 2011, 4, 736–740. [Google Scholar] [CrossRef]
  22. Niu, J.; Shao, R.; Liang, J.; Dou, M.; Li, Z.; Huang, Y.; Wang, F. Biomass-Derived Mesopore-Dominant Porous Carbons with Large Specific Surface Area and High Defect Density as High Performance Electrode Materials for Li-Ion Batteries and Supercapacitors. Nano Energy 2017, 36, 322–330. [Google Scholar] [CrossRef]
  23. Zafar, M.S.; Zahid, M.; Athanassiou, A.; Fragouli, D. Biowaste-Derived Carbonized Bone for Solar Steam Generation and Seawater Desalination. Adv. Sustain. Syst. 2021, 5, 2100031. [Google Scholar] [CrossRef]
  24. Shahid, M.K.; Kim, J.Y.; Choi, Y.G. Synthesis of Bone Char from Cattle Bones and Its Application for Fluoride Removal from the Contaminated Water. Groundw. Sustain. Dev. 2019, 8, 324–331. [Google Scholar] [CrossRef]
  25. Müller, M.; Obuz, H.E.; Keber, S.; Tekmanli, F.; Mettke, L.N.; Yagmurlu, B. Concepts for the Sustainable Hydrometallurgical Processing of End-of-Life Lithium Iron Phosphate (LFP) Batteries. Sustainability 2024, 16, 11267. [Google Scholar] [CrossRef]
  26. Park, J.-K. (Ed.) Principles and Applications of Lithium Secondary Batteries; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  27. Fortune Business Insights Lithium Iron Phosphate Battery Market Size, Share & Industry Analysis, By Type (Portable Battery, Stationary Battery), By Application (Automotive, Industrial, Energy Storage System, Consumer Electronics, and Others), and Regional Forecast, 2024–2032. Available online: https://www.fortunebusinessinsights.com/amp/lithium-ion-li-ion-phosphate-batteries-market-102152 (accessed on 28 February 2025).
  28. Seroka, N.S.; Luo, H.; Khotseng, L. Biochar-Derived Anode Materials for Lithium-Ion Batteries: A Review. Batteries 2024, 10, 144. [Google Scholar] [CrossRef]
  29. Parviz, Z.; Salimi, P.; Javadian, S.; Gharibi, H.; Morsali, A.; Bayat, E.; Leoncino, L.; Lauciello, S.; Proietti Zaccaria, R. Fabrication of Sustainable Hybrid MOF/Silica Electrodes for Current Lithium-Ion Batteries and Beyond. ACS Appl. Energy Mater. 2022, 5, 15155–15165. [Google Scholar] [CrossRef]
  30. Chine, M.K.; Sediri, F.; Gharbi, N. Hydrothermal Synthesis of V3O7.H2O Nanobelts and Study of Their Electrochemical Properties. Mater. Sci. Appl. 2011, 2, 964–970. [Google Scholar] [CrossRef]
  31. Yang, C.; Zhang, J.; Han, S.; Wang, X.; Wang, L.; Yu, W.; Wang, Z. Compositional Controls on Pore-Size Distribution by Nitrogen Adsorption Technique in the Lower Permian Shanxi Shales, Ordos Basin. J. Nat. Gas Sci. Eng. 2016, 34, 1369–1381. [Google Scholar] [CrossRef]
  32. ALOthman, Z.A. A Review: Fundamental Aspects of Silicate Mesoporous Materials. Materials 2012, 5, 2874–2902. [Google Scholar] [CrossRef]
  33. Zhao, S.; Liao, Z.; Fane, A.; Li, J.; Tang, C.; Zheng, C.; Lin, J.; Kong, L. Engineering Antifouling Reverse Osmosis Membranes: A Review. Desalination 2021, 499, 114857. [Google Scholar] [CrossRef]
  34. Akindoyo, J.O.; Beg, M.D.H.; Ghazali, S.; Akindoyo, E.O.; Jeyaratnam, N. Synthesis of Hydroxyapatite through Ultrasound and Calcination Techniques. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2017; Volume 203. [Google Scholar] [CrossRef]
  35. Akindoyo, J.O.; Ghazali, S.; Beg, M.D.H.; Jeyaratnam, N. Characterization and Elemental Quantification of Natural Hydroxyapatite Produced from Cow Bone. Chem. Eng. Technol. 2019, 42, 1805–1815. [Google Scholar] [CrossRef]
  36. Boskey, A.; Pleshko Camacho, N. FT-IR Imaging of Native and Tissue-Engineered Bone and Cartilage. Biomaterials 2007, 28, 2465–2478. [Google Scholar] [CrossRef]
  37. Liu, F.; Wang, R.; Cheng, Y.; Jiang, X.; Zhang, Q.; Zhu, M. Polymer Grafted Hydroxyapatite Whisker as a Filler for Dental Composite Resin with Enhanced Physical and Mechanical Properties. Mater. Sci. Eng. C 2013, 33, 4994–5000. [Google Scholar] [CrossRef] [PubMed]
  38. Zahid, M.; Del Río Castillo, A.E.; Thorat, S.B.; Panda, J.K.; Bonaccorso, F.; Athanassiou, A. Graphene Morphology Effect on the Gas Barrier, Mechanical and Thermal Properties of Thermoplastic Polyurethane. Compos. Sci. Technol. 2020, 200, 108461. [Google Scholar] [CrossRef]
  39. Yang, Y.; Zhang, F.; Wei, K.; Zhai, B.; Wang, X. Porous Carbon Microspheres with Controlled Porosity and Graphitization Degree for High-Performance Supercapacitor. J. Electroanal. Chem. 2022, 918, 116449. [Google Scholar] [CrossRef]
  40. Muraleedharan Pillai, M.; Kalidas, N.; Zhao, X.; Lehto, V.-P. Biomass-Based Silicon and Carbon for Lithium-Ion Battery Anodes. Front. Chem. 2022, 10, 882081. [Google Scholar] [CrossRef]
  41. Wan, Y.; Liu, Y.; Chao, D.; Li, W.; Zhao, D. Recent Advances in Hard Carbon Anodes with High Initial Coulombic Efficiency for Sodium-Ion Batteries. Nano Mater. Sci. 2023, 5, 189–201. [Google Scholar] [CrossRef]
  42. Salimi, P.; Vercruysse, W.; Chauque, S.; Yari, S.; Venezia, E.; Lataf, A.; Ghanemnia, N.; Zafar, M.S.; Safari, M.; Hardy, A.; et al. Lithium-Metal-Free Sulfur Batteries with Biochar and Steam-Activated Biochar-Based Anodes from Spent Common Ivy. Energy Environ. Mater. 2024, 7, e12758. [Google Scholar] [CrossRef]
  43. Deng, J.; Li, M.; Wang, Y. Biomass-Derived Carbon: Synthesis and Applications in Energy Storage and Conversion. Green Chem. 2016, 18, 4824–4854. [Google Scholar] [CrossRef]
  44. Zhao, W.; Zhao, C.; Wu, H.; Li, L.; Zhang, C. Progress, Challenge and Perspective of Graphite-Based Anode Materials for Lithium Batteries: A Review. J. Energy Storage 2024, 81, 110409. [Google Scholar] [CrossRef]
  45. Toki, G.F.I.; Hossain, M.K.; Rehman, W.U.; Manj, R.Z.A.; Wang, L.; Yang, J. Recent Progress and Challenges in Silicon-Based Anode Materials for Lithium-Ion Batteries. Ind. Chem. Mater. 2024, 2, 226–269. [Google Scholar] [CrossRef]
  46. Barbosa Nogueira, M.J.; Chauque, S.; Sperati, V.; Savio, L.; Divitini, G.; Pasquale, L.; Marras, S.; Franchi, P.; Paciornik, S.; Proietti Zaccaria, R.; et al. Untreated Bamboo Biochar as Anode Material for Sustainable Lithium Ion Batteries. Biomass Bioenergy 2025, 193, 107511. [Google Scholar] [CrossRef]
  47. Han, S.-W.; Jung, D.-W.; Jeong, J.-H.; Oh, E.-S. Effect of Pyrolysis Temperature on Carbon Obtained from Green Tea Biomass for Superior Lithium Ion Battery Anodes. Chem. Eng. J. 2014, 254, 597–604. [Google Scholar] [CrossRef]
  48. Christensen, J.; Newman, J. Stress Generation and Fracture in Lithium Insertion Materials. J. Solid State Electrochem. 2006, 10, 293–319. [Google Scholar] [CrossRef]
  49. Feng, J.; Chernova, N.A.; Omenya, F.; Tong, L.; Rastogi, A.C.; Stanley Whittingham, M. Effect of Electrode Charge Balance on the Energy Storage Performance of Hybrid Supercapacitor Cells Based on LiFePO4 as Li-Ion Battery Electrode and Activated Carbon. J. Solid State Electrochem. 2018, 22, 1063–1078. [Google Scholar] [CrossRef]
  50. Xu, J.; Deshpande, R.D.; Pan, J.; Cheng, Y.-T.; Battaglia, V.S. Electrode Side Reactions, Capacity Loss and Mechanical Degradation in Lithium-Ion Batteries. J. Electrochem. Soc. 2015, 162, A2026. [Google Scholar] [CrossRef]
  51. Brutti, S.; Hassoun, J.; Scrosati, B.; Lin, C.-Y.; Wu, H.; Hsieh, H.-W. A High Power Sn–C/C–LiFePO4 Lithium Ion Battery. J. Power Sources 2012, 217, 72–76. [Google Scholar] [CrossRef]
  52. Balogun, M.-S.; Qiu, W.; Luo, Y.; Meng, H.; Mai, W.; Onasanya, A.; Olaniyi, T.K.; Tong, Y. A Review of the Development of Full Cell Lithium-Ion Batteries: The Impact of Nanostructured Anode Materials. Nano Res. 2016, 9, 2823–2851. [Google Scholar] [CrossRef]
  53. Zhang, X.; Aravindan, V.; Kumar, P.S.; Liu, H.; Sundaramurthy, J.; Ramakrishna, S.; Madhavi, S. Synthesis of TiO2 Hollow Nanofibers by Co-Axial Electrospinning and Its Superior Lithium Storage Capability in Full-Cell Assembly with Olivine Phosphate. Nanoscale 2013, 5, 5973–5980. [Google Scholar] [CrossRef]
  54. Zhu, G.-N.; Chen, L.; Wang, Y.-G.; Wang, C.-X.; Che, R.-C.; Xia, Y.-Y. Binary Li4Ti5O12-Li2Ti3O7 Nanocomposite as an Anode Material for Li-Ion Batteries. Adv. Funct. Mater. 2013, 23, 640–647. [Google Scholar] [CrossRef]
  55. Varzi, A.; Bresser, D.; von Zamory, J.; Müller, F.; Passerini, S. Lithium-Ion Batteries: ZnFe2O4-C/LiFePO4-CNT: A Novel High-Power Lithium-Ion Battery with Excellent Cycling Performance (Adv. Energy Mater. 10/2014). Adv. Energy Mater. 2014, 4, 1400054. [Google Scholar] [CrossRef]
  56. Ma, Y.; Younesi, R.; Pan, R.; Liu, C.; Zhu, J.; Wei, B.; Edström, K. Constraining Si Particles within Graphene Foam Monolith: Interfacial Modification for High-Performance Li+ Storage and Flexible Integrated Configuration. Adv. Funct. Mater. 2016, 26, 6797–6806. [Google Scholar] [CrossRef]
  57. Choi, S.; Cho, Y.-G.; Kim, J.; Choi, N.-S.; Song, H.-K.; Wang, G.; Park, S. Mesoporous Germanium Anode Materials for Lithium-Ion Battery with Exceptional Cycling Stability in Wide Temperature Range. Small 2017, 13, 1603045. [Google Scholar] [CrossRef]
  58. Yang, C.; Tartaglino, U.; Persson, B.N.J. Influence of Surface Roughness on Superhydrophobicity. Phys. Rev. Lett. 2006, 97, 116103. [Google Scholar] [CrossRef]
  59. Zafar, M.S.; Gatto, F.; Mancini, G.; Lauciello, S.; Pompa, P.; Athanassiou, A.; Fragouli, D. Biocomposite Cryogels for Photothermal Decontamination of Water. Langmuir 2023, 39, 7793–7803. [Google Scholar] [CrossRef]
  60. Liu, M.; Wang, S.; Wei, Z.; Song, Y.; Jiang, L. Bioinspired Design of a Superoleophobic and Low Adhesive Water/Solid Interface. Adv. Mater. 2009, 21, 665–669. [Google Scholar] [CrossRef]
  61. Liu, X.; Zhou, J.; Xue, Z.; Gao, J.; Meng, J.; Wang, S.; Jiang, L. Clam’s Shell Inspired High-Energy Inorganic Coatings with Underwater Low Adhesive Superoleophobicity. Adv. Mater. 2012, 24, 3401–3405. [Google Scholar] [CrossRef]
Figure 1. The schematic of the complete process for producing biochar (CBW8) from cattle bone waste.
Figure 1. The schematic of the complete process for producing biochar (CBW8) from cattle bone waste.
Sustainability 17 03005 g001
Figure 2. Morphological analyses and pore size distribution of CBW8. (a) Cross-sectional SEM images at different magnifications (from left to right: 100 µm, 50 µm, and 10 µm). (b) BET pore size distribution for CBW8. (c) BET N2 adsorption/desorption isotherms for CBW8.
Figure 2. Morphological analyses and pore size distribution of CBW8. (a) Cross-sectional SEM images at different magnifications (from left to right: 100 µm, 50 µm, and 10 µm). (b) BET pore size distribution for CBW8. (c) BET N2 adsorption/desorption isotherms for CBW8.
Sustainability 17 03005 g002
Figure 3. (a) TGA of CBW8 and XPS analyses. (b) Full spectrum scans of CBW8 showing the high resolution deconvoluted (c) C1s and (d) O1s spectra.
Figure 3. (a) TGA of CBW8 and XPS analyses. (b) Full spectrum scans of CBW8 showing the high resolution deconvoluted (c) C1s and (d) O1s spectra.
Sustainability 17 03005 g003
Figure 4. (a) FTIR, (b) XRD, and (c) Raman spectra of CBW8.
Figure 4. (a) FTIR, (b) XRD, and (c) Raman spectra of CBW8.
Sustainability 17 03005 g004
Figure 5. Electrochemical performance of the CBW8 electrode in lithium half-cells. (a) Galvanostatic charge/discharge profiles for the second cycle, calculated for the four current densities highlighted in the figure. (b) Rate performance at different current densities, and (c) cycling performance at 0.5 A g−1. (d) CV curves at different scan rates ranging from 0.1 to 8 mV s−1. The tests were performed at room temperature in the voltage range 0.01–3 V. The variation of the discharge capacity with cycling is related to the thermal oscillation. Electrolyte: LP30 (1 M LiPF6 in EC: DMC).
Figure 5. Electrochemical performance of the CBW8 electrode in lithium half-cells. (a) Galvanostatic charge/discharge profiles for the second cycle, calculated for the four current densities highlighted in the figure. (b) Rate performance at different current densities, and (c) cycling performance at 0.5 A g−1. (d) CV curves at different scan rates ranging from 0.1 to 8 mV s−1. The tests were performed at room temperature in the voltage range 0.01–3 V. The variation of the discharge capacity with cycling is related to the thermal oscillation. Electrolyte: LP30 (1 M LiPF6 in EC: DMC).
Sustainability 17 03005 g005
Figure 6. Electrochemical results of the prelithiated CBW8/LFP full-cell in the potential range 0.8–3.9 V. (a) The CV profile of the prelithiated CBW8/LFP full-cell at the scan rate of 0.1 mV s−1 for the first three cycles. (b) Long-term cycling test at various C-rates (1 C = 170 mA/gLFP). (c) Galvanostatic charge/discharge profile of the cell at different C-rates, and (d) gravimetrical energy density according to the mass of the active material in the cathode, and to the overall mass of cathode and anode (active material). The analyses were performed at room temperature. Electrolyte: LP30. LFP = 2 mg (delectrode = 11 mm), CBW8 = 1.3 mg (delectrode = 12 mm).
Figure 6. Electrochemical results of the prelithiated CBW8/LFP full-cell in the potential range 0.8–3.9 V. (a) The CV profile of the prelithiated CBW8/LFP full-cell at the scan rate of 0.1 mV s−1 for the first three cycles. (b) Long-term cycling test at various C-rates (1 C = 170 mA/gLFP). (c) Galvanostatic charge/discharge profile of the cell at different C-rates, and (d) gravimetrical energy density according to the mass of the active material in the cathode, and to the overall mass of cathode and anode (active material). The analyses were performed at room temperature. Electrolyte: LP30. LFP = 2 mg (delectrode = 11 mm), CBW8 = 1.3 mg (delectrode = 12 mm).
Sustainability 17 03005 g006
Table 1. Comparison of the main electrochemical properties of LFP-based LIBs.
Table 1. Comparison of the main electrochemical properties of LFP-based LIBs.
AnodeMass Ratio (Anode/Cathode)ICE%Working Voltage (V)Initial Discharge Capacity (mAh g−1)Specific Capacity
(mAh g−1)
Ref.
Sn-C1:2-2.8150120 at 3 C[51]
Graphene1:4893165160 at 1 C[52]
TiO2 hollow nanofibers1:1.3681.4140110 at 0.1 A g−1, 80 at 0.2 A g−1, 50 at 0.5 A g−1, and 30 at 1 A g−1[53]
Binary Li4 Ti5O12-Li2 Ti3O7--1.812575 at 0.08 A g−1[54]
ZnFe2O4-C1:1.5-2.112080 at 9.6 C[55]
Si-Graphene-83.23157130 at 1 C[56]
Ge-932.810590 at 0.5 C[57]
CBW81:1.5803167165 at 0.5 C, 155 at 1 C, 142 at 2 C, 95 at 5 C, 65 at 10 CThis work
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

Zafar, M.S.; Salimi, P.; Ricci, M.; Zia, J.; Zaccaria, R.P. Environmentally Sustainable Anode Material for Lithium-Ion Batteries Derived from Cattle Bone Waste: A Full-Cell Analysis with a LiFePO4 Cathode. Sustainability 2025, 17, 3005. https://doi.org/10.3390/su17073005

AMA Style

Zafar MS, Salimi P, Ricci M, Zia J, Zaccaria RP. Environmentally Sustainable Anode Material for Lithium-Ion Batteries Derived from Cattle Bone Waste: A Full-Cell Analysis with a LiFePO4 Cathode. Sustainability. 2025; 17(7):3005. https://doi.org/10.3390/su17073005

Chicago/Turabian Style

Zafar, Muhammad Shajih, Pejman Salimi, Marco Ricci, Jasim Zia, and Remo Proietti Zaccaria. 2025. "Environmentally Sustainable Anode Material for Lithium-Ion Batteries Derived from Cattle Bone Waste: A Full-Cell Analysis with a LiFePO4 Cathode" Sustainability 17, no. 7: 3005. https://doi.org/10.3390/su17073005

APA Style

Zafar, M. S., Salimi, P., Ricci, M., Zia, J., & Zaccaria, R. P. (2025). Environmentally Sustainable Anode Material for Lithium-Ion Batteries Derived from Cattle Bone Waste: A Full-Cell Analysis with a LiFePO4 Cathode. Sustainability, 17(7), 3005. https://doi.org/10.3390/su17073005

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