Masfar Alkhatani*a,
Mzoun M. Almutairib,
Mariam H. Alalwanb,
Hawraa K. Algharashb,
Raghad A. Aldossaryb,
Yahya A. Alzahrania,
Sultan Alenzia,
Ibtisam S. Almalkia,
Ghazal S. Yafic,
Abdulmalik M. Alessaa,
Faisal S. Alghannama,
Abdulaziz Aljuwayra,
Nouf K. Al-Saleem*b,
Anwar Alanazia and
Masaud Almalkia
aFuture Energy Technologies Institute, King Abdulaziz City for Science and Technology (KACST), P.O. Box 6086, Riyadh 11442, Saudi Arabia. E-mail: mqhtani@kacst.gov.sa
bDepartment of Physics, College of Science and Humanities, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Jubail, Saudi Arabia. E-mail: nkalsaleem@iau.edu.sa
cDepartment of Chemistry, King Saud University, P.O. Box 2455, Riyadh, 11451, Saudi Arabia
First published on 23rd June 2025
We present a cost-effective approach for converting petroleum coke into high-quality graphene via thermal desulfurization followed by electrochemical exfoliation. The synthesized graphene exhibited structural integrity and optoelectronic properties comparable to those of graphite-derived materials. Integrated as rear electrodes in perovskite solar cells (PSCs), the petcoke-derived graphene enabled devices to achieve a power conversion efficiency (PCE) of approximately 19.1%, with an open-circuit voltage (Voc) exceeding 1.13 V and a fill factor (FF) of ∼78%. The petcoke-derived graphene PSCs exhibited superior photovoltaic performance compared to those fabricated with commercial graphene, and demonstrated comparable efficiency to devices employing conventional gold electrodes. These findings highlight the potential of waste-derived graphene as a scalable, sustainable, and economically viable alternative to noble metal contacts, aligning with circular economy principles and advancing low-cost photovoltaic technologies.
To address these limitations, perovskite solar cells (PSCs) have emerged as one of the most transformative technologies in next-generation photovoltaics.14–16 Since their first report in 2009 with a modest 3.8% PCE, PSCs have witnessed an unprecedented rise in performance, reaching certified efficiencies exceeding 26.7% as of 2025.17 Their remarkable optoelectronic properties—including strong light absorption, long carrier diffusion lengths, tunable bandgaps, and solution-processability at low temperatures position PSCs as strong candidates for low-cost, high-efficiency, and flexible energy harvesting devices.14–16,18
Nonetheless, several challenges impede the commercial realization of PSCs. Key issues include stability under environmental stressors (e.g., humidity, oxygen, light, and heat),14,15 scalability of large-area fabrication, and high material costs, particularly those arising from noble metal electrodes such as gold (Au) and silver (Ag).19 In this context, there is a growing interest in replacing conventional metal contacts with cost-effective, earth-abundant alternatives to improve both the economic and environmental profiles of PSCs.20,21
Carbon-based materials, especially graphene, have attracted considerable attention as promising substitutes due to their outstanding electrical conductivity, chemical stability, mechanical flexibility, and optical transparency.20–24 Graphene has been integrated into various parts of PSC architectures, including as transparent electrodes, hole/electron transport layers, and interfacial modifiers, with demonstrated improvements in both performance and device longevity.20 Particularly, its use as a counter electrode has shown potential for cost reduction and enhanced moisture stability. However, most reported studies rely on graphene produced via chemical vapor deposition (CVD) or reduction of graphene oxide (GO) processes that are often expensive, multistep, and environmentally taxing, thereby limiting large-scale implementation.20
To address these limitations, researchers are actively exploring low-cost and sustainable feedstocks for graphene production. One compelling alternative is petroleum coke (petcoke), a carbon-rich byproduct of oil refining that is typically relegated to high-emission industrial uses such as metallurgical processes and fuel combustion.25–27 Owing to its high fixed-carbon content, low volatile impurities, and global abundance, petcoke represents a largely underutilized waste stream with considerable potential for upcycling into high-value nanomaterials. Recent studies have demonstrated that electrochemical exfoliation of petcoke in aqueous electrolytes can yield conductive few-layer graphene through a simple, potentially scalable, and environmentally benign process, provided that the petcoke undergoes appropriate thermal pretreatment.25–27 This post-thermal activation step plays a critical role in reducing impurities, enhancing graphitic ordering, and improving exfoliation efficiency.
In this study, we adapted the post-thermal activation step and reported an efficient approach to synthesize few-layer graphene with a much better yield from locally sourced petroleum coke using electrochemical exfoliation in sulfate-based electrolytes. The electrochemical technique itself promotes ion intercalation and subsequent exfoliation under mild conditions, offering a viable and sustainable pathway for large-scale graphene synthesis using a low-value carbon feedstock. The produced graphene was systematically characterized using structural, spectroscopic, and electronic techniques to assess its suitability as an electrode material. To evaluate its practical applicability, we integrated the petcoke-derived graphene as a rear electrode in planar perovskite solar cells, replacing conventional gold contacts. The photovoltaic performance of these devices was then systematically compared to reference devices fabricated with commercially available graphene and noble metal electrodes. Notably, the PSCs employing petcoke-derived graphene exhibited better performance compared to those using commercial graphene, and demonstrated comparable efficiency and stability to devices with traditional gold electrodes. These findings underscore the potential of petcoke upcycling as a cost-effective and scalable route for producing high-quality graphene for next-generation photovoltaic technologies.
A compact TiO2 layer was deposited via spray pyrolysis from a precursor solution containing titanium diisopropoxide bis(acetylacetonate), acetylacetone, and ethanol. The substrates were maintained at 450 °C during and after deposition to ensure proper adhesion and crystallization. A mesoporous TiO2 layer was subsequently applied using a commercial TiO2 paste (Dyesol 18NR-T) via spin-coating at 4000 rpm for 30 seconds, followed by annealing at 450 °C for 30 minutes.
The perovskite active layer was prepared from a precursor solution comprising 1.6 M PbI2, 1.51 M FAI, 0.04 M PbBr2, 0.33 M MACl, and 0.04 M MABr dissolved in DMF:
DMSO (8
:
1 v/v). The solution was deposited using a two-step spin-coating process (2000 rpm for 10 s, then 6000 rpm for 30 s), with chlorobenzene applied as an anti-solvent during the final 18 seconds. Films were annealed at 100 °C for 10 minutes and then at 150 °C for an additional 10 minutes. After that, the hole transport layer (HTL) was fabricated by spin-coating a solution of 102.72 mg spiro-OMeTAD in chlorobenzene, doped with Li-TFSI and 4-tert-butylpyridine (tBP), at 4000 rpm for 30 seconds.
The transfer was conducted via a press-transfer method, in which the plasma-treated graphene film was manually aligned over the device surface and laminated using a soft polydimethylsiloxane (PDMS) stamp under gentle pressure (∼2–5 N cm−2) at room temperature for approximately 1 minute. After transfer, the assembled device was annealed at 80 °C for 10 minutes to further enhance bonding between the graphene electrode and the hole transport layer. While freestanding graphene electrodes are often associated with increased series resistance and reduced fill factor (FF) due to suboptimal interfacial contact, the combined surface activation and mechanical lamination techniques employed here mitigated these limitations. This approach ensured uniform contact, minimized interfacial voids or wrinkles, and enabled efficient charge extraction, contributing to the relatively high FF values obtained in our devices.
To enhance the crystalline structure of the petroleum coke used in this study, a high-temperature pre-treatment was conducted at 1400 °C for 10 hours under an inert argon atmosphere. This specific thermal condition was selected based on its dual effect: it effectively facilitates the desulfurization of high-sulfur petroleum coke and simultaneously promotes graphitization by aligning aromatic carbon domains into ordered graphitic layers.29,30 X-ray diffraction (XRD) analysis was performed on the coke samples before and after the thermal treatment. As shown in Fig. 1(b), the annealed sample exhibited a marked increase in the intensity and sharpness of the (002) diffraction peak, indicative of improved graphitic ordering. Quantitative analysis revealed that the degree of crystallinity increased to approximately 72%, confirming the transformation of the disordered carbon domains into a more ordered, graphite-like structure.31 This enhancement in crystallinity significantly improves the material's suitability for electrochemical exfoliation and contributes to the production of high-quality graphene nanosheets.
Detailed morphological and compositional characterization of graphene-like structures obtained via electrochemical exfoliation of thermally treated petroleum coke were carried out using a transmitted electron microscope (TEM). Fig. 2(a) and (b) display low magnification images of transparent, ultrathin, and crumpled nanosheets morphological traits typically associated with few-layer graphene. Furthermore, high-resolution TEM in Fig. 2(c) unveils a well-defined, hexagonally arranged atomic lattice, characteristic of graphitic carbon. This lattice structure closely resembles that observed in electrochemically exfoliated graphene derived from natural graphite, suggesting that the exfoliation process preserves the intrinsic crystallinity of the precursor material. The selected area electron diffraction (SAED) pattern shown in the inset of Fig. 2(c) exhibits sharp, symmetric, and hexagonally arranged diffraction spots, unequivocally confirming the presence of a crystalline sp2-bonded carbon framework.27,29 The clarity and symmetry of the SAED pattern further attest to the high degree of structural order and the retention of graphitic domains after exfoliation. Collectively, these observations confirm the successful formation of few-layer graphene with preserved crystallinity from petroleum coke via electrochemical exfoliation.
Fig. 2(d) shows the corresponding energy-dispersive X-ray spectroscopy (EDX) spectrum, providing elemental insights into the composition of the exfoliated sheets. The dominant peak at ∼0.28 keV corresponds to carbon (C), affirming the carbon-rich nature of the product. A secondary peak appears at ∼0.53 keV, corresponding to oxygen (O). This signal likely arises from surface oxygen-containing functional groups such as hydroxyl (–OH), carboxyl (–COOH), and carbonyl (CO) moieties introduced during electrochemical exfoliation in aqueous or mildly oxidative conditions. These oxygen functionalities can also contribute to improved dispersibility and potential redox activity in energy storage applications. A weak yet distinguishable sulfur (S) peak at approximately 2.3 keV suggests the presence of residual sulfur species. While the intensity of this peak is low indicating that the high-temperature (1400 °C) thermal treatment was effective in removing a significant portion of sulfur the presence of trace sulfur implies incomplete desulfurization, which is common in high-sulfur petroleum coke precursors. Further optimization of the pretreatment conditions could potentially eliminate these residues. In addition to those peaks, strong copper (Cu) peaks near 8.0 keV and 9.0 keV are attributed to the copper TEM grid used during imaging. These peaks are instrumental artifacts and are not representative of the sample's intrinsic composition.
To further investigate the structural evolution during exfoliation, Raman spectroscopy was employed to analyze both the thermally treated petroleum coke and the resulting graphene,28,32,33 as shown in Fig. 3. Raman spectroscopy is a powerful and widely used technique for characterizing carbon-based materials, particularly in distinguishing between graphitic domains, structural disorder, and the number of graphene layers. The characteristic Raman bands observed include the D band (∼1350 cm−1), associated with defect-induced breathing modes of sp2 carbon rings; the G band (∼1580 cm−1), corresponding to the in-plane vibration (E2g mode) of sp2-hybridized carbon atoms; and the 2D band (∼2700 cm−1), which serves as a fingerprint for the number of graphene layers and the stacking order.32 Raman spectra were collected using a 532 nm excitation laser on both samples drop-cast onto glass substrates. The spectrum of the annealed petroleum coke (red curve) displays a prominent D band, a broadened G band, and a weak, broad 2D band. This indicates the presence of partially graphitized domains with a significant degree of disorder, likely due to defects, edge states, and residual functional groups such as hydroxyl and epoxide moieties.
In contrast, the Raman spectrum of the exfoliated graphene (black curve) reveals a significantly sharper and more intense 2D band centered at ∼2700 cm−1, indicative of few-layer graphene. The I2D/IG intensity ratio ranges from 0.5 to 0.7, suggesting the presence of two to five graphene layers with relatively low stacking disorder. Furthermore, the ID/IG ratio, which varies between 0.8 and 0.9 across different regions of the sample, reflects the presence of some residual defects, yet within a range acceptable for conductive applications.32,33 These findings confirm the successful exfoliation of graphene sheets from petroleum coke and the partial restoration of sp2 graphitic domains. To evaluate the scalability and batch-to-batch reproducibility of our electrochemically exfoliated graphene, we developed a standardized protocol with tight control over key parameters, including electrolyte composition, applied voltage, exfoliation duration, and post-annealing conditions. These parameters were optimized to ensure consistency across production runs. To confirm reproducibility, three independent exfoliation batches were performed under identical conditions. The resulting graphene films demonstrated consistent electrical performance, with sheet resistance values varying by less than ±7%, as determined using a four-point probe method. Morphological analyses using TEM revealed a uniform flake thickness distribution, with the majority of flakes exhibiting 4–6 layers.
Prior to integrating coke-derived graphene into perovskite solar cell architectures, it was essential to evaluate its electrical properties to ensure its suitability as a conductive electrode material. To this end, freestanding films were fabricated via vacuum filtration from dispersions of the exfoliated graphene sheets. The baseline electrical conductivity of the thermally treated but non-exfoliated petroleum coke was measured at 65 S m−1, reflecting limited structural order and high defect density. Following electrochemical exfoliation, a notable enhancement in conductivity was observed, with the coke-derived graphene films reaching a conductivity of 125 S m−1, more than doubling the initial value. This improvement is attributed to the increased restoration of sp2 carbon domains and the partial removal of insulating oxygenated functional groups during the exfoliation process. To further optimize the electrical performance, post-synthesis thermal annealing was employed at elevated temperatures. The exfoliated films were annealed under inert atmosphere at 500 °C and 1100 °C for 12 hours. After annealing at 500 °C, the conductivity increased to 225 S m−1, and further rose to 310 S m−1 upon annealing at 1100 °C. The achieved value is in the range of conductivity needed for Li ion batteries.34,35 This progressive enhancement is primarily attributed to the thermal removal of residual oxygen-containing functional groups (e.g., hydroxyl, carboxyl, and epoxide) as shown in the FTIR spectra illustrated in Fig. 1S,† as well as adsorbed impurities that act as scattering centers. The additional increase observed beyond 500 °C may result from molecular rearrangement and further graphitization, leading to improved π-conjugation and reduced defect density in the carbon lattice.27,36
Next, we fabricated two PSC devices with and without coke derived graphene electrode and their photovoltaics was performed and compared. The general and common structure of PSCs, as illustrated in Fig. 4(a), consists of multiple functional layers: a cathode at the top, typically made of conductive materials like gold or silver, which collects electrons; a hole transport layer (HTL) that facilitates the movement of positive charge carriers (holes) towards the cathode; an absorber layer made of perovskite material, such as methylammonium lead iodide (CH3NH3PbI3), which is responsible for light absorption and generating electron–hole pairs; an electron transport layer (ETL) that transports electrons to the anode; and a glass substrate coated with a transparent conductive oxide like ITO or FTO, serving as the anode.37 The working principle involves sunlight absorption by the perovskite layer, generating excitons that separate into free electrons and holes. Electrons are transported through the ETL to the anode while holes move through the HTL to the cathode, creating an electric current.38,39
To evaluate the structural and optical properties of the synthesized perovskite layer prior to full device fabrication, the perovskite films were deposited on fluorine-doped tin oxide (FTO) substrates sequentially coated with compact TiO2 (c-TiO2) and mesoporous TiO2 (m-TiO2). The detailed fabrication process is outlined in the Materials and methods section. Structural and morphological analyses were carried out using scanning electron microscopy (SEM), X-ray diffraction (XRD), UV-Vis absorption spectroscopy, and photoluminescence (PL) measurements, as presented in Fig. 4. The SEM images, shown in Fig. 4(b), reveal a densely packed perovskite film with well-defined grains and significantly enhanced grain size. Larger grains are known to reduce grain boundary density, thereby minimizing non-radiative recombination pathways and enhancing charge carrier transport factors that directly contribute to improved photovoltaic performance.24,38
XRD analysis, presented in Fig. 4(c), provides insights into the crystallographic structure of the perovskite layer. The diffraction peaks observed at 13.90°, 19.73°, 24.16°, 27.85°, 31.17°, 34.31°, 39.86°, and 42.29° correspond to the (001), (011), (111), (002), (012), (112), (022), and (122) crystal planes of the α-phase formamidinium lead iodide (α-FAPbI3), respectively.24 Notably, the prominent diffraction peak at 13.90°, corresponding to the (001) orientation, is present in both control and PEAI-modified perovskite films, indicating the dominant presence of the photoactive black phase. Optical characterization of the film was conducted using UV-Vis absorption spectroscopy, as shown in Fig. 4(d). The absorption spectrum exhibits a sharp absorption edge at approximately 815 nm, which corresponds to an optical band gap of 1.55 eV, consistent with values reported for high-quality α-FAPbI3 perovskite films. Complementary photoluminescence (PL) spectroscopy also revealed a strong and distinct emission peak, closely aligned with the absorption edge, further confirming the formation of a phase-pure perovskite structure with minimal trap-assisted recombination.
To investigate the feasibility of petroleum coke-derived graphene as a cost-effective alternative to noble metal electrodes, perovskite solar cells (PSCs) were fabricated using graphene as the rear electrode and benchmarked against control devices employing thermally evaporated gold. Both types of devices shared an identical n–i–p configuration: FTO/c-TiO2/m-TiO2/perovskite/PEAI/spiro-OMeTAD/electrode. Photovoltaic characterization revealed that perovskite solar cells (PSCs) employing petcoke-derived, electrochemically exfoliated graphene as the rear electrode exhibited a power conversion efficiency (PCE) of approximately 19.1%. The devices achieved an open-circuit voltage (Voc) exceeding 1.12 V, a fill factor (FF) approaching 75%, and a short-circuit current density (Jsc) of ∼23.1 mA cm−2, as shown in Fig. 5(a) and (b) and summarize in Table 1. These performance metrics suggest efficient charge extraction and low interfacial recombination losses at the graphene–perovskite interface. The enhancement in Voc is attributed to favorable energy level alignment, where literature-reported work function values for electrochemically exfoliated graphene (typically ∼4.6–4.9 eV)40,41 are positioned between the valence band maximum of MAPbI3 (∼5.4 eV) and the HOMO level of spiro-OMeTAD (∼5.1 eV). This alignment is expected to promote effective hole extraction and minimize interfacial energy barriers.
PSC device structure | Jsc (mA cm−2) | Voc (V) | FF (%) | PCE (%) |
---|---|---|---|---|
FTO/TiO2/perovskite/spiro-OMeTAD/petcoke-derived graphene | 23.1 | 1.12 | 75 | 19.1 |
FTO/TiO2/perovskite/spiro-OMeTAD/commercial graphene | 22.5 | 1.1 | 73 | 18 |
FTO/TiO2/perovskite/spiro-OMeTAD/Au | 25.3 | 1.13 | 76 | 22 |
Consequently, the use of graphene contributes to the observed improvements in both Voc and FF, which are typically associated with suppressed interfacial recombination and enhanced carrier transport. Although direct characterization techniques such as time-resolved photoluminescence (TRPL) and electrochemical impedance spectroscopy (EIS) are routinely used to probe carrier dynamics and quantify non-radiative losses, these measurements were not conducted in the present study due to instrumental constraints. Nevertheless, the elevated Voc and the absence of pronounced hysteresis in the current–voltage (J–V) curves provide indirect evidence of reduced non-radiative recombination at the graphene–perovskite interface. Future work will incorporate TRPL and EIS analyses to enable a more comprehensive understanding of the interfacial charge dynamics and to further validate the role of petcoke-derived graphene in defect passivation and charge transfer enhancement.
We evaluated the long-term operational stability of the fabricated PSC device employing exfoliated graphene electrodes. The devices were subjected to continuous illumination under 1 sun equivalent intensity using an LED solar simulator in an inert nitrogen (N2) atmosphere. As shown in Fig. 5(c), the normalized power conversion efficiency (PCE) remained above 90% of the initial value for over 900 hours, demonstrating excellent operational durability. These results highlight the intrinsic stability of the graphene-based PSCs under prolonged operating conditions.
To contextualize this performance, we compared petcoke-derived graphene-based PSC device to similar PSC architectures using commercial graphene and gold (Au) electrodes. As summarized in Table 1, the Au-based device delivered the highest PCE of 22.0%, attributed to its higher Jsc ∼ 25.3 mA cm−2, while Voc and FF remained similar (∼1.13 V and 76%) see Fig. 2S.† The device employing commercial graphene exhibited a lower PCE of 18.0%, confirming that our exfoliated, petcoke-derived graphene offers superior performance among carbon-based alternatives. Notably, while the Au-based control device exhibits the highest PCE and Jsc, our graphene-based PSC stands out by maintaining high Voc and FF while replacing a costly, diffusion-prone noble metal with a sustainable, low-cost carbon electrode. The slightly lower Jsc in graphene-based devices may stem from differences in electrical conductivity or interfacial contact quality. Further optimization, such as doping, conductivity enhancement, or improved interfacial engineering may help close this performance gap. This comparison underscores the strong potential of exfoliated graphene derived from petroleum coke as a scalable and environmentally sustainable electrode material for next-generation perovskite solar cells.
To assess the performance of our fabricated perovskite solar cell (PSC) employing petcoke-derived exfoliated graphene as the rear electrode, we benchmarked its photovoltaic characteristics against previously reported graphene-based PSCs21–24 that incorporate graphene or graphene-derivative rear contacts. Table 2 summarizes key photovoltaic parameters, including short-circuit current density (Jsc), open-circuit voltage (Voc), fill factor (FF), and power conversion efficiency (PCE), for representative device architectures.
PSC device structure | Jsc (mA cm−2) | Voc (V) | FF (%) | PCE (%) | Ref. |
---|---|---|---|---|---|
Glass/FTO/TiO2/perovskite/spiro-OMeTAD/graphene | 19.17 | 0.96 | 67.22 | 12.37 | 21 |
Graphene-based flexible perovskite solar cells | 21 | 0.99 | 72 | 15 | 22 |
FTO/SnO2/perovskite/spiro-OMeTAD/graphene layer | 22.69 | 1.05 | 77 | 18.25 | 23 |
FTO/TiO2/perovskite/spiro-OMeTAD/graphene oxide | 25.06 | 0.984 | 69 | 17 | 24 |
FTO/TiO2/perovskite/spiro-OMeTAD/petcoke-derived graphene | 23.1 | 1.12 | 75 | 19.1 | This work |
As seen in the Table 2, our device demonstrates a high open-circuit voltage of 1.12 V and a fill factor of 75%, contributing to a competitive PCE of 19.1%. This performance surpasses many of the previously reported graphene-based PSCs, particularly those using untreated graphene or graphene oxide, which typically suffer from lower FF and Voc due to interfacial energy mismatches or poor film conductivity. While some devices employing optimized SnO2 scaffolds exhibit slightly higher FF, the Voc achieved in our work is among the highest reported for graphene-based electrodes, underscoring the favorable energy alignment between the petcoke-derived graphene and the hole transport layer. The enhanced device performance can be attributed to both the improved conductivity of the thermally annealed graphene and the effective interface formed between the exfoliated graphene and the perovskite/spiro-OMeTAD layers. This comparative analysis highlights the viability of petroleum coke as a low-cost and scalable precursor for high-performance graphene electrodes in PSCs, providing an eco-friendly alternative to conventional gold contacts without compromising efficiency.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5na00441a |
This journal is © The Royal Society of Chemistry 2025 |