Open Access Article
Abdelaziz Gouda*abc,
Juan Manuel Restrepo-Flórez
d,
Abhinav Mohanabc,
Otavio Augusto Titton Dias
c,
Andrew Wanga,
Camilo J. Viasus Perez
ae,
Jessica Ye
ac,
Tamlyn Slocombe
a,
Vijay K. Tomer
c,
Jiabao Shenf,
Nhat Truong Nguyen
g,
Nazir Kheranif,
Mohamad Hmadehah,
Alán Aspuru-Guzik
abi,
Mohini Sain*bc and
Geoffrey Ozin
*ab
aDepartment of Chemistry, University of Toronto, 80 Saint George St., Toronto, Ontario M5S 3H6, Canada. E-mail: Abdelaziz.gouda@utoronto.ca; g.ozin@utoronto.ca
bDepartment of Applied Chemistry and Chemical Engineering, University of Toronto, 200 College St., Toronto, ON M5S 3E5, Canada. E-mail: m.Sain@utoronto.ca
cDepartment of Mechanical and Industrial Engineering, University of Toronto, 5 King's College Rd, Toronto, ON M5S 3G8, Canada
dDepartment of Chemical Engineering, University of Florida, Gainesville, Florida 32611, USA
eGrupo de Investigación en Química de Coordinación y Bioinorgánica, Departamento de Química, Facultad de Ciencias, Universidad Nacional de Colombia, Av. Carrera 30 # 45-03, Sede Bogotá, Colombia
fDepartment of Electrical and Computing Engineering, University of Toronto, 10 King's College Road, Toronto, Ontario M5S 3G4, Canada
gDepartment of Chemical and Material Engineering, Gina Cody School of Engineering and Computer Science, Concordia University, Montreal, Quebec, Canada
hDepartment of Chemistry, Faculty of Arts and Sciences, American University of Beirut, P.O. Box 11-0236, Beirut 1107 2020, Lebanon
iDepartment of Computer Science, University of Toronto, Sandford Fleming Building, 10 King's College Road, Toronto, Ontario M5S 3G4, Canada
First published on 13th November 2025
The transition to sustainable hydrogen production is critical to decarbonizing the global energy system and reducing reliance on carbon-intensive methods such as steam methane reforming (SMR). Methane pyrolysis has emerged as a lower-emission alternative, yet its viability is constrained by the extreme temperatures required (600–1200 °C) and associated energy demands. Here, we present methane photolysis, CH4(g) + hv → C(s) + 2H2(g), as a novel, light-driven pathway for simultaneous hydrogen and carbon nanotube production. Operating at ambient conditions and powered by LED illumination, this scalable and continuous process selectively yields clean hydrogen with a maximum hydrogen production rate of 17.74 ± 1.71 mol cm−2 h−1 and high-value carbon nanotubes (CNTs) achieving yields of approximately 8.0 g CNT per g catalyst at a production rate of ∼0.150 ± 0.001 g h−1, while circumventing the thermal and carbon burdens of conventional methods. Compared to water electrolysis, our approach requires approximately 70% less energy per kilogram of H2 produced. For CNTs, the energy savings reach ∼74% relative to conventional methods such as fluidized-bed chemical vapor deposition. Techno-economic and life-cycle assessments demonstrate its potential as a scalable, energy-efficient alternative for decentralized hydrogen and carbon nanomaterial synthesis, with implications for cleaner fuel production and circular carbon utilization.
Broader contextThe global shift toward sustainable hydrogen production is essential for decarbonizing energy systems and achieving climate goals. While current industrial methods like steam methane reforming and high-temperature pyrolysis can produce hydrogen at scale, they are energy-intensive and carbon-emitting. This study introduces a light-driven methane photolysis process that operates at ambient conditions to simultaneously generate clean hydrogen and valuable multi-walled carbon nanotubes (MWCNTs). Unlike conventional approaches, this method eliminates the need for high temperatures, reducing energy input by up to 70% for hydrogen and 74% for CNTs. Techno-economic and life cycle analyses show that the light-driven process not only lowers the environmental footprint but also enhances economic feasibility, particularly under solar or LED operation. Beyond hydrogen, the co-produced MWCNTs offer sustainable value as next-generation materials in batteries and electrical systems, potentially displacing copper. By integrating system energy requirements, carbon valorization, and economic viability, this approach represents a transformative advancement in clean hydrogen and carbon nanomaterial production. |
Steam-methane reforming
![]() | (1) |
One such promising alternative is methane pyrolysis, which can produce hydrogen without generating CO2 by decomposing methane (CH4) into hydrogen (H2) and solid carbon.6–8
Methane pyrolysis
![]() | (2) |
Unlike SMR, methane pyrolysis avoids direct CO2 emissions by utilizing high temperatures (600–1200 °C) to induce the necessary chemical reactions. While this process offers a cleaner alternative to SMR, the intensive thermal requirements result in substantial energy inputs, limiting its economic viability and scalability, especially when considering the broader goal of decarbonizing the energy sector. Consequently, there is an increasing demand for methods that can overcome the energy-intensive nature of traditional pyrolysis while still enabling efficient, low-carbon hydrogen production.
Another promising thermochemical approach employs concentrated solar to crack methane around 1600 K in a particle flow reactor using light intensities equivalent to 1720 suns to produce hydrogen (H2) and carbon filaments.9 Solar thermal cracking of methane of this kind has been explored in different reactor configurations, such as rotary beds and roll-to-roll reactors.10–12
Methane photolysis, described herein by contrast, is a photocatalytic (photochemical/photothermal) process in which light under ambient conditions is used to dissociate methane into hydrogen and carbon nanotubes. The direct use of photons eliminates the need for high temperatures, significantly reducing the energy input required compared to the aforementioned methods. The carbon nanotube coproduct brings dual value to the process with applications in energy storage, electronics, and advanced materials science.13–15 The process can be scaled and operated continuously with light-emitting diodes, thereby circumventing solar intermittency. It outperforms methane steam reforming and water electrolysis in energy requirements, environmental impact, and cost-effectiveness, as validated by system energy requirements calculations, techno-economic and life-cycle analyses (TEA–LCA). A comparative summary of similar work in the literature, including methane pyrolysis and concentrated solar-driven cracking approaches, is presented in Table S1. To the best of our knowledge, this is the first report of methane photolysis driven by low-intensity light irradiation that enables the simultaneous production of high-value CNTs and clean hydrogen, with CNTs mechanically separable from the catalyst, eliminating the need for chemical post-processing.
Control tests were performed to monitor the activity of each component of the reaction system including the empty reactor, catalyst, and photo-grown MWCNT, to confirm catalyst optimization and rule out external contributions (Fig. 1b). 10% Ni/Al2O3 showed an optimum H2 rate compared to pure Ni particles and pure MWCNT, which could be related to the metal/support interactions that improved the thermal and chemical stability of the catalyst.22 Nickel was selected for this study as it is widely reported in the literature as the most active monometallic catalyst for methane cracking, exhibiting high methane conversion and hydrogen yield across a wide range of conditions.7,23–25 Its strong activity in C–H bond activation, thermal stability, and compatibility with Al2O3 support made it a suitable choice for our system. That said, we have also experimentally tested Co, Fe, and Cu catalysts supported on Al2O3 under similar conditions (Fig. S4). Ni consistently demonstrated the highest catalytic activity and favourable formation of ordered solid carbon structures among these candidates.
Various Ni loadings on Al2O3 were investigated to study the effect on the H2 rate and CNT formation (Fig. S5a). 10% Ni loading yielded MWCNT and had the highest H2 production rate compared to 1%, 5% and 20% Ni; all catalysts have been tested under the same experimental conditions. PXRD patterns of the spent 10% and 20% Ni/Al2O3 showed the presence of the graphitic carbon (002) at 26.2°, while 1% and 5% did not (Fig. S5b). These results confirm the importance of the Ni particle size and distribution on photocatalytic H2 and MWCNT formation. The influence of Ni particle size on conventional thermal methane cracking has been extensively documented in the literature.26–30
For this study, Ni/γ-Al2O3 was deliberately chosen due to its well-established strong metal–support interaction, which is widely reported to promote high dispersion of Ni particles, suppress sintering, and provide consistent methane decomposition activity and stability. Strong metal–support interaction with γ-Al2O3 also enables the formation of smaller Ni nanoparticles, which are more resistant to aggregation and deactivation.31–33
Ni/γ-Al2O3 additionally favors the formation of ordered carbon structures by supporting base-growth of carbon nanotubes (CNTs) for the strong metal–support interaction with better thermal stability for the gamma alumina support, facilitating separation and enhancing catalyst durability.33 This is supported by studies showing that compared to inert (SiO2) or basic (MgO, CaO) supports, γ-Al2O3 enables higher CNT quality and stability, while basic supports may favor amorphous carbon and inert supports tend to result in larger Ni particles and tip-grown CNTs with faster deactivation. Carbon-based supports can complicate mechanistic interpretation due to simultaneous carbon formation from the support itself.34
In addition, the acidity/basicity and high surface area of γ-Al2O3 facilitate the balance between catalytic activity and carbon structure, as also demonstrated in recent catalyst design studies.
We have performed direct experimental comparisons under identical conditions for 10% Ni/γ-Al2O3, 10% Ni/SiO2, and 10% Ni/MgO. The results show that both Ni/γ-Al2O3 and Ni/SiO2 give high hydrogen production rates, with Ni/SiO2 the most active and Ni/MgO showing the lowest activity (Fig. S6). TEM imaging confirms that Ni/γ-Al2O3 supports base-grown CNTs, which allow easier CNT separation, whereas Ni/SiO2 leads primarily to tip-growth CNTs due to weaker metal–support interactions (Fig. S7). 10% Ni/SiO2 catalysts exhibit notably higher activity for methane photolysis than Ni/γ-Al2O3, under typical conditions. This increased activity is attributed to weaker metal–support interactions on SiO2, which expose more active nickel sites for reaction.35 However, this same weak interaction leads to faster catalyst deactivation due to rapid carbon accumulation that blocks the nickel surfaces. In contrast, Ni/γ-Al2O3 demonstrates greater long-term stability and enhanced resistance to coking, owing to stronger metal–support interactions.36 Additionally, the weak interactions in Ni/SiO2 favor carbon nanotube (CNT) formation through tip-growth mode, which complicates product separation and diminishes CNT quality by increasing structural disorder and metal contamination35,37 (Fig. S7b). The lowest activity of 10% Ni/Mgo is related to the strong interaction between nickel and the MgO support leads to the formation of stable NiO–MgO solid solutions or spinel-type structures, which significantly inhibit the reduction of nickel oxide to its active metallic Ni form.38 As a result, fewer active Ni0 sites are available to catalyze methane decomposition.39 Additionally, the basic nature of MgO support can stabilize nickel in its oxidized state and prevent the dispersion and accessibility of metallic Ni particles, further lowering activity. Enhancing Ni/MgO activity often requires the addition of promoters such as chromium or copper, which improve the reduction of Ni and increase the number of accessible active sites.40 In the absence of such promoters, Ni/MgO remains much less active than Ni/SiO2 or Ni/Al2O3 for methane photolysis.38–40
The dependence of reaction rate on light intensity is a key indicator of the dominant reaction mechanism. A linear relationship typically suggests a photochemical process, where photon absorption directly generates reactive species (e.g., excited electrons, radicals) that drive the reaction. In such cases, the generation rate of these species and therefore the overall reaction rate is directly proportional to the photon flux (i.e., light intensity, I), and can be expressed as:
The dependence of reaction rate on light intensity
| Rate ∝ I | (3) |
Each absorbed photon can initiate an elementary reaction, so doubling the light intensity approximately doubles the reaction rate, resulting in a straightforward linear correlation.41,42
In contrast, an exponential or super-linear dependence indicates a photothermal mechanism, where light energy is converted into localized heat, which accelerates the reaction through thermal activation. This behavior follows the Arrhenius law:
![]() | (4) |
In practice, both mechanisms can co-exist. The dominant pathway depends on the light intensity and the material's response to irradiation. Even a modest increase in light intensity can lead to a substantial local temperature rise, shifting the reaction behavior from photochemical to photothermal in nature.43,44
Varied irradiation intensities were utilized to determine the relationship between incident light intensity and H2 production rate as well as to evaluate the relative photochemical and photothermal contributions of the 10% Ni/Al2O3 for methane photolysis (Fig. 1c). Incident light intensity showed a linear rate dependence and photochemical behaviour at the light intensities below 8 W cm−2 and an exponential Arrhenius-type photothermal behaviour at intensities above 8 W cm−2 (Fig. S3c and d).45,46
Charge separation plays a critical role in photochemical mechanisms. While our current study focuses primarily on the photothermal pathway, dominant at light intensities above 8 W cm−2, further investigation into charge carrier dynamics, such as through transient absorption spectroscopy or electron paramagnetic resonance (EPR) studies, is necessary to fully disentangle and quantify photochemical contributions under lower-intensity illumination. Electron and hole scavenger experiments can provide another quantitative insight into the relative contributions of photochemical and photothermal effects in light-driven methane decomposition. As discussed in the literature,47,48 a pronounced drop in conversion upon the addition of scavengers is indicative of photochemical involvement, whereas negligible changes confirm a dominantly photothermal mechanism.43
An IR camera was used to measure the bulk surface temperature of the catalyst under various light intensities in reactor with a CaF2 window (Fig. 1d). The maximum measured temperature was 410 °C due to the intrinsic light absorbance of the material and light intensities used. Bandpass filters were used to evaluate the wavelength dependency of the reaction. Incident intensities were maintained to ensure all samples had the same incident photon flux regardless of wavelength. The total incident optical power on the sample was carefully normalized across all filter conditions, not the transmitted photon flux above the cut-off. This normalization ensures that each experiment is conducted under identical total irradiance conditions, which is a standard approach in photochemical and photothermal comparisons. 495 nm high-pass filter displayed the highest H2 production rate and demonstrated a linear relationship between irradiation intensity and H2 production (Fig. 1e and Fig. S3e). Our catalyst (10% Ni/Al2O3) has a strong and broad absorption band in the visible range (Fig. S8a), particularly between 420–600 nm, which we attribute to plasmonic coupling in Ni nanoparticles (20–80 nm).49,50 The broadband Xe lamp (Fig. 1a) emits significant intensity in the UV, but the catalyst has relatively low absorption in this region due to a minimum near 344 nm. Shifting the light spectrum toward the visible range, either with an AM1.5 filter or long-pass filters, enhances overlap with the catalyst's strong absorption band, increasing photon absorption and thus promoting photothermal activation. Filtering out UV and blue photons (<495 nm) eliminates photons that are poorly absorbed or contribute primarily via slower, and less efficient non-radiative pathways. The remaining photons (≥495 nm) coincide with the broad plasmonic absorption peak, leading to enhanced local heating via fast plasmonic decay processes. Furthermore, based on our Tauc analysis (Fig. S8b), the apparent optical band gap of the catalyst is ∼2.9 eV (∼429 nm). Photons with energies above this threshold may excite interband transitions or be lost to inefficient relaxation. In contrast, longer-wavelength photons (≥495 nm) are better matched to plasmonic and free-carrier heating processes, yielding higher local temperatures and thus higher reaction rates.51–53 As the cutoff wavelength increases beyond 590 nm, the photon energy drops below 2.1 eV, and while still within the absorption band, these photons have significantly less energy per photon. As a result, their contribution to local heating diminishes, reducing the thermal activation of methane cracking. Additionally, the optical absorption of the catalyst begins to decline past ∼600 nm, further limiting the efficacy of these photons in sustaining elevated reaction temperatures.
The pristine 10% Ni/Al2O3 samples were heated to the maximum surface temperatures measured under light irradiation by IR camera to observe the highest catalyst activity under purely thermal conditions. The reaction rates were higher under light irradiation compared to dark (Fig. 1f), confirming that photocatalysis was the driving force behind the observed activity increase.46
Stability under photocatalytic conditions was determined by varying incident light intensity (Fig. S9a–c). Hydrogen production rates decreased over time until it matched the activity of purely the formed carbon nanotubes. The observed decline in activity is due to the progressive blocking of active sites by the accumulated MWCNTs on the catalyst surface. This encapsulation effect reduces catalyst accessibility until the carbon deposits are removed. To evaluate the reusability of the catalyst, extensive cycling tests were performed over ten consecutive reaction cycles, as shown in Fig. S9d. After each cycle, the deposited CNT layer was mechanically removed to expose the active catalyst surface. The results clearly indicate that the hydrogen production rate gradually decreases during each run due to surface coverage by carbon deposits. However, following mechanical carbon removal, the catalytic activity was consistently restored to its initial level. These observations confirm that catalyst deactivation is primarily physical rather than chemical in nature, arising from surface blockage rather than irreversible structural or compositional degradation. Thus, the Ni/Al2O3 catalyst remains stable and can be readily regenerated through simple removal of the carbon layer.
To validate the scalability of our approach, we developed a customized reactor scaled 36 times larger than the bench-scale model. This advanced system incorporates a 400W white LED light source, integrated with a gas chromatograph (GC) and H2 and CH4 sensors, as illustrated in Fig. S10. Remarkably, the scaled-up system maintained the same selectivity for H2 and MWCNTs (Fig. S11). These findings confirm the feasibility and practicality of our approach for large-scale hydrogen and CNT production, preserving both high selectivity and operational efficiency. We have also conducted an energy requirements calculation of our system (Fig. S12, S13 and Table S2). Our results demonstrate that our photocatalytic approach significantly reduces energy requirements, requiring approximately 70% less energy for H2 production compared to water electrolysis and at least 74% less energy for CNT production compared to commercial methods. Additionally, our approach offers cost-effectiveness and substantial reductions in greenhouse gas (GHG) emissions, as validated through our techno-economic analysis (TEA) and life cycle assessment (LCA) (discussed in detail in a later section). These findings underscore the potential of our technology as a scalable and sustainable alternative for hydrogen and CNT production.
X-ray photoelectron spectroscopy (XPS) was used to determine the surface chemistry and the elemental composition of the catalyst prior to and post thermal and methane photolysis reactions. The spent catalyst post-irradiation showed more elemental carbon than the pristine sample. These results confirm the formation of the MWCNT following methane photolysis (Fig. 3d and Table S3). The sharpness of the C 1s peak in the post-methane photolysis indicates a high C–C
:
C
C ratio and lower C
O content compared to the pristine catalyst and confirms the ordered nature of the formed carbon (Fig. 3e). Lack of changes to the Ni 2p peaks in the high-resolution XPS spectra after methane photolysis confirms no structural change on the catalyst following reaction at low irradiation (Fig. S14).
Thermogravimetric analysis (TGA) was used to confirm the crystallinity of the formed MWCNT by measuring the sample's weight change over time as a function of temperature. Spent, fresh catalyst, and commercial MWCNT showed no thermal degradation between 150–550 °C in the amorphous carbon thermal degradation range, indicating high MWCNT crystallinity (Fig. 3f). We also conducted energy-dispersive X-ray spectroscopy (EDX) to assess the purity of the separated CNTs and quantify residual nickel and alumina impurities. Across more than 15 scanned spots, the nickel content did not exceed an average of 2%, while alumina remained below 8% (Fig. S15).
We observed that the thermal cracking of methane (methane pyrolysis) at 500 °C produced a mixture of ordered carbon, such as CNTs, and disordered carbon, like carbon black (Fig. 4a). In contrast, methane photolysis yielded only MWCNTs (Fig. 4b). The average diameter of the formed MWCNT was about 17 ± 5 nm (Fig. 2c). HRTEM images confirm the multiwalled structure of this photo-grown CNT (∼40 wall) and also reveal an interlayer spacing of ∼3.33 Å, similar to single crystal graphite interlayer spacing58 (Fig. 4d and e).
ID/IG ratios in Raman spectroscopy identified the quality of the formed carbon.58 High ID/IG indicates more disordered carbon, while low ID/IG ratios indicate a more ordered sp2 hybridized carbon. The pristine catalyst showed no carbon presence and did not feature any ID/IG ratio (Fig. 4f). The post-irradiation spent catalyst featured a lower ID/IG ratio compared to post-thermal spent catalyst (Fig. 4f). Increasing the irradiance from 8 W cm−2 to 20 W cm−2 increased ID/IG, likely resulting from higher local temperatures under higher light intensities (Fig. 4f). The separated MWCNT collected from the irradiated catalyst, showed a slightly higher ID/IG compared to commercial Sigma MWCNT due to the presence of surface oxygen functional groups (Fig. S16d).
The C 1s high-resolution XPS spectrum of the mechanically separated MWCNT showed C–C and C–OH, C
O, O-C
O and π–π* peaks at 284.3, 284.7, 286, 290 eV respectively (Fig. S16a). The O 1s high-resolution XPS spectra of the mechanically separated MWCNT confirmed the formation of oxygen functional groups (oxygen, C–O, C–OH, and C
O at 533.2, 531.6, and 530.1 eV, respectively) on the MWCNT surfaces through changes in the integrated peak areas (Fig. S16b). FTIR spectrum of the mechanically separated O-MWCNT, confirms the formation of oxygen functional groups on the MWCNT surfaces (Fig. S16c). Oxygen functionalization of MWCNT was attributed to the high local temperatures involving chemical reactions between the CNT surface and oxygen-containing species from the catalyst (Ni's high local temperature may facilitate the oxidation of CNT using oxygen from the alumina support), trace oxygen in methane (3.7 UHP Methane, 99.7%), or surface defects.59 The high local temperature weakens and activates specific C–C and C–H bonds on the CNT surface, particularly at defect sites, edges, or unsaturated bonds, making them more reactive.59
Brunauer–Emmett–Teller (BET) surface area analysis and Barrett–Joyner–Halenda (BJH) pore size and pore volume distribution determined textural properties such as surface area, the pore size and pore volume distribution of the pristine and spent catalyst, and the separated MWCNT (Fig. S16, S17 and Table S5). The nitrogen adsorption–desorption isotherm of the separated MWCNTs exhibits a low uptake profile with a Type IV-like hysteresis loop (Fig. S16e), which is characteristic of mesoporous materials and commonly observed in CNT-based structures.60 The BJH pore size distribution (Fig. S16f) reveals a broad range of pore widths extending into the macroporous region, with apparent pores >200 nm. While such macroporosity may reflect interstitial voids between entangled CNT bundles or structural heterogeneity.61
As expected, the fresh catalyst exhibits higher N2 uptake than both spent samples, consistent with textural degradation resulting from CNT formation. The N2 adsorption isotherm curve for the photolysis-treated catalyst shows more uptake, especially at higher relative pressures, indicating higher surface area and/or less pore blockage due to the formation of CNT on the former's surface (Fig. S17e). This supports our interpretation that photolysis leads to CNT formation on the external surface rather than inside pores, thereby preserving more accessible surface area compared to the pyrolysis-treated catalyst, which may be more clogged or blocked. The fresh catalyst shows a sharp peak around ∼5 nm (mesopores), with a drop in pore volume after reaction (Fig. S17f). Both spent catalysts show narrowed and flattened distributions, consistent with pore blockage due to CNT deposition (Fig. S17f).
Comparatively, methane photolysis forms base-grown CNTs, where the metal is attached firmly to the support and the CNT stems from it (Fig. 4b and Fig. S17b). This growth difference can be attributed to high local temperatures as opposed to bulk temperatures during reactions.63,64 High local temperatures from photothermal heating drive methane decomposition and facilitate the nucleation and growth of MWCNTs by enabling carbon atoms to diffuse and arrange themselves into cylindrical graphitic layers with enhanced crystallinity and fewer defects, thereby improving their mechanical and electrical properties and resulting in facile mechanical separation without degradation. The methane photolysis reaction over 10% Ni/Al2O3 can be described in three key steps: methane activation, carbon nucleation, and carbon deposition to form carbon nanomaterials.65 Methane molecules adsorb dissociatively on the Ni metal particle surface through C–H bond photoactivation. After taking further steps, carbon atoms lead to the simultaneous release of molecular hydrogen. The adsorbed carbon atoms diffuse on the surface or through the bulk Ni metal to a suitable area for CNT formation. Previous studies in the literature showed the effect of the Ni particle size and support interaction on the formation of different carbon nanomaterials including carbon nanotubes, nano-onions, and nanofibers.22 The base-growth mechanism can be attributed to strong Ni-support interaction forcing carbon precipitation on only the exposed surfaces of the Ni particles rather than the interface between Ni and support. This results in carbon deposition and growth on the surface of Ni particles at the gas–solid interface. As a result, the Ni particle remains fixed to the support, and close-ended MWCNT grows on the Ni surface.22
We have also investigated the effect of different light intensities on the formation of MWCNT. Our findings indicate that the d-spacing (interlayer distance) of the MWCNTs decreases from 0.38 nm to 0.33 nm as the light intensity increases from 8 to 20 W cm−2, with no complete CNT formation observed at light intensities ≤3 W cm−2 (Fig. S18a). The local temperature generated by light during methane photolysis significantly influences the formation of MWCNTs and its d-spacing.66–68 At low light intensities (≤8 W cm−2), incomplete decomposition of methane resulted in carbon atoms being deposited in a less organized fashion, leading to increased d-spacing of 0.38 nm due to the presence of structural defects, amorphous carbon, and weakly bound graphene layers (Fig. S18b).
As light intensity increases to 20 W cm−2, high local temperatures increase methane decomposition efficiency, resulting in a higher flux of reactive carbon species. This enhances graphitization and a reduction in d-spacing to 0.33 nm as graphene layers form more tightly packed and well-ordered structures (Fig. S18c and d). However, excessively high light intensities can produce high local temperatures, which may cause thermal stress or non-uniform carbon deposition, potentially introducing new defects that increase d-spacing. Therefore, an optimal light intensity range exists for methane photolysis that minimizes d-spacing and maximizes the structural integrity of the MWCNTs. The increased degree of graphitization was confirmed through PXRD analysis of the spent catalyst under varying light intensities. The (002) peak intensity of the formed MWCNTs progressively increased with light intensity, indicating enhanced structural order and graphitization69 (Fig. S3e). The Raman spectra of the MWCNTs formed at different light intensities showed an increasing ID/IG ratio. This change is observed when increasing the irradiance from 8 W cm−2 to 20 W cm−2 (Fig. 4f).
| * + 2CH4(g) → *_C2H2 + 3H2(g) |
Following the formation of adsorbed C2H2, subsequent additions and dehydrogenations of CH4 can lead to an initial layer of CNT as shown in Fig. S20. In our model, proceeding from an adsorbed C2H2 fragment to this first layer of CNT (given by 12 sp2 carbon atoms with C–C pairs alternating closer to and further from the surface and the pairs further from the surface capped by H atoms) corresponds to adding and releasing another 10 CH4 and 18 H2 molecules, respectively. The net energy change is +22.43 eV or approximately +2.24 eV for each additional carbon added.
For the two energetically demanding steps mentioned above, we further performed linear response time-dependent DFT (TD-DFT) to investigate possible pathways for light to facilitate the reaction. While the 2nd dehydrogenation is the most energetically demanding step, we found major excitations across from the computed absorption profile of the *_(C2H4 + 2H) intermediate (structure 9 in Fig. 5c) correspond to transitions that involve an increase in electron density on one adsorbed hydrogen and a decrease in the other, suggesting that light can help facilitate the formation of a proton and hydride towards the formation of H2 heterolytically. We also observe transitions that involve an increase in electron density between the adsorbed C2H4 fragment and the Ni cluster. This suggests the light can also help facilitate a stronger bonding interaction between the carbon fragment and the metal cluster, which helps prevent the desorption of the adsorbed C2H4 as ethylene prematurely and allows the carbon fragment to remain on the surface for further dehydrogenation and building of the initial CNT layer. For the complete absorption profile and the selected difference densities, we refer to Fig. S21. For the *_2CH3 intermediate (structure 6 in Fig. 5c), we observe multiple transitions – again within major excitations across the computed absorption spectrum (Fig. S22) – that lead to a decrease in electron density on one CH3 fragment and an increase in another, suggesting that these excitations could help form CH3+ and CH3− fragments that facilitate heterolytic C–C coupling. Together, these results demonstrate that light plays a critical role in facilitating the key steps of CNT nucleation: C–C bond formation, dehydrogenation, and preventing early desorption of ethylene fragments.
The CNT anode's initial charge capacity (during lithium insertion) closely matches its subsequent discharge capacity (during lithium extraction), suggesting a stable electrochemical performance (Fig. 6c). Notably, the self-standing anode exhibits consistent and reversible lithium-ion insertion and extraction. Furthermore, it maintains minimal voltage hysteresis over 10 cycles, demonstrating stable cycling performance and an extended cell lifetime. When compared to CNT-based anodes reported in the literature, our electrode shows lower voltage hysteresis and longer cycle life, thus demonstrating its excellent performance.75 To investigate the reduction and oxidation behavior of the electrode, cyclic voltammetry (CV) of CNT/Li half cells was measured at 0.2 mV s−1. The CV curves reveal essential processes within the cell, with significant anodic and cathodic peaks indicating Li+ extraction and insertion, respectively.76 The cathodic peak at ∼0.03 V is attributed to the intercalation of lithium ions into graphitic layers, while the anodic peak at 0.21 V is due to the deintercalation of lithium ions (Fig. S23).77 This observation aligns with the CV graphs reported by Liu et al.78 and Zhao et al.,79 which exhibited a similar trend for CNT. In the initial cycle, no new strong peaks appeared, and the peak potentials remained stable, indicating the electrochemical stability of the CNT anode across the entire voltage range.79 The CV curves for the first three cycles mostly overlap, except for the reduction peak around ∼1.0 V, which is present in the first and second cycles but disappears in the third. This suggests that the formation of the solid electrolyte interphase (SEI) effectively prevents further decomposition of the electrolyte after the initial cycles.79
EIS measurements were taken to evaluate the resistive properties of the electrodes, as shown in Fig. 6d. The EIS data between 2 MHz and 0.01 Hz were fitted using MEISP software based on an equivalent circuit model. The impedance information was modelled using an electrical equivalent circuit shown in the inset of Fig. 6d, which shows accurate reproductions of the arcs representing Rs, Rct, and Rsei.73 The intersection on the real axis at high frequency, referred to as the equivalent series resistance (Rs), indicates the combined ionic resistance of the electrolyte and the electronic resistance of the active material in the electrode. The semicircle observed at high frequencies represents the resistance of the solid electrolyte interface (Rsei). In addition to this high-frequency semicircle, and the mid semicircle, attributed to the charge transfer impedance between the electrolyte and the electrode, another distinct arc, identified as the low-frequency Warburg impedance tail, signifies the impedance resulting from the diffusion of ions into the active materials.80 From the EIS fitting of the graph, it can be observed that the values of Rs, Rsei, and Rct are 7.72 Ω, 54.76 Ω, and 119.73 Ω, respectively. The values of Rsei and Rct demonstrate the dominance of the cell impedance, with the increase primarily attributable to charge-transfer resistance. In addition, it can be seen that Rsei is relatively large, indicating that the formation of the SEI plays a significant role in influencing the insertion and extraction of lithium ions.81
First, methane is fed to a reactor, where a supported nickel-based material is used to catalyze the photolysis reaction. The required light is supplied via an ancillary LED system powered by photovoltaics. The reactor system is designed considering a modular approach, in which the total number of reactor vessels (modules) is determined based on the amount of energy available per modular reactor (δSE), details on the calculation of the required number of modules are presented elsewhere.1 To ensure a 24/7 operation, a battery system is installed. Batteries are sized based on the duration of an average night at the plant location (15.41 hours). For this study, we have chosen a place in Southern California, close to the town of Daggett. Location and night length were established based on the analysis by Peng and Maravelias.82 The system is sized such that ∼11.3 Tons H2/h can be obtained; this size is in line with other works for the production of hydrogen and advanced carbon materials from methane.83
There are two reactor outlets (1) a mostly gaseous stream (stream 5) consisting of unconverted methane, hydrogen, and elutriated solids, and (2) a solid stream containing most of MWCNT and catalyst (stream 4). As MWCNT grow on the surface of the catalyst, it is necessary to remove a fraction of the catalyst from the reactor to recover this product. The small fraction of solids elutriated in the gas stream are removed using a cyclone. The gas stream without solids (stream 6), consisting of hydrogen and methane, is fed to a pressure swing adsorption (PSA) unit. This unit operates at 35 bar and produces pure hydrogen that can be sold to the market (stream 10), the PSA recovery is 85%.83 The other outlet stream of the PSA unit consists of unconverted methane along with unrecovered hydrogen (stream 11). This stream is recycled back to the reactor. To ensure that the gas phase in the reactor consists mostly of methane, we impose a constraint limiting the fraction of hydrogen in the reactor feed to be below a predefined limit (δH2 = 20%). The solid stream produced by the reactor (stream 4) is separated into three streams, one containing catalyst (stream 13) that is recycled back to the reactor; one consisting of MWCNT (stream 12) that is collected, and a stream consisting of CO2 (stream 14).
The process to separate the catalyst from the MWCNT is still in the early stages of development. Rather than describing a specific set of unit operations, we adopt a more general approach, in which we assume that it is possible to (1) recover a functional catalyst to be recycled once all carbon attached to its surface is removed (2) that a fraction of the carbon attached to the catalyst surface can be recovered to be sold (3) that the fraction of MWCNT that cannot be recovered from the catalyst surface can be removed via incineration without affecting the catalyst activity, and (4) that the catalyst has a shelf life of a year and it is replaced continuously by adding/removing new/old catalyst to/from the system. We note that the reusability of the catalyst is backed by the results presented in the experimental section. The magnitude of these streams is such that the reactor catalyst load is replaced in a year. Despite the generality of the modeling approach, we note that the system proposed is consistent with the separation scheme reported by Parmar and coworkers,84 this scheme consists of three operations (1) sonication for the removal of the MWCNT (2) magnetic separation of the catalyst fraction (3) thermal regeneration of the catalyst to remove residual carbon.
Capital costs of major pieces of equipment are estimated using different approaches: (1) for equipment commonly used in the chemical industry (cyclone, compressor) we use known cost correlations.85 (2) For pressure swing adsorption, we use available cost estimations from the literature and the six-tenths rule for scaling.86 (3) For batteries, we use reference costs based on DOE estimations for energy storage and a linear scaling relation87 (4) for novel equipment requiring a tailored design (LED-based reactor and catalyst recovery system), we use a reasonable initial estimate and perform a sensitivity analysis to determine the influence that the capital cost of these pieces of equipment has on the minimum selling price of the main product. The operating costs of the system are estimated based on material and energy balances.
The economics of this system is a strong function of several physical/economic parameters whose value is susceptible to change as technology develops. We have grouped these parameters into five categories as shown in Table 1. The value for the parameters in the base case analysis, along with a range of possible values is shown. This range is established such we can explore the consequences of technological improvements over the next 30 years, and the effect of overoptimistic assumptions in the base case design. While the meaning of many of the parameters is evident, some of them require further explanation: The light-to-chemical efficiency (ηCH) denotes the fraction of energy contained in photons that is transformed into chemical energy. The LED power conversion efficiency (ηLED) denotes the amount of energy that is converted into light with respect to the electric energy inputs to the LED system. The energy available per modular reactor (δSE) characterizes the total amount of energy that the LED system can supply to each of the modular reactors. Finally, the fraction of carbon recovered (λC) determines the fraction of MWCNT that can be recovered as a pure and usable product. The reference values for most parameters are determined based on the current state of the art for the involved technology. For example, for roundtrip battery efficiency (ηB), we use a value of 85%, which is consistent with current estimations for Li-ion batteries.87 A few of the parameters were assumed based on a reasonable estimate. Of particular importance is the fraction of carbon recovered (λC), in this case, we adopt a conservative initial estimate to reflect the uncertainty of the process. Likewise, the light-to-chemical efficiency is assumed to be 10%, a low value that can improve as better optothermal chemical reactor designs are developed.
| Category | Parameter | Symbol | Units | LB | BC | UB | Ref. |
|---|---|---|---|---|---|---|---|
| LED/batteries | Round trip battery efficiency | ηB | % | 80 | 85 | 95 | 87 |
| Light-to-chemical efficiency | ηCH | % | 5 | 10 | 80 | Assumed | |
| LED power conversion efficiency | ηLED | % | 20 | 32.5 | 86 | 88 | |
| Energy per modular reactor | δSE | MW | 0.05 | 0.1 | 1 | 45 | |
| Capital cost of batteries | θB | $ per KWh | 104 | 414 | 500 | 87 | |
| Reactor capital cost | θR | $MM per [kg s−1]0.6 | 0.5 | 1 | 10 | Assumed | |
| Catalyst recovery | Fraction of carbon recovered | λC | % | 10 | 25 | 90 | Assumed |
| Catalyst regeneration capital cost | θCAT | $MM per [kg s−1]0.6 | 0.5 | 1 | 10 | Assumed | |
| Feedstock and byproduct costs | Selling price for H2 | λCAT | $ per kg | 0.5 | 1 | 10 | 89 |
| Purchase price of CH4 | λCH4 | $ per kg | 0.1 | 0.16 | 0.48 | 90 | |
| Utilities | Electricity price | λE | $ per KWh | 0.03 | 0.06 | 0.09 | 91 |
| Chemistry | Conversion | χ | % | 1 | 8.7 | 90 | This work |
First, we study the economics of the base case (BC). A breakdown of the minimum selling price (MSP) for the base case is shown in Fig. 7(B). An MSP of $106.3 per kg was obtained for MWCNT assuming a hydrogen price of $1 per kg. These results highlight the potential of the proposed technology, considering that the market price for MWCNT has been estimated to be between $100 per kg and $1000 per kg.92 Furthermore, these results are obtained using the assumption of a hydrogen price of $1 per kg, a price in line with DOE targets.89 To further explore the relation between the MSP for the MWCNT and the hydrogen byproduct, we present a sensitivity analysis on the supplementary information (Fig. S24); we show that the MSP of MWCNT is relatively incentive to the price of H2. For reference, if the price of H2 drops to zero, the MSP of MWCNT increases by ∼1.25%. Conversely, if the price of H2 increases five times the MSP of MWCNT is reduced by 5.25%. The annualized capital and operating costs are shown in Fig. 7(c). The dominant operating cost is the electricity (∼79%) and the dominant capital cost is the batteries (∼87%). The analysis of capital cost reveals that a major limitation to the deployment of the proposed technology is the high capital investment required (∼47 billion for the proposed design).
To identify the critical parameters affecting the MSF, we perform a sensitivity analysis. The results are presented in Fig. 7(c), where we show the percentage change in the MSP of MWCNT when the value of each parameter evaluated changes to the lower or upper bounds in Table 1. Since batteries dominate the capital cost and are the major component of the MSP, it is not surprising that the parameters related either to the price of energy storage (capital cost of batteries), or the amount of energy that needs to be stored (light to chemical efficiency and LED power conversion efficiency) are the most important economic drivers. In addition to these parameters, we also find that the fraction of MWCNT that can be recovered significantly influences the economics of the system. We also highlight that the capital cost of the reactor may have an important impact on the MSP if significantly larger capital costs in comparison with the base case are required. Conversely, the cost of the catalyst recovery system is not critical for the economics of the system, an increase of 10× with respect to the base case only has a marginal effect on the MSP. These results point toward the need to develop better optothermal chemical reactors for this system both requiring low capital costs and leading to a more efficient utilization of photons.
In Fig. 8, we explore the effect that incremental improvements in the parameters impacting the cost will have on the MSP. In the figure, we show both the breakdown of the MSP, and the total capital investment (TCI) required. Dramatic improvements in the MSP and TCI can be obtained if the projected improvements in key parameters are achieved. In this respect, we note that the cost of energy storage in batteries is projected to drop by more than 75% in the next 25 years.93 Furthermore, the power conversion efficiency is also projected to increase to close to 86%,88 in our calculations, we are conservative and use a more moderate 65%. Along the same lines, the efficiency of batteries has been reported to be as high as 95%,87 but for the projection, we use a conservative 90%. Finally, we note that the SunShot target for utility-based photovoltaics electricity is $0.03 per KWh.91 Regarding the development of our system itself, we assume three main improvements (1) that ηCH can be increased by 3×, this is tied to the development of better reactors (2) that conversion can be raised without major changes in selectivity to 50%, this is contingent on the development of better catalysts (3) that the fraction of carbon recovered can be as high as 90%, this depends on the design and optimization of the catalyst recovery system and (4) that the available energy per modular reactor can be increased 5×, this is also contingent on the reactor design.
![]() | ||
| Fig. 8 Breakdown of the minimum selling price of MWCNT as a function of changes in the key parameters associated with the system. | ||
There are two important conclusions from Fig. 8. First, we see that the MSP in the most optimistic case can be reduced by two orders of magnitude to as low as $1.2 per kg, this is encouraging, considering that the market price of MWCNT can drop if the offer for this type of material starts to increase as this and other competing technologies develop. The second point is that the total capital investment (TCI) will also reduce as the technologies evolve, in the most optimistic case a ∼20× reduction is observed, bringing the TCI close to 2 billion.
![]() | ||
| Fig. 9 Global warming potential for the proposed technology under different scenarios. S1 and S3: base case for technology, S2 and S4: best case for technology development. | ||
Supplementry information (SI) is available. See DOI: https://doi.org/10.1039/d5ee02120k.
| This journal is © The Royal Society of Chemistry 2025 |