Jose Luis Mendoza-Cortes
Jose L. Mendoza-Cortes | |
---|---|
Alma mater |
|
Known for | Theoretical Physics & Chemistry / Computational Physics / Material Science / Computational Engineering |
Awards |
|
Scientific career | |
Fields | Physics, Chemistry, Materials Science and Engineering, Scientific Computing, Computational Mathematics |
Institutions | |
Thesis | Design of Molecules and Materials for Applications in Clean Energy, Catalysis and Molecular Machines Through Quantum Mechanics, Molecular Dynamics and Monte Carlo Simulations. (2012) |
Doctoral advisor | William A. Goddard III |
Other academic advisors | Martin Head-Gordon (PD) |
Jose L. Mendoza-Cortes is a theoretical condensed matter physicist and material scientist specializing in computational physics, materials science, chemistry, and engineering. His studies include methods for solving Schrödinger's or Dirac's equation, machine learning equations, among others. These methods include the development of computational algorithms and their mathematical properties.[2]
Because of graduate and post-graduate studies advisors, Dr. Mendoza-Cortes' academic ancestors are Marie Curie and Paul Dirac.[3] His family branch is connected to Spanish Conquistador Hernan Cortes and the first viceroy of New Spain Antonio de Mendoza.
Mendoza is a big proponent of renaissance science and engineering, where his lab solves problems, by combining and developing several areas of knowledge, independently of their formal separation by the human mind.[2] He has made several key contributions to a substantial number of subjects (see below) including Relativistic Quantum Mechanics, models for Beyond Standard Model of Physics, Renewable and Sustainable Energy, Future Batteries, Machine Learning and AI, Quantum Computing, Advanced Mathematics, to name a few.
Education
Throughout his school years he earned top honours in the national Knowledge Olympiad at the primary-school level and, in high school, went on to win national and international Olympiads in chemistry, informatics, mathematics, and physics. He completed perfect problems in the 34th International Chemistry Olympiad at Groningen, Netherlands 2002.[4]
Mendoza completed his B.Sc. in chemistry and physics from Tec de Monterrey (ITESM), Monterrey, Mexico in 2008. During this time, he had an interchange program in the last two years of his B.Sc. to finish all the master's degree classes at the University of California, Los Angeles. Following this, he moved to Pasadena, California to complete his M.Sc. at California Institute of Technology (CalTech)in 2010. After the completion of his M.Sc., he stayed at Caltech and completed his Ph.D. in physics in 2012. His research advisor was William Goddard III and his dissertation title is "Design of Molecules and Materials for Applications in Clean Energy, Catalysis and Molecular Machines Through Quantum Mechanics, Molecular Dynamics and Monte Carlo Simulations."[5] He completed his postdoctoral studies at University of California, Berkeley.
Career
During his undergraduate studies, Dr. Mendoza was awarded the Newcomb Cleveland Prize of the American Association for the Advancement of Science (AAAS) is annually awarded to the author(s) of outstanding scientific paper published in the Research Articles or Reports sections of Science. This is AAAS's oldest and most prestigious award. Specifically, Dr. Mendoza-Cortes synthesized and designed the first 3D-Covalent organic framework (COF), ever, COF-103 and COF-108, helping unleash this new field. Besides synthesizing them, Dr. Mendoza-Cortes created the computational models that would simulate their X-ray pattern, thus identifying and characterizing their chemical structures.[6]
Following his graduation, Mendoza joined the Caltech & Joint Center for Artificial Photosynthesis (JCAP) as a staff scientist until 2013 and then as a postdoctoral fellow at the California Institute of Technology, where he served until 2014. He started the theory and simulations arm of JCAP at Caltech and then moved to UC Berkeley.
In 2015, he started as a Faculty with Florida State University at the Department of Physics, Scientific Computing, Chemical and Biomedical Engineering, Materials Science and Engineering until 2020. During this time, he was also a scientist at the National High Magnetic Field Laboratory and Condensed Matter group. He is credited with starting and developing the first class in Quantum Computing and Machine Learning at Florida State University.
Dr. Mendoza is currently a Faculty at the Department of Physics and Astronomy & Chemical Engineering and Material Science at Michigan State University. He created several courses combining Machine Learning, Physics, Chemistry, Materials Science, and Quantum modeling to create materials starting at the atomic scale.
His work and reputation have already led to significant national attention as he is the only researcher to be named four times in a row to the prestigious Scialog Fellowship (2020–2023) for his contributions to the development of negative emissions technologies.[7] This is a fellowship is for only 50 faculty per year including both the US and Canada. His works on the amphidynamic behavior in oligo-functionalized covalent-organic frameworks were selected as one of the 2018 Emerging Investigators collection from the Royal Society of Chemistry.[8] He was also the recipient of the GAP awards in 2018 from Florida State University for his work on creating the database to reliably predict which compounds will produce materials with the most desirable properties for a given purpose.[9]
He was part of the American Physical Society (APS) national committee on diversity and inclusion (9 persons), which developed the Bridge program; which has now expanded into the Inclusive Graduate Education Network (IGEN) which is made of 30 societies (including ACS, MRS, APS), corporations, and national laboratories, which is considered one of the most influential programs in post-graduate education for minorities in the USA.[10][11]
Dr. Mendoza's research has been featured in Forbes,[12] CNBC,[13] MRS Bulletin,[14] C&EN News, Public Radio, Laser Focus World magazine, and the DOE Highlights, to name a few. This work has been disseminated across more than 60 national and international invited and keynote lectures at scientific meetings and universities all over the world.
Published work
As an independent researcher, Dr. Mendoza-Cortes' work has been cited over 9,000 times with an average of over 157 citations/paper, as well as Erdős number = 5, H-index = 32, and i10-index = 45.[15]
Beyond standard model of physics
In 2024 I. M. Morris, K. Klink, J. T. Singh, S. S. Nicley, J. N. Becker and J. L. Mendoza-Cortes proposed using synthetic diamond doped with extremely rare, "pear-shaped" radioisotopes to probe fundamental symmetries of nature.[16] The article appears in the themed issue "Diamond for quantum applications" and was featured on the journal's cover. Physics motivation: Detecting a permanent electric dipole moment (EDM) in an elementary particle or nucleus would signal new physics beyond the Standard Model. Certain isotopes with large nuclear octupole deformations, notably 229Pa, amplify EDM-related effects via an enhanced Schiff moment. Because such isotopes are produced only in minute quantities at rare-isotope facilities, any solid-state sensor must trap and repeatedly interrogate single atoms with high efficiency. Diamond-defect platform: Density-functional calculations show that substituting 229Pa into a carbon vacancy creates a colour centre whose electronic states lie deep in the diamond band gap, leading to narrow optical lines and long spin coherence times. Lanthanide proxies (e.g., Pr3+, Tb3+) form isoelectronic defects with nearly identical crystal fields, providing non-radioactive testbeds for growth and spectroscopy before radioactive implantation. The defect structure combines molecule-like localisation with the mechanical and chemical robustness of diamond, enabling laser cooling, high-fidelity read-out and integration into nanophotonic resonators. Impact and outlook: The study outlines a quantum-sensing roadmap that merges rare-isotope production (for example at MSU's Facility for Rare Isotope Beams) with diamond-based quantum technology. Beyond EDM searches, rare-isotope colour centres could advance nuclear-clock development and tests of time-reversal symmetry. Philosophical Transactions of the Royal Society, the world's oldest continuously published scientific journal, has previously featured seminal work by Isaac Newton, Dorothy Hodgkin, Alan Turing, Charles Darwin, Michael Faraday, James Clerk Maxwell, and Stephen Hawking, among others. The 2024 cover highlights how engineered diamond defects continue that tradition by opening new windows on fundamental physics.
See also: | Electric dipole moment | Colour centre | Nitrogen-vacancy center | Facility for Rare Isotope Beams | ISOLDE | Radioactive Isotope Beam Factory | Grand Accélérateur National d'Ions Lourds | Facility for Antiproton and Ion Research | RAON |
Quantum computing
In 2025 the Mendoza-Cortés group released an open-access e-book, "Machine Learning and Quantum Computing Guide for Humans," which packages more than 300 pages of tutorials and executable Jupyter notebooks into a single, continuously updated resource.[17] Designed for students with only basic linear-algebra and Python backgrounds, the guide integrates: Quantum-algorithm notebooks: step-by-step implementations of the Deutsch–Jozsa, Grover and quantum-Fourier-transform algorithms in Qiskit, plus optimisation examples on D-Wave quantum annealers. Hardware primers: concise introductions to IBM Q superconducting qubits, trapped-ion systems and photonic processors, emphasising current qubit counts, coherence times and error-mitigation strategies. Bridging topics: quantum machine learning (QML), variational quantum eigensolvers, quantum support-vector machines and near-term "Noisy Intermediate-Scale Quantum" (NISQ) applications. Because all examples run either on free cloud back-ends or local simulators, the guide functions as a "lab manual" for classrooms lacking dedicated quantum hardware. It has been adopted in graduate seminars at Florida State University, Michigan State University and downloaded thousands of times within its first year, filling a gap between highly technical monographs and marketing-level introductions.
See also: | Quantum computing | Qiskit | D-Wave Systems | Quantum machine learning | Machine learning |
Advanced mathematics
Recent work by E. R. Dolores-Cuenca and J. L. Mendoza-Cortés extends classic number-theory results to the language of partially ordered sets (posets) and their operads, broadening the bridge between combinatorics and analytic number theory.[18] Operad of finite posets. The authors treat finite posets as elements of an operad whose algebra naturally encodes operations such as disjoint union. Mapping each poset to its order polytope places the combinatorial data in a geometric setting that is amenable to generating-function techniques. Zeta-value identities. By generalising a family of series identities, they derive relations among multiple zeta values (MZVs) that mirror the poset's decomposition into chains and antichains. The approach reproduces, and unifies, several classical formulas for ζ(n) that trace back to Srinivasa Ramanujan. Eulerian numbers reinterpretation. Ramanujan's results on Eulerian numbers are re-proved by showing that the relevant series inherit the algebraic structure of disjoint unions of points inside the operad. This yields shorter, conceptually transparent proofs compared with analytic methods. Linear independence criteria. The poset framework provides a new perspective on why certain ζ-value combinations are conjectured to be linearly independent over the rationals; specific independence results follow by analysing how operadic compositions factor through order polytopes. Together with Mendoza-Cortés's broader programme on the mathematical foundations of machine learning and quantum algorithms, this work illustrates how modern algebraic tools, operads, polytopes and posets, can unlock fresh insights into long-standing problems that resist classical analytic approaches.
See also: | Eulerian number | Multiple zeta values | Operad | Partially ordered set | Srinivasa Ramanujan |
Machine learning (ML) and Artificial Intelligence (AI)
The mathematics of ML and AI
Polychronisation, a term coined in computational neuroscience, describes how neurons (or abstract signal-processing nodes) can fire in precisely timed patterns when multiple inputs coincide. In 2023 Eric Dolores-Cuenca and colleagues turned this biological idea into a rigorous graph-theoretic framework.[19] Model. A node emits an output if it receives (i) an external stimulus or (ii) two synchronous inputs whose summed amplitude exceeds a fixed threshold. Collections of such time-locked firings are called polychrony groups. Cascades and "profit''. When the number of nodes triggered by prior activations exceeds the number of nodes that were externally stimulated, the authors define a cascade with a profit equal to that excess. Graph dictionary. Each cascade maps to a directed graph dubbed a chinampa (inspired by the interlinked agricultural islets of pre-Hispanic Mexico). This correspondence allows polychrony dynamics to be analysed via classic graph and poset tools. Topological classification. By enumerating chinampas with profit 0 and 1, the study links their counting problem to computing Stanley order-polynomials of certain partially ordered sets, opening algebraic routes to predict cascade statistics. Algorithms. The paper supplies polynomial-time procedures to (i) predict whether a given vertex will activate at a prescribed time and (ii) reconstruct an entire cascade from minimal initial data, capabilities relevant to explainable artificial-intelligence models that rely on signal-timing codes rather than weight magnitudes. This synthesis of nonlinear signal timing with combinatorial topology enriches the mathematical toolkit underlying machine-learning interpretability, neural-coding theory and asynchronous circuit design.
See also: | Spike-timing-dependent plasticity | Graph theory | Partially ordered set | Stanley symmetric function | Neural coding |
Machine-learning handbook
In 2025 the Mendoza-Cortés group released an open-access e-book, Machine Learning and Quantum Computing Guide for Humans, aimed at readers outside traditional computer-science programmes.[17] Scope and structure. The 300-page handbook couples concise theory summaries with interactive Jupyter notebooks covering staple algorithms: linear and logistic regression, k-nearest neighbours, decision trees, random forests, support-vector machines, convolutional and recurrent neural networks, Bayesian optimisation, genetic algorithms, non-negative tensor factorisation and more. Domain-specific examples. Each chapter ends with case studies drawn from art style transfer, materials discovery, quantum-state tomography, medical image segmentation, and game-level generation, illustrating how to tailor hyper-parameters and data pipelines for non-CS disciplines. Quantum-computing modules. Dedicated sections introduce superposition, entanglement and qubit noise, then walk readers through implementing Grover's search, the quantum Fourier transform and variational quantum eigensolvers on IBM Q and D-Wave back-ends via Qiskit and Ocean SDK notebooks. Best-practice workflow. The guide emphasises reproducibility (version-controlled notebooks, environment files), model interpretability (SHAP, LIME) and benchmarking against automated ML platforms, helping users decide when a quick "AutoML" pass suffices and when custom tuning yields significant gains. Adopted in cross-disciplinary courses at Michigan State University and Florida State University, the resource fills the gap between beginner tutorials and advanced monographs, enabling engineers, artists and scientists to deploy modern ML and quantum tools on their own research problems.
See also: | Machine learning | Quantum computing | Qiskit | Automated machine learning | Jupyter notebook |
Order-theoretic neural networks
A 2024 preprint by E. Dolores-Cuenca, A. Guzmán-Sáenz, S. Kim, S. López-Moreno and J. L. Mendoza-Cortés introduced an explicit connection between order theory, the study of partially-ordered sets (posets), and modern deep-learning architectures.[20] Building on earlier work that equates integer-valued neural networks with tropical-geometry maps, the authors: Order polytopes as network diagrams. Every finite poset with n elements has an associated order polytope, a convex region inside the unit n-cube whose coordinates satisfy the poset's relations. The paper shows that integer-weight ReLU networks whose decision regions are order polytopes can be viewed as poset neural networks (PNNs). Poset filters. Specific four-element posets generate tropical polynomials equivalent to 2 × 2 convolutional filters. These poset filters can be inserted into ordinary (real-valued) networks to perform feature extraction or pooling without introducing extra trainable parameters. Training behaviour. In image-classification tests, poset pooling achieved finer weight updates than average-, max- or mixed-pooling, yielding accuracy comparable to maxout layers while keeping the model size unchanged. Algebraic composability. The authors endow poset neural networks with an operad algebra: composing networks corresponds to Minkowski sums and convex-envelope operations on their Newton polytopes, providing a combinatorial handle on depth, width and expressivity. Significance. The work suggests that order-theoretic tools, long used in combinatorics and lattice theory, can give interpretable, geometry-aware building blocks for deep learning. Because poset filters involve no extra parameters, they may offer lightweight alternatives for edge devices or for networks where parameter count is critical.
See also: | Order theory | Partially ordered set | Tropical geometry | Convolutional neural network | Pooling layer |
Machine-learning insight for molecular thermodynamics
Predicting thermodynamic data for tens of thousands of candidate molecules is a rate-limiting step in materials and drug discovery. In 2022 the lab of Dr. Mendoza-Cortes and coworkers showed that carefully designed classification experiments can serve as a rapid "scouting" tool before committing to high-accuracy quantum calculations.[21] Methodology. Instead of tackling difficult regression tasks head-on, the authors first cast the problem as a binary classification: is a molecule's enthalpy (or Gibbs free energy, zero-point energy, heat capacity) above or below the median value in the QM9 data set (133 885 small organic molecules)? Descriptor benchmarking. Multiple molecular representations were compared: a high-resolution sorted Coulomb matrix, graph neural-network embeddings and the low-resolution atomic-composition vector (element counts only). When paired with an optimiser tuned for imbalanced classes, atomic composition reached nearly the same classification accuracy as the Coulomb matrix, despite being ~100 × smaller. Transferable insight. The classification pipeline reveals, at negligible cost, which descriptors encode enough physics for a given property, guiding researchers toward efficient surrogate models for later high-precision regression. Practical takeaway. For bulk thermodynamic functions, coarse descriptors may suffice for rapid screening, freeing costly quantum resources for the most promising candidates. The study illustrates how machine-learning design choices, descriptor granularity, loss function, optimiser, translate into physical insight, providing a template for other chemical-property predictions.
See also: | Machine learning | Coulomb's law | Thermodynamic databases | Surrogate model | List of datasets for machine-learning research |
Data-driven machine learning force-fields for 2-D materials
Atomistic simulations are essential for predicting how two-dimensional (2-D) crystals stretch, wrinkle and fracture, but their accuracy depends on the interatomic potential chosen. In 2021 the lab of Dr. Mendoza-Cortes and co-workers introduced a multi-objective, genetic-algorithm workflow that tunes the parameters of widely used analytical potentials so they reproduce large-deformation pathways as well as small-strain properties.[22] Framework highlights: Multi-objective genetic algorithm searches parameter space to minimise simultaneous errors in lattice constants, elastic moduli, phonon spectra and fracture stresses. Correlation & principal-component analysis identify redundant or weakly coupled targets, focusing effort on the most informative training set. Iterative screening adds new targets only when they improve transferability. MoSe2 case study. Four potential families, Buckingham, Stillinger-Weber, Tersoff and modified REBO, were re-fitted; the Tersoff form emerged as the best compromise, capturing both small-strain phonons and ultimate tensile strength within <5 % of density-functional-theory values. Impact. The authors provide a transferable protocol (and open-source code) that other groups can apply to graphene, phosphorene, MoS2 and even bulk solids, accelerating reliable modelling of mechanical, thermal and fracture phenomena across nanotechnology.
See also: | Interatomic potential | Two-dimensional materials | Molybdenum diselenide | Genetic algorithm |
Hessian-informed machine-learning interatomic potentials
In 2025, A. Rodriguez, J. S. Smith (from NVIDIA) and J. L. Mendoza-Cortés published a study exploring whether adding Hessian matrix information (the second derivatives of the potential-energy surface) to the training data of machine-learning interatomic potentials (MLIPs) improves their accuracy and transferability.[23] The authors compared several neural-network MLIPs trained on a benchmark set of small, reactive molecules, supplying different combinations of energy, force and Hessian data. When Hessians were included: Accuracy. Predictions of energies, forces and vibrational spectra for non-equilibrium geometries improved significantly, leading to better description of reaction pathways and transition states. Data efficiency. Comparable accuracy could be achieved with fewer training structures, because the Hessian embeds additional curvature information about the potential-energy surface. Applicability. Improvements translated into more reliable molecular-dynamics trajectories and lower-error nudged-elastic-band (NEB) calculations, both crucial for modelling chemical reactions and materials behaviour. The main drawback was computational cost: evaluating full Hessians during training is substantially more expensive than calculating energies and forces alone. The authors therefore suggested Hessian-augmented training when high fidelity is required (for example, in drug-design, materials discovery or reactive chemistry studies), and force-only training when resources are limited. The work provides practical guidance for researchers choosing between training strategies and highlights the growing role of higher-order quantum-chemical data in next-generation MLIPs.
See also: | Machine learning potential | Potential energy surface | Molecular dynamics | Hessian matrix |
Database of chemical reactions with transition states and Hessians
OpenReACT-CHON-EFH (Open Reaction dataset of atomic ConfiguraTions containing the elements C, H, O and N with reference Energies, Forces and Hessians) is a 2025 open-access benchmark designed for training and stress-testing machine-learning interatomic potentials (MLIPs). It accompanies the Hessian-informed MLIP study by Rodriguez et al.[23] and is the first large-scale collection that provides full second-derivative information for thousands of chemically diverse reactions. The dataset is released under a Creative Commons licence on Figshare.[24] RTP – The core of OpenReACT: each of the 11 961 elementary reactions is represented by its optimised reactant, transition state and product, complete with energy, gradient and Hessian labels. IRC – 600 minimum-energy paths provide a rigorous benchmark for how well a trained potential generalises to unseen regions of the potential-energy surface. NMS – Normal-mode displacements create realistic, thermally perturbed structures for evaluating numerical stability and error growth in molecular-dynamics simulations. Because every geometry carries energies, forces and curvatures, OpenReACT-CHON-EFH enables: 1.benchmarking new MLIP architectures that explicitly learn Hessians; 2.testing force-field extrapolation along complete reaction pathways; and 3. training surrogate models for reactive molecular-dynamics with improved vibrational fidelity.
See also: | Chemical reaction | Reaction mechanism | Quantum chemistry | Dataset | Machine learning potential
Superconductivity
Transition-metal–intercalated bilayer graphene: a blueprint for magnetically mediated superconductivity
Intercalating alkali or alkaline-earth metals between graphene sheets is known to induce superconductivity by donating electrons that hybridise with carbon π-bands. In 2019 the lab of Dr. Mendoza-Cortes and coworkers extended this concept to the entire first-row transition-metal (TM) series, using density-functional theory to predict how each element would affect bilayer graphene when inserted in a "honeycomb" registry between the layers.[25] Electronic structure. For many TMs the carbon π-bands become degenerate and hybridised with partially filled TM d bands, forming a small electron pocket at the K-point, an electronic feature reminiscent of superconducting graphite intercalation compounds. Magnetic exchange. By computing total energies for ferromagnetic and antiferromagnetic alignments, the authors extracted intra-layer exchange couplings, revealing that Ti, V and Cr favour moderate ferromagnetism while Mn and Fe stabilise antiferromagnetic order. Design implication. Because the same TM d states that supply carriers also supply spin excitations, the system could host an unconventional superconducting phase mediated by magnetic fluctuations rather than phonons, all while preserving graphene's Dirac-like dispersion. The study offers a computational roadmap for experimentalists seeking to engineer magneto-electric Dirac materials and explore novel superconducting mechanisms in two-dimensional carbon frameworks.
See also: | Graphene | Graphite intercalation compound | Unconventional superconductivity | Fluctuation–dissipation theorem |
Ultra-fast science
In a 2025 report by the labs of Dr. Cocker and Dr. Mendoza-Cortes reported that few-cycle terahertz (THz) pulses can trigger a reversible, surface-confined topological phase transition in the layered semimetal tungsten ditelluride (WTe2), a material long suspected to host Weyl fermions.[26] Method. Intense THz fields were localised at the apex of a scanning-tunnelling-microscope (STM) tip, resonantly pumping an inter-layer shear phonon. Hybrid-functional density-functional-theory (DFT) calculations confirm that the vibration couples to a polar distortion that toggles the surface between two crystallographic stackings. Sub-atomic imaging. Differential STM images resolved a vertical displacement of 7±3 pm (Picometre), the smallest real-space shift ever captured, setting a new record for spatial resolution in surface science. Electronic consequences. Tunnelling spectroscopy shows the disappearance of characteristic electron- and hole-pocket signatures, consistent with the light-induced annihilation of the surface Fermi-arc states that signify a Weyl-semimetal phase. The ability to switch topological states on and off in a single atomic layer without bulk heating paves the way for ultra-compact devices whose contact resistances approach the quantum limit. Because the work couples record-breaking imaging precision with a striking demonstration of light-controlled quantum matter, it is expected to captivate both specialists and the wider public, offering vivid insight into how photons can sculpt matter on the smallest conceivable length scales.
See also: | Terahertz radiation | Weyl semimetal | Scanning tunnelling microscope | Fermi arc | Phase transition |
Relativistic quantum mechanics
Alleviating nuclear waste
Managing long-lived nuclear-waste streams requires a detailed picture of how the heaviest elements, the actinides, bind to organic chelators. In 2023 the Mendoza-Cortes lab and co-workers developed and deployed state-of-the-art four-component relativistic quantum chemistry to an isostructural series of Pu, Am, Cf and Bk complexes with the redox-active ligand DOPO (2,4,6,8-tetra-tert-butyl-1-oxo-1H-phenoxazin-9-olate).[27] Methodological advance. All-electron Dirac–Hartree–Fock calculations were performed on molecules containing up to 421 electrons and 6 131 basis functions, among the largest actinide systems ever treated with full four-component formalisms. Comparisons with scalar- and spin–orbit-ZORA hybrid-DFT show that lower-cost ZORA models reproduce geometries and frontier-orbital energetics reasonably well, providing a practical screening tool before high-level refinement. f-Orbital contribution. Analysis of natural population and molecular-orbital compositions reveals that 5f character increases from Pu to Bk in the HOMO–LUMO manifold, underscoring the growing role of relativistic stabilisation in later actinides. Actinide–ligand bonding energies. Spin–orbit coupling alters bond strengths by up to 15 kcal mol-1; the study's energy decomposition pinpoints how relativistic contractions of 6d/7s shells compete with 5f participation in σ and π donation. Experimental relevance. Calculated bond lengths for Pu, Am and Bk complexes match available crystallography within 0.02 Å, validating theory where measurements exist and supplying predictions for the uncharacterised Cf analogue. By pushing relativistic electronic-structure theory to previously inaccessible molecule sizes, the work supplies quantitative benchmarks for designing extractants, separations and transmutation strategies in the nuclear-fuel cycle.
See also: | Actinides | Relativistic quantum chemistry | Dirac equation | Nuclear waste | Ligand (chemistry) |
Separation of nuclear waste
One route to reducing the radiotoxicity of Cold-War–era stockpiles is to selectively extract actinide ions (Th, Pa, U) from the complex mixtures stored in ageing tanks. In 2018 Ashley Gannon and colleagues combined a high-throughput virtual-screening algorithm with relativistic density-functional theory (DFT) to design hundreds of candidate chelating agents, organic molecules that wrap around a metal ion and hold it in solution.[28] Screening workflow. Tens of thousands of ligand geometries were generated by systematically "linking" two catecholamide units, well-known actinide binders, and then filtered by machine-learning models trained on smaller DFT data sets. Relativistic accuracy. Promising hits were refined with four-component relativistic DFT, which is essential for heavy elements whose 5f, 6d and 7s electrons experience strong spin–orbit coupling. Secondary-sphere tuning. The study tested the hypothesis that strengthening secondary-coordination-sphere (SCS) interactions, through either covalent linkers or non-covalent hydrogen bonding, would sharpen selectivity. Calculations confirmed that carefully placed linkers can boost binding-energy differences among Th, Pa and U by tens of kJ mol-1. Transferable linkers. Several top-ranking linkers are chemically compatible with other ligand scaffolds, offering a modular path toward bespoke extractants for the wider actinide series. This integrated computational-discovery pipeline illustrates how modern quantum chemistry and data science can accelerate the design of separation agents, complementing experimental efforts to clean up legacy nuclear waste.
See also: | Actinides | Chelation therapy | Nuclear reprocessing | Relativistic quantum chemistry |
Undertanding bonding in later-actinides (heavy elements)
Covalency in the 5f-orbitals of the late actinides (americium onward) remains controversial because few well-defined, air- and moisture-free molecular examples exist. In 2019 a large collaboration of laboratories, including the lab of Dr. Mendoza-Cortes and co-workers prepared the first nonaqueous, isostructural family of tris-chelate complexes that spans both the lanthanides (Ce, Nd, Sm, Gd) and the later actinides (Am, Bk, Cf). Each complex, formulated M(DOPO)3, where DOPO = 2,4,6,8-tetra-tert-butyl-1-oxo-1H-phenoxazin-9-olate, is built from a redox-active dioxophenoxazine ligand bound to a single f-element metal centre.[29] Experimental overview – All seven compounds were crystallographically characterised; the lanthanide members were additionally probed by 1H NMR and SQUID magnetometry. Computational insight – Relativistic CASSCF calculations (validated against UDFT-ZORA geometries) show that, unlike many transition-metal DOPO analogues, the Am, Bk and Cf complexes are predominantly ionic, with f-electrons localised on the metal rather than delocalised onto the ligand π* system. Periodic trends – Systematic comparison across the f-block reveals small but measurable changes in M–O and M–N bond lengths and in optical absorption energies, providing a rare benchmark for theories of 5f covalency. Methodological advance – The work demonstrates that UDFT-ZORA geometries give reliable starting points for multireference (CASSCF) treatments of large actinide systems, a practical route when crystal structures are unavailable. By uniting synthesis, spectroscopy and high-level relativistic calculations, the study delivers a new platform for testing fundamental ideas about f-orbital participation in metal–ligand bonding, knowledge essential for designing actinide separation agents, catalysts and materials.
See also: | Actinides | Lanthanide | Covalent bond | Post–Hartree–Fock |
Relativistic effects on NMR shielding in heavy-atom molecules
A 2023 study by the Mendoza-Cortes lab and the Aucar lab analysed how relativistic quantum mechanics influences the nuclear magnetic-resonance (NMR) shielding constants (σ) of molecules that contain very heavy transition metals, specifically cadmium, platinum and mercury.[30] Background: In light atoms, NMR chemical shifts can often be interpreted with non-relativistic theories. For 5 d and 6 p elements such as Pt and Hg, however, electrons move fast enough that relativistic corrections, spin–orbit coupling, Darwin terms and mass–velocity terms, become comparable to, or larger than, ordinary magnetic shielding contributions. Methodology: The authors employed the Linear Response with Elimination of Small Components (LRESC) framework and introduced a variant that uses localised molecular orbitals (LRESC-Loc). Localisation casts shielding contributions in chemically intuitive fragments; core shells, lone pairs and chemical bonds, rather than delocalised canonical orbitals, helping chemists visualise where relativistic effects originate. Calculations were carried out on: Square-planar and octahedral platinum halides PtX44 and PtX66 (X = F, Cl, Br, I) and Tellurium-bridged complexes CdCl22Te22Y22H66 and HgCl22Te22Y22H66 with Y = N or P. Key findings: Platinum halides. In PtX44 the leading relativistic term is the one-body spin–orbit (σSO(1)SO(1)) contribution. In PtX66, by contrast, mass–velocity and Darwin terms (σMv/DwMv/Dw) overtake spin–orbit effects, altering ligand-dependent trends in NMR shielding across the halogen series. Cadmium and mercury complexes. Replacing nitrogen with phosphorus ligands leaves most relativistic contributions unchanged, except for σSO(1)SO(1). In Cd complexes this term remains positive, whereas in Hg analogues it switches sign because Te–metal bond orbitals dominate the spin–orbit interaction. Chemical insight via localisation. The LRESC-Loc approach partitions shielding into core, lone-pair and bond components, revealing which bonds or shells are chiefly responsible for up- or down-field shifts, information difficult to extract from delocalised orbital analyses. Significance: Understanding the balance of relativistic mechanisms is essential for interpreting heavy-atom NMR spectra in catalysis, materials science and bio-inorganic chemistry. The LRESC-Loc model supplies a chemically transparent tool for predicting and rationalising these effects, complementing experimental spectroscopic studies.
See also: | Nuclear magnetic resonance spectroscopy | Chemical shift | Relativistic quantum chemistry | Spin–orbit coupling |
Spin–orbit coupling and spin-current
Designing next-generation quantum spin Hall (QSH) materials requires an electronic-structure method that treats spin–orbit coupling (SOC) and electron correlation on equal footing. In 2022 the lab of Dr. Mendoza-Cortes and collaborators extended spin-current density-functional theory (SCDFT) to two-dimensional systems and demonstrated its impact on the prototypical bismuth bilayer.[31] Why SCDFT? Standard DFT includes SOC only through one-electron terms; SCDFT adds the three spin-current densities (Jx, Jγ, J_z) to the exchange–correlation potential, capturing many-body effects that arise when moving electrons generate spin flows. Bismuth (001) bilayer under strain. Conventional DFT misplaces the Dirac cone associated with the QSH state. SCDFT predicts a strain-driven s + pz ↔ px + i py band inversion and recovers the Dirac crossing exactly at the Γ-point, in quantitative agreement with first-order k·p perturbation theory. Qualitative improvement. Including spin currents changes the topology of the valence band, not merely shifting energies, highlighting a critical ingredient for accurately modelling SOC-dominated 2D materials. Broader impact. The work opens a pathway to design strain-tunable topological insulators and other SOC-governed quantum materials by providing a practical, relativistic density-functional framework applicable to large-scale simulations.
See also: | Quantum spin Hall effect | Density functional theory | Topological insulator | Spin–orbit interaction |
Quantum mechanics
Quantum-level design rules for hydrogen-storage adsorbents
Physisorption in porous solids is considered the safest, lightest and most energy-efficient route to on-board hydrogen storage, yet the underlying H2–surface interactions are notoriously weak (≈ 4–12 kJ mol-1) and highly quantum-mechanical. In 2020 the lab of Dr. Mendoza-Cortés mapped those interactions with high-level quantum calculations on 240 model complexes that mimic the binding sites found in metal–organic frameworks (MOFs) and covalent-organic frameworks (COFs).[32] Chemical space explored. 12 transition metals (e.g., Sc, Ti, V, Co, Ni, Pd). Multiple oxidation states, spin configurations and chelating ligands. Square-planar, tetrahedral and octahedral geometries representative of MOF secondary-building units. Key findings. Dispersion and electrostatics dominate the binding energy; covalent Kubas-type donation is less common than often assumed. Dozens of sites deliver ≥ 10 kJ mol-1 adsorption, considered the sweet spot for room-temperature storage at moderate pressures (≤ 100 bar). Metals with empty d-orbitals (Sc3+, Ti3+/4+) flanked by π-acidic ligands give the strongest, yet still fully reversible, interactions. Design implications. The study supplies quantitative guidelines for crafting linker chemistry and metal choice in porous frameworks so that each pore wall presents multiple "Goldilocks" sites, pushing materials toward the U.S. DOE targets of 50 g L-1 volumetric and 6.5 wt % gravimetric capacity. By clarifying, at an explicitly quantum level, how H2 binds to realistic catalytic and adsorptive motifs, the work accelerates the rational design of next-generation hydrogen tanks for fuel-cell vehicles and grid storage.
See also: | Dihydrogen complex | Sigma bond | Physisorption | Hydrogen storage | Metal–organic framework |
Sustainability and renewable energy
Artificial photosynthesis
In 2023, the Mendoza-Cortes lab and coworkers created an eleven-member library of titanium-based metal–organic frameworks (MOFs) whose pore size and electronic structure can be tuned independently, allowing a systematic exploration of structure–reactivity trends in photoredox catalysis.[33] Framework design. All members share the formula Ti6O9(linker)3, where the linkers are linear oligo-p-arylene dicarboxylates (n = 1–4 rings). Additional variants incorporate up to 20 mol % amine-functionalised linkers, producing a multivariate (MTV) series. Structural features. Advanced powder X-ray diffraction and total-scattering analyses reveal parallel 1-D Ti6O9 rod "nanowires" packed into a hexagonal net; increasing linker length widens the pore aperture without altering topology. Steric vs. electronic tuning. Steric (pore size). Longer linkers enhance benzyl-alcohol uptake, accelerating photocatalytic oxidation. Electronic (HOMO–LUMO gap). Electron-donating amine groups narrow the optical band gap, boosting visible-light absorption and charge separation. Catalytic performance. The best-performing MTV-MOF oxidises benzyl alcohol nearly 20 × faster than the benchmark MOF MIL-125(Ti) under identical conditions, showing how simultaneous pore-engineering and electronic functionalisation act synergistically. The study demonstrates that decoupling steric and electronic effects in isoreticular MOF series provides a rational route to high-efficiency, earth-abundant photocatalysts for green chemical transformations.
See also: | Metal–organic framework | Photoredox catalysis | Titanium dioxide | Heterogeneous catalysis | Green chemistry |
A tunable photocatalyst for solar water splitting
Birnessite, a naturally occurring, layered manganese oxide, attracts interest as an earth-abundant analogue of the Oxygen-evolving complex in Photosystem II. In 2015, the Mendoza-Cortes Lab performed and developed first-principles calculations to show how intercalating different metal cations between the MnO₂ sheets allows the material’s electronic structure to be “dialled in” for visible-light water splitting.[34] Cation intercalation. Thirteen guest ions (Li⁺, Na⁺, K⁺; Be²⁺, Mg²⁺, Ca²⁺, Sr²⁺, Zn²⁺; B³⁺, Al³⁺, Ga³⁺, Sc³⁺, Y³⁺) were modelled. Jahn–Teller distortions around Mn³⁺ centres shift the valence and conduction bands, narrowing the indirect gap from ~2.6 eV down to ~2.2 eV and the direct gap from ~3.1 eV to ~2.5 eV. Promising compositions. Intercalants such as Sr²⁺, Ca²⁺, B³⁺ and Al³⁺ align the band edges straddling the H₂O/O₂ and H⁺/H₂ redox levels, making the material thermodynamically capable of overall water splitting under sunlight. Anhydrous B-birnessite offers a direct gap ideally positioned for visible-light absorption. Dimensional effect. Isolating a single MnO₂ layer converts the gap from indirect to direct, further enhancing optical absorption—an effect absent in the bulk. The work highlights cation exchange and layer exfoliation as practical levers for turning a common mineral into a low-cost, manganese-based photo-anode that mimics the light-harvesting and catalytic functions of natural photosynthesis.
See also: | Photocatalytic water splitting | Manganese oxide | Band gap | Jahn–Teller effect |
Predicting solar-energy materials for artificial photosynthesis
In 2023, the Mendoza-Cortes lab published a high-throughput strategy for discovering earth-abundant semiconductor photocatalysts suitable for overall water splitting and other solar-fuel reactions.[35] The workflow, dubbed SALSA (Substitution-Approximation evoLutionary Search with Ab-initio calculations), combines three computational filters: 1. Composition generation. An ionic-translation model trained on the Inorganic Crystal Structure Database proposes >30 000 previously unreported ABX and ABC2 semiconductors composed only of non-critical elements (for example Si, Fe, Ti, Ca, N, O, S). 2. Aqueous redox stability. Thermodynamic Pourbaix analysis removes phases that would corrode under the hydrogen-evolution reaction (HER) or oxygen-evolution reaction (OER) conditions used in artificial photosynthesis. 3. Electronic-structure screening. Density-functional-theory calculations obtain accurate band-gap energies and band-edge positions, retaining compounds that straddle the water-splitting potentials while absorbing visible light. From the initial pool the authors isolated dozens of low-cost candidates, including oxynitrides, sulphides and selenides, with predicted light-trapping, redox-stable and catalytically active properties comparable to state-of-the-art but scarce materials such as CdTe or GaInP2. The study demonstrates that data-driven substitution plus ab-initio refinement can accelerate the discovery of photocatalysts for converting sunlight directly into chemical fuels, a key goal of artificial photosynthesis research.
See also: | Artificial photosynthesis | Photocatalytic water splitting | Solar fuel | High-throughput screening |
Hydrogen storage
In 2023 M. Djokic and J. L. Mendoza-Cortes introduced a family of multi-binding-site covalent-organic frameworks (MSUCOFs) designed for room-temperature hydrogen storage and delivery.[36] The work tackles a long-standing obstacle in the hydrogen economy: meeting the US Department of Energy (DOE) ultimate system targets of 50 g L-1 volumetric and 6.5 wt % gravimetric capacity for on-board storage in light-duty vehicles.[37] Design strategy: Quantum-mechanical screening – The authors built 24 virtual COFs by combining three organic linkers with seven transition metals (Co, Cu, Fe, Ni, Mn, Pd and Pt). Grand-canonical Monte-Carlo simulations (0–700 bar, 298 K) based on quantum-derived force fields predicted hydrogen uptake and binding enthalpies for each structure. Multiple binding motifs – By chelating metal centres inside nano-porous channels, MSUCOFs present both strong Kubas-type metal sites and weaker physisorption sites within the same framework, enabling high capacity without cryogenic cooling. Performance highlights: Meets DOE targets at 298 K – Six MSUCOFs built from the abundant first-row metals Co, Ni, Mn and Fe already satisfy the DOE 2025 benchmarks in silico, while Pd- and Pt-based analogues show similar or lower capacities.[36] Cost advantage – The superior performance of first-row metals suggests that economically viable, large-scale hydrogen tanks could rely on readily available elements rather than precious metals. Room-temperature cycling – Simulations predict reversible adsorption–desorption over hundreds of cycles without structural collapse, a key requirement for vehicular refuelling infrastructure. Impact: The MSUCOF concept provides a material-level pathway to reach near-ambient hydrogen-storage targets, complementing high-pressure tanks and cryo-compressed approaches. Because the frameworks can be synthesised from inexpensive building blocks and tuned chemically, they have attracted interest for fuel-cell vehicles, stationary renewable-energy buffering, and portable hydrogen supplies. The paper was featured on the cover of Energy & Fuels, underscoring its significance for sustainable-fuel research.[36]
See also: | Hydrogen storage | Covalent organic framework | Metal–organic framework | Hydrogen economy |
Room-temperature hydrogen storage
Meeting the 2015 US Department of Energy target of ≈40 g L⁻¹ H₂ at 298 K requires sorbents whose binding enthalpy is strong enough (≈15–20 kJ mol⁻¹) to out-perform simple compression yet still release the gas on demand. In 2016, the Mendoza-Cortes Lab proposed a design strategy in which lightweight covalent-organic frameworks (COFs) are functionalised with chelating linkers that bind abundant first-row transition metals (Sc–Cu).[38] Quantum‐mechanical screening. First-principles calculations were implemented to compute H₂ binding enthalpies for 60 metal-linker complexes. First-row metals provided interactions comparable to, or stronger than, precious metals. Porous-material realisation. Thirty COF structures incorporating the best linkers were modelled with QM-based force fields and grand-canonical Monte-Carlo simulations over 0–700 bar. Best performers. Co-, Ni- and Fe-chelated COFs achieved projected working capacities up to ~40 g L⁻¹ at 700 bar and 298 K, double the uptake of unmodified frameworks and above previously reported MOF or COF sorbents. Design guideline. The authors identify an enthalpy “sweet spot’’ and show how pore size and metal spacing create cooperative binding sites, offering a blueprint for Pt-free hydrogen-storage materials. The study was highlighted in MRS Bulletin for demonstrating that cheap, earth-abundant metals can out-perform noble metals in physisorptive hydrogen storage and for mapping an unexplored high-pressure regime relevant to fuel-cell vehicles.[39]
See also: | Hydrogen storage | Covalent organic framework | Monte Carlo method | Transition metal |
Hydrogen production (HER)
Platinum is the benchmark catalyst for the hydrogen-evolution reaction (HER) but is too expensive for large-scale water electrolysis. In 2017, the labs of Prof. Mendoza-Cortes and collaborators reported a centimetre-scale, low-temperature (300 °C) route to stack alternating layers of graphene with alloyed transition-metal dichalcogenides, WxMo1–xS₂, and showed that the resulting vertical heterostructures outperform state-of-the-art WS₂ and MoS₂ catalysts by at least a factor of two.[40] Synthesis and structure. Wet-chemical deposition followed by gentle annealing yields wafer-like films in which graphene sheets alternate with alloy layers of composition W0.4Mo0.6S₂, producing a large density of vertically exposed TMD edges. Electrocatalytic metrics. The best alloy/graphene stack delivers an onset potential of 96 mV at 10 mA cm⁻² and a Tafel slope of 38.7 mV dec⁻¹, values that rival or exceed many noble-metal-free HER catalysts. Activity remains stable after 1 000 cyclic-voltammetry sweeps. Origin of activity. First-principles calculations show that alloying W into MoS₂ lowers the hydrogen-adsorption barrier by optimally overlapping W- and Mo-d orbitals with H-s orbitals; the calculated minimum occurs near the experimentally identified x ≈ 0.4. Graphene provides high conductivity and protects the sulfide layers. Significance. The process is scalable, low-cost, and avoids high-temperature sulfurisation, pointing toward practical, Pt-free electrodes for alkaline and acidic water electrolysers.
See also: | Hydrogen evolution reaction | Graphene | Transition metal dichalcogenide | Heterostructure |
Fuel cells
In October 2024, the labs of Mendoza-Cortes at MSU and B. E. Koel at Princeton reported that electrochemical cycling can transform defect-rich platinum diselenide (DEF-PtSe2) into an ultrastable electrocatalyst for the oxygen reduction reaction (ORR).[41] By subjecting DEF-PtSe2 to 42 000 potential cycles in an O2-saturated electrolyte, the authors triggered a "self-restructuring" process that deposits nanometre-scale Pt clusters onto the two-dimensional lattice and exposes selenium apical sites. The restructured material achieved 1.3 × higher specific activity and 2.6 × higher mass activity than a commercial Pt/C benchmark, and retained most of its performance after 126 000 accelerated-durability cycles. Hybrid density-functional calculations attributed the durability to a synergy between the in-situ-formed Pt nanoparticles and the defective chalcogenide surface, which together optimise the adsorption energies of key ORR intermediates. The study shows that defect engineering and reaction-driven restructuring can unlock high catalytic activity in otherwise semiconducting, van-der-Waals–stacked TM-dichalcogenides. Significance: - Demonstrates a non-carbon 2D support that maintains ORR activity over >105 electrochemical cycles. - Provides a route to lower platinum usage per watt in proton-exchange-membrane fuel cells. - Highlights the emerging role of Pt–chalcogenide chemistry in durable air-electrode design.
See also: | Oxygen reduction reaction | Electrocatalyst | Fuel cell | Transition metal dichalcogenide | Two-dimensional materials |
Water splitting
In 2025, the labs of Dr. Mendoza-Cortés and Dr. Zdilla reported that atomic-scale ordering of metal ions inside two-dimensional potassium nickel–cobalt oxides dramatically improves their performance as oxygen-evolution-reaction (OER) electrocatalysts.[42] Uniform–composition nanosheets are sluggish OER catalysts. Randomly mixed Co/Ni layers show moderate gains, but an alternating stack of pure CoO2 and pure NiO2 sheets cuts the overpotential by more than 400 mV, reaching values competitive with state-of-the-art transition-metal oxides. Mechanistic insights: Hybrid-functional density-functional-theory calculations attribute the enhancement to: Inter-layer potential steps that arise when electronically dissimilar Co and Ni sheets alternate, accelerating hole transport to the active surface. Fermi-level pinning within catalytic CoO2 layers, which stabilises high-valent Co sites crucial for O–O bond formation. Reduced orbital localisation compared with homogeneously mixed solids, lowering kinetic barriers for electron transfer. These findings support the idea that atomic pre-organisation, rather than simply adjusting bulk composition, can be a decisive lever for tuning catalytic activity in layered materials. Broader impact: Layer-by-layer assembly is readily extendable to other transition-metal oxides and redox processes, suggesting a general route to boost electrocatalyst efficiency for water splitting, CO2 reduction and related energy-conversion reactions.
See also: | Oxygen evolution | Electrocatalyst | Layered double hydroxide | Water splitting |
Scalable water-electrolysis technology
Platinum-group metals are the benchmark electrocatalysts for the hydrogen-evolution reaction (HER), but their cost encourages the search for earth-abundant substitutes. In 2018, the Mendoza-Cortes lab and collaborators reported a binder-free, S-doped molybdenum phosphide nanoporous layer (abbreviated S-MoP NPL) that rivals noble metals across the full pH range.[43] Synthesis. A molybdenum foil is anodised to produce a porous oxide, then phosphidised and sulfur-doped by a two-step chemical-vapour-deposition route. The result is a self-supported MoP layer perforated by nano-pores and decorated with sulfur atoms. Electro-catalytic performance. Overpotential of 86 mV at 10 mA cm⁻² and Tafel slope of 34 mV dec⁻¹ in 0.5 M H₂SO₄ (acidic medium). Comparable activity is retained in alkaline (1.0 M KOH) and neutral (phosphate-buffered) electrolytes, making the catalyst “pH-universal”. Long-term chronopotentiometry shows minimal degradation, attributed to the robust nanoporous architecture. Mechanistic insight. Density-functional-theory calculations indicate that sulfur atoms modulate the Mo–P electronic structure, lowering the free-energy barrier for the Volmer–Heyrovsky steps and stabilising Mo–H intermediates; H₂ formation is more favourable on S-MoP than on MoS₂. The study demonstrates that anion-doping can substantially boost the activity and stability of non-precious metal phosphides, suggesting a practical route to scalable water-electrolysis technology.
See also: | Hydrogen evolution reaction | Electrocatalyst | Water electrolysis |
Photocatalytic ammonia production
A 2024 study led by labs of Dr. Mendoza-Cortes, Dr. K.E. Knowles and M.C. Hatzell showed that common nitrogen-containing surface ligands, especially oleylamine, can create false positives in the photocatalytic "fixation" of dinitrogen (N2) on metal-oxide nanocrystals.[44] Ligands that remain bound after nanoparticle synthesis can desorb or decompose under illumination, releasing ammonia that researchers may mistake for genuine nitrogen-reduction activity. Key findings: Ammonia artefacts: Photocatalytic tests on ligand-capped TiO2 and other metal-oxide nanocrystals produced milligram-scale ammonia even when the N2 feed was replaced with argon, proving that the detected nitrogen originated from the ligand, not the gas phase. DFT validation: Hybrid density-functional calculations showed that the valence orbitals of adsorbed amines overlap weakly with the semiconductor band edges, making true N2 activation on these surfaces thermodynamically unlikely, consistent with the experimental artefact. Magnitude of error: If unrecognised, contamination can inflate apparent ammonia yields by two orders of magnitude, undermining claims of efficient "photofixation". Mitigation: The authors recommend nitrogen-free syntheses (for example, using oleic acid or halide precursors) and detailed reporting of washing steps, ligand identities and control experiments under inert atmospheres. Broader impact: The warning extends beyond photocatalytic nitrogen reduction (pNRR) to related reactions such as electro- or photochemical urea synthesis, which also quantifies ammonia as an intermediate. The work has prompted calls for stricter protocols and independent verification to ensure that future catalyst reports are not compromised by hidden nitrogen sources.
See also: | Photocatalysis | Nitrogen fixation | Ammonia production | Ligand (chemistry) | Nanocrystal |
Carbon dioxide reduction
Electro- and photocatalytic CO2 reduction are central to emerging negative-emissions technologies (NETs), yet progress is slowed by the gulf between what can be synthesised in the laboratory and what can be accurately simulated on a computer. In 2022, the lab of Dr. Mendoza-Cortés and collaborators outlined a research roadmap that knits together materials synthesis, multiscale simulation and advanced operando characterisation to speed discovery of NET catalysts.[45] Combinatorial complexity. Heterogeneous CO2-reduction catalysts span variables in composition, nanostructure, support, defect density and electrolyte environment; fully sampling this space experimentally is impractical. Simulation bottlenecks. High-accuracy quantum methods handle ~102 atoms, whereas real catalyst particles contain 104–106 atoms. Machine-learning potentials and coarse-grained continuum models must therefore be coupled seamlessly to bridge length and time scales. Data-centric synthesis. The authors advocate closed-loop workflows in which high-throughput synthesis feeds operando spectroscopy/ microscopy; the resulting data train surrogate models that suggest the next round of experiments, an approach already common in drug discovery but rare in catalysis. Operando insight. Techniques such as ambient-pressure X-ray photoelectron spectroscopy, electrochemical STM and ultrafast optical probes can reveal dynamic catalyst restructuring, helping simulations target physically relevant surface states. The perspective argues that only by integrating these multi-dimensional tools can the field deliver catalysts with the activity, selectivity and durability required for gigaton-scale CO2 removal.
See also: | Negative emissions technology | Photoelectrochemical reduction of carbon dioxide | Electrochemical reduction of carbon dioxide |
Homogeneous selective electrochemical CO2 reduction
Electro-reducing carbon dioxide to carbon monoxide is a key step in closing the carbon cycle, but molecular catalysts often make unwanted H₂. In 2018, the labs of Dr. Mendoza-Cortés and Dr. Head-Gordon performed quantum-mechanical calculations to explain why the tetraaza complex [CoII(N₄H)]²⁺ selectively converts CO₂ to CO with a Faradaic efficiency of ≈45 %.[46] Ligand preference controls selectivity. After two-electron reduction the CoI centre binds CO₂⁻ far more strongly than either H₂O or H⁺, suppressing proton reduction. Proposed catalytic cycle: 1. Two-electron reduction of the precatalyst, 2. CO₂ binds to give [CoIN₄H]⁺–CO₂⁻, 3. Protonation forms a metallocarboxylic acid intermediate, 4. A second CO₂ molecule acts as a Lewis acid, promoting C–O bond cleavage to release CO and bicarbonate while regenerating the CoII species. Why formate is not made. Competing pathways to HCOO⁻ have activation barriers >25 kcal mol⁻¹, whereas the CO-forming route is <15 kcal mol⁻¹, matching the exclusive formation of CO seen experimentally. Testable prediction. The mechanism implies that the turnover frequency should increase linearly with CO₂ partial pressure, an experiment the authors suggest for future verification. The study provides atomistic insight into how CO₂ itself can assist C–O bond breaking, guiding the design of next-generation molecular catalysts for selective carbon-monoxide production.
See also: | Electrochemical reduction of carbon dioxide | Transition metal complex | Carbon dioxide | Carbon capture and storage |
Sustainable polymers
Epoxide ring-opening polymerisation (ROP) offers a route to high-value, recyclable polyethers for adhesives, elastomers and biomedical devices, but the chemistry has long lacked the mechanistic insight. Complementary aluminium-based catalyst platforms developed in the labs of Mendoza-Cortés, Ferrier and Lynd laboratories is redefining how earth-abundant aluminium catalysts can deliver precision control in the ring-opening polymerisation (ROP) of epoxides, reactions that convert three-membered cyclic ethers into high-value polyethers.
Nitrogen–aluminium adduct catalysts (2023)
A 2023 study by labs of Dr. Mendoza-Cortés and Dr. Ferrier Jr. clarified how nitrogen–aluminium (N-Al) adducts catalyse the ring-opening polymerisation (ROP) of epoxides, reactions that convert simple, three-membered cyclic ethers into high-value polyethers used in adhesives, elastomers and biomedical devices.[47] Background: Unlike vinyl monomers, where RAFT and related "living" techniques give predictable molecular weights, epoxide ROP lacks a single, general platform. Aluminum compounds are attractive because aluminum is earth-abundant, inexpensive and non-toxic, but their catalytic mechanisms have remained poorly defined. Mechanistic findings: Adduct formation. Spectroscopic data show that dialkyl- or alkoxide-aluminum precursors form reversible N-Al adducts with amine initiators; density-functional-theory (DFT) calculations locate a stable five-coordinate aluminium centre that activates the epoxide ring. Chain-growth control. Kinetic studies reveal first-order dependence on monomer and catalyst, with tunable rates (10-3–10-1 s-1) by varying the Lewis basicity of the nitrogen ligand. Narrow dispersities (Đ ≈ 1.1–1.3) and linear Mn growth demonstrate living/controlled behaviour analogous to RAFT. End-group fidelity and architecture. The platform supports block, star and gradient polyethers through sequential monomer feeds and multifunctional amines, while maintaining predictable molecular weights up to 100 kDa. DFT reaction profile. Computations map the full catalytic cycle-epoxide coordination, ring opening, alkoxide propagation and catalyst regeneration, and rationalise why N-substituent electronics modulate both rate and selectivity. Significance: The combined experimental–theoretical picture establishes N-Al adducts as a general, customisable epoxide-polymerisation toolkit, comparable in scope to RAFT for vinyls or ROMP for cyclic olefins. The insight guides future catalyst design aimed at sustainable, aluminium-based routes to advanced polyethers without relying on precious metals.
Classical Vandenberg catalyst revisited (2018)
For half a century chemists have used the aluminum “Vandenberg catalyst’’ to make isotactic polyethers, but the detailed reason for its regio- and isoselectivity was unclear. In 2018, the Mendoza-Cortés and Lynd laboratories combined spectroscopy, kinetics and density-functional theory (DFT) to unravel the mechanism.[48] Resting and transition states. Quantum mechanical calculations shows the active species is a rigid bis(μ-oxo) dialuminium (BOD) complex that inserts epoxide through a mono(μ-oxo) dialuminium (MOD) transition state. Origin of isoselectivity. The BOD framework simultaneously coordinates the ultimate and penultimate oxygen atoms of the growing chain; inserting an epoxide of opposite configuration would break a favourable secondary Al–O interaction, imposing a ~2 kcal mol⁻¹ penalty that biases the reaction toward isotactic enchainment. Experimental support. Model BOD and MOD complexes reproduce the predicted NMR and IR signatures. A synthetically prepared BOD analogue polymerises allyl-glycidyl ether to give the same isotactic-rich polyether as the classical catalyst. In-situ ¹H-NMR during polymerisation yields an activation enthalpy ΔH‡ ≈ 21 kcal mol⁻¹ and an exothermic epoxide-binding step (ΔH ≈ –4 kcal mol⁻¹), matching the computed profile. The work replaces earlier speculative schemes with a quantitative picture, guiding the design of next-generation aluminum catalysts for stereospecific polyether synthesis.
See also: | Ring-opening polymerization | Polyether | Living polymerization | Aluminium compounds | Isotactic polymer |
Future batteries
High-voltage lithium batteries
Conventional ether-based polymer electrolytes are attractive for flexible or solid-state lithium batteries, but they degrade at both extremes of the cell: Anode side – low potentials trigger uncontrolled anionic polymerisation, thickening the interphase and raising impedance. Cathode side – oxidative breakdown begins well below the >4 V potentials needed for next-generation high-energy cathodes. In 2019, the lab of Dr. Mendoza-Cortés and Dr. Archer (at Cornell) reported two complementary fixes that push ether electrolytes into the high-voltage regime.[49] Cationic chain-transfer agents (CTAs). Added to the electrolyte, CTAs intercept growing anionic chains at the lithium metal surface, suppressing runaway polymerisation and stabilising the solid-electrolyte interphase. Designer cathode-electrolyte interphases (CEIs). Pre-formed films of anionic polymers and supramolecular additives coat high-voltage cathodes, blocking oxidative attack on the ether matrix up to ≈5 V. Significance: Cells employing both strategies cycle stably with nickel-rich layered oxides (NMC811) at 4.5 V, demonstrating that ether-based polymer electrolytes can operate well above previously accepted voltage limits. The work provides a general blueprint for chemically engineering both anode and cathode interphases to unlock safer, high-energy solid-state batteries.
See also: | Solid-state battery | Lithium metal battery |
Sodium-ion batteries
Graphite, the standard anode in lithium-ion cells, does not readily host sodium ions because Na+–graphite intercalation is thermodynamically unstable. In 2022 Dipobrato Sarbapalli and co-workers demonstrated that a one-side fluorinated few-layer graphene (F-FLG) film overcomes this limitation, enabling reversible Na+ intercalation at room temperature.[50] Experimental evidence. Cyclic voltammetry showed well-defined intercalation/de-intercalation peaks corresponding to stoichiometries near NaC14–C18, far higher than the ~NaC186 limits reported for pristine graphite. In-situ Raman spectroscopy tracked shifts in the G- and 2D-bands, confirming staging behaviour as Na+ entered and left the graphene galleries. Ion-sensitive scanning electrochemical microscopy provided direct spatial maps of Na+ flux into the F-FLG electrode. Mechanistic insight. Fluorination stabilises Na+ between graphene layers, while a pre-formed Li-based solid-electrolyte interphase (SEI) lowers kinetic barriers, highlighting the interplay between surface chemistry and interphase engineering. Impact. The work suggests that targeted surface modifications, rather than wholesale changes to carbon architecture, can adapt existing graphitic anodes for sodium-ion battery chemistries, an attractive route given sodium's abundance.
See also: | Sodium-ion battery | solid-electrolyte interphase | Graphene |
Potassium-ion batteries
Lithium-ion technology struggles to scale for stationary storage because of cost and resource limits. One promising alternative is the potassium-ion battery (KIB), but K⁺ ions intercalate sluggishly into graphitic anodes and tend to plate on the surface, causing poor rate capability and rapid failure. In 2018, the Mendoza-Cortes lab in collaboration with the Rodriguez-Lopez lab demonstrated that an ultrathin few-layer graphene (FLG) electrode can be “pre-conditioned’’ in a Li⁺ electrolyte so that the resulting Li⁺-derived solid-electrolyte interphase (SEI) acts as a gatekeeper for fast, reversible K⁺ intercalation.[51] Electrochemical performance. The Li-SEI-modified FLG exhibits discrete staging peaks in cyclic voltammetry even at scan rates up to 100 mV s⁻¹, confirming true intercalation rather than potassium plating. Charge–discharge tests reach ~360 C (full discharge in 10 s) and retain capacity for 1 000 cycles at 10 C. Spectroscopic confirmation. In situ Raman spectroscopy tracks the G- and 2D-band shifts characteristic of K⁺ staging, while time-of-flight secondary-ion mass spectrometry shows that the Li-based SEI is permeable to K⁺. Mechanistic insight. Density-functional calculations indicate that the Li-rich interphase lowers the kinetic barrier for K⁺ insertion and suppresses dendritic deposition, explaining the high rate capability and cycling stability. The study introduces a simple SEI-engineering strategy, conditioning carbon electrodes in Li⁺ before K⁺ use, that could accelerate the development of low-cost, high-power potassium-ion batteries for grid and fast-charging applications.
See also: | Potassium-ion battery | Solid-Electrolyte interphase | Graphene | Intercalation (chemistry) |
Lithium/Potassium Batteries (Co-intercalation)
Lithium-ion intercalation into graphite underpins most commercial batteries, yet its fundamental thermodynamics had never been shown to follow a textbook Nernst equation. In 2021 the labs of Dr. Rodríguez-López and Dr. Mendoza-Cortés used micrometre-grain few-layer graphene (FLG) as an electroanalytical platform to demonstrate that Li+ staging indeed exhibits a near-Nernstian slope (≈ 55 mV decade-1) and to probe, for the first time, the co-intercalation of Li+ and K+ into the same carbon galleries.[52] Electrochemical signatures. Cyclic voltammetry of FLG in Li+ electrolytes shows staging peaks that shift linearly with log [Li+], mirroring the Nernst prediction for a one-electron process. Co-intercalation behaviour. Introducing K+ alters the peak positions with shallower slopes (~30 mV decade-1), indicating competitive occupancy of graphene layers by both cations. Diffusion kinetics. Potentiostatic intermittent titration (PITT) finds that alkali-ion diffusion coefficients in FLG are composition-dependent, falling as K+ content rises. Theoretical support. Density-functional-theory calculations reveal that Li+/K+ mixtures can form thermodynamically favourable domains within a single graphene layer; binding energy increases with Li-rich areas and decreases as –K–Li– ordering grows. The study positions ultrathin graphitic electrodes as versatile probes for multi-ion intercalation thermodynamics, informing the design of dual-alkali and sodium–potassium batteries as well as next-generation graphite alternatives.
See also | Lithium-ion battery | Graphene | Intercalation (chemistry) | Nernst equation |
Beyond lithium-ion batteries
Efforts to develop rechargeable batteries beyond the dominant lithium-ion technology have increasingly focused on more abundant elements such as sodium, potassium, magnesium and calcium. In 2024, Yu-Hsiu Lin and José L. Mendoza-Cortés reported a density-functional-theory (DFT) study that maps out how these ions insert, or intercalate, between layers of graphene that have been selectively modified on one side with fluorine or related surface groups.[53] Key results: Staging mechanism: The authors analysed the first three "stages" of intercalation, the progressive filling of distinct gallery spaces between graphene sheets, and identified a previously unrecognised binding site that is thermodynamically favoured for all six ions studied (Li, Na, K, Be, Mg, Ca). Surface modifiers (SMs): One-side fluorination lowers the energy cost of inserting large Na+ and K+ ions, suggesting a strategy for high-capacity sodium-ion and potassium-ion batteries. Layer dependence: The calculated binding energy varies strongly with the number of graphene layers: beyond a threshold thickness the gain from additional layers saturates, defining an optimal electrode architecture. Electronic properties: Hybrid-functional DFT combined with projected density-of-states (PDOS) and Mulliken-population analysis revealed charge transfer pathways among the ion, the surface modifier and the graphene host, clarifying why certain ions (for example Ca2+) are electronically incompatible with specific staging configurations. Significance: The study provides design rules for few-layer graphene (FLG) electrodes that could accommodate both monovalent and divalent ions, widening the palette of elements usable in next-generation batteries. It also shows how staging intercalation, well known in graphite–lithium chemistry, extends to heavier or multivalent ions when surface chemistry and layer count are tuned together.
See also: | Sodium-ion battery | Potassium-ion battery | Multivalent battery | Graphene | Intercalation (chemistry) |
Surface-engineered graphene anodes for Li+, Na+ and K+ batteries
Graphite is the standard negative-electrode material in today's lithium-ion batteries, yet it fails to store sodium (Na+) or potassium (K+) efficiently at room temperature. In 2020 the lab of Dr. Mendoza-Cortes and co-workers used first-principles calculations to show that atomically thin, surface-modified few-layer graphene (FLG) can overcome that limitation.[54] Interface design. Three strategies were modelled: 1. Covalent doping of the outermost graphene layer with hetero-atoms. 2. Laminating FLG with single-side fluorinated graphene. 3. Laminating FLG with a Janus hydrofluorographene sheet (H-F asymmetry). Thermodynamic outcome. All modifications shift the Fermi level and introduce charge-transfer interactions that stabilise the insertion of Li+, Na+ and K+ between graphene layers; for Na+ the process becomes exothermic, enabling reversible storage that pristine graphite cannot achieve. Capacity gain. Calculated specific capacities surpass those of conventional graphite, indicating head-room for higher-energy sodium- and potassium-ion batteries using carbonaceous hosts. Electronic rationale. Density-of-states and charge-density analyses trace the improved affinity to mid-gap states created by dopants or by the electrostatic dipole of the fluorinated overlayer. The results position surface chemistry, rather than bulk structural change, as a key lever for adapting graphene-based anodes to multivalent or heavier alkali-ion chemistries, broadening the palette of post-lithium energy-storage technologies.
See also | Lithium-ion battery | Sodium-ion battery | Potassium-ion battery | Graphene | Intercalation (chemistry)
Solid electrolytes
Solid materials are usually thought of as rigid, yet in an amphidynamic crystal certain parts of the lattice can move rapidly while the backbone remains fixed. In 2018, the Mendoza-Cortes lab and co-workers showed how to build this behaviour into a porous covalent-organic framework (COF) by decorating its pores with short oligo-(ethylene oxide) (OEO) chains of different lengths.[55] Preserved crystallinity. Because COFs assemble through strong covalent bonds, adding flexible OEO chains does not disrupt the long-range order; the resulting materials keep the predictable honey-comb topology characteristic of 2-D COFs. Two kinds of motion. Framework rotors – Phenylene rings built into the backbone undergo hindered 180 ° flips, but only slowly at room temperature. Side-chain rotors – The attached OEO chains spin almost freely inside the one-nanometre pores, an example of fast “molecular-gear” motion within an otherwise rigid crystal. How it was measured. The team used ¹³C solid-state NMR relaxation to quantify rotation rates and activation energies, and quantum-mechanical calculations to visualise the atomistic motions. Why it matters. Combining a sturdy scaffold with internal moving parts opens the door to COF-based solid electrolytes, adaptive sorbents and molecular machines whose dynamics can be tuned simply by changing side-chain length.
See also: | Covalent organic framework | Molecular machine | Solid-state nuclear magnetic resonance | Poly(ethylene oxide) |
Solid polymer electrolytes
Solid polymer electrolytes (SPEs) that conduct only Li+ ions are attractive for high-energy, non-flammable solid-state batteries, but most candidates suffer from low room-temperature conductivity. In 2021 the lab of Dr. Mendoza-Cortes and collaborators reported the first crystalline, anionic helical covalent polymer (HCP) whose backbone folds into densely packed double helices, creating one-directional, negatively charged nanochannels that transport lithium ions without added salt.[56] Structure. Single-crystal X-ray diffraction reveals a double-helix arrangement; the aligned anionic channels act as built-in ion pathways. Electrolyte performance. Li+ conductivity: 1.2 × 10-3 S cm-1 at 25 °C (without external Li salt). Activation energy: 0.14 eV. Electrochemical stability window: 0.2 – 5 V (vs Li/Li+). Ionic-liquid enhancement. Swelling the polymer with a non-volatile ionic liquid boosts conductivity by > 1,000× while retaining non-flammability. Battery test. An all-solid-state cell with an NMC-811 cathode cycled reversibly, demonstrating practical viability. The authors propose that the "helical-channel" design paradigm could be generalised to other covalent frameworks, opening a new route toward high-performance, single-ion conducting solid-state electrolytes.
See also: | Solid-state battery | Polymer electrolyte | Lithium-ion battery |
Zinc–Air batteries
Rechargeable zinc–air batteries (ZABs) need bifunctional electrocatalysts that drive both the oxygen-reduction reaction (ORR) during discharge and the oxygen-evolution reaction (OER) during recharge. In 2018, the Mendoza-Cortés and colleagues reported an “apically dominant’’ strategy that lifts the activity of nitrogen-doped carbon-nanotube (NCNT) arrays well above that of conventional base-grown tubes or even platinum-group-metal benchmarks.[57] Design concept. Rather than grow nanotubes from their bases, the team encapsulated CoNi nanoparticles inside the tips (apical domain) of vertically aligned NCNTs on nickel foam (denoted CoNi@NCNT/NF). Electrocatalytic performance. The apically doped electrode shows: 1. ORR half-wave potential and OER over-potential superior to commercial Pt/C + IrO₂ catalysts. 2. ZAB coin cell with peak power density ≈ 127 mW cm⁻², energy density ≈ 845 Wh kgZn⁻¹ and stable cycling for > 90 h. Mechanism. Quantum mechanical simulations indicate that the highest ORR activity arises from the synergy between graphitic-N sites in the NCNT walls and newly created apical active sites adjacent to the embedded CoNi particles. The work demonstrates how precise placement of transition-metal nanoparticles inside carbon nanotubes can overcome the hydrophobicity and limited active-site exposure of base-grown arrays, offering a scalable route to low-cost, high-performance air electrodes.
See also: | Zinc–air battery | Oxygen reduction reaction | Oxygen evolution | Carbon nanotube |
Next generation materials
2D-materials
In 2023 a consortium led by several lab, including the lab of Prof. Mendoza-Cortes, published an extensivestudy that maps the fast-moving landscape of two-dimensional (2D) materials, spanning fundamental theory to prototype devices.[58] The article synthesises progress across six interconnected fronts: Theory and modelling. Defect and intercalant formation pathways are described with first-principles and machine-learning (ML) potentials, revealing how atomic-scale disorder can be harnessed for catalysis, sensing and spin–orbit engineering. Machine-learning synthesis. Data-driven retrosynthesis guides precursor choice, furnace profiles and in-situ metrology, accelerating the growth of high-quality graphene, transition-metal dichalcogenides (TMDCs) and MXenes. Emergent 2D families. Advances include room-temperature 2D magnets, epitaxial low-symmetry crystals (e.g. ReS2) and oxidation/strain-gradient engineering that tailors band gaps and exciton funnels. Optical & phonon phenomena. Spatial inhomogeneity, probed via multidimensional Raman/photoluminescence imaging, enables tunable valley polarisation and ultra-low-loss phonon waveguides; ML algorithms extract hidden correlations from hyperspectral data cubes. Mixed-dimensional heterostructures. Stacking 2D layers with 0D quantum dots or 3D ferroelectrics yields reconfigurable logic/memory elements, while high-quality magnetic topological-insulator films show robust quantum anomalous Hall states. Twistronics & quantum transport. Small-angle homojunctions exhibit flat-band–driven superconductivity and correlated insulators beyond graphene, expanding the twistronics paradigm to TMDCs and black phosphorus. Perspectives sections highlight open challenges in scalable synthesis, interfacial heat management and quantum-grade defect control, charting a roadmap toward industrial and quantum-technology deployment of 2D materials.
See also: | Two-dimensional materials | MXene | Twistronics | Quantum anomalous Hall effect | Applications of artificial intelligence |
Twistronics: Twist-engineered transition-metal dichalcogenide bilayers
Twistronics, the control of electronic properties through the relative twist of adjacent two-dimensional (2D) layers, was first popularised by the discovery of superconductivity in magic-angle graphene bilayers in 2018. Similar ideas are now being applied to semiconducting transition-metal dichalcogenides (TMDCs), whose rich chemistry offers additional "knobs" such as chalcogen substitution and heterostructure design. In 2025, the Mendoza-Cortes lab reported a systematic first-principles study of how twisting and stacking affect the electronic structure of six MX2 TMDCs (M = Mo, W; X = S, Se, Te).[59] Using range-separated hybrid density-functional theory they examined 30 bilayer and heterobilayer combinations under both relaxed and low-strain conditions. Twist tuning: Critical angles generate symmetric Moiré patterns that flatten bands and convert indirect-gap semiconductors into direct-gap materials. Stacking registry: Bernal-like (60°) shifts can stabilise heterostructures and lower the gap, as demonstrated for MoTe2/WSe2. Strain sensitivity: Allowing full relaxation often restores the indirect gap, whereas low-strain constraints preserve twist-induced direct gaps, highlighting the interplay between lattice relaxation and electronic structure. Device relevance: The ability to toggle between metallic, indirect-gap and direct-gap states suggests routes to valleytronic devices, broadband photodetectors and flexible optoelectronics based on a single material platform. The authors conclude that combining heterostructure chemistry with controlled twist angles provides a versatile toolkit for tailoring band-edge positions, gap character and band-flatness, key parameters for next-generation 2D electronics and photonics.
See also: | Twistronics | Transition metal dichalcogenide | Moiré pattern | Two-dimensional materials | Density functional theory |
2-D electronic devices
Efficient organic and 2-D electronic devices require materials whose molecular geometries change very little when they gain or lose an electron; that is, they have small internal reorganisation energies (λ). In 2018, the Mendoza-Cortes lab and collaborators performed quantum mechanical calculations to show that rectangular, hydrogen-passivated silicene nanoclusters (abbreviated H-SiNCs) can achieve exceptionally low λ values while retaining large electron affinities, making them promising n-type conductors and spintronic building blocks.[60] Key findings. Internal reorganisation energies for electron transport fall below 50 meV for suitably sized clusters, comparable to or better than many high-mobility organic semiconductors. All studied H-SiNCs exhibit large electron affinities (stabilised anionic states), indicating intrinsic n-type behaviour. Tuning the cluster dimensions along the zig-zag and arm-chair lattice directions independently provides a handle for optimising both λ and frontier-orbital energies. Implications. The combination of low λ and high electron affinity suggests H-SiNCs could serve as contact layers, channel materials or spin filters in future silicon-compatible nano-electronics and 2-D spintronic devices.
See also: | Silicene | Reorganization energy | n-type semiconductor | Spintronics |
2D-Heterostructures as “electronically transparent’’ electrocatalysts
Two-dimensional graphene is impermeable to molecules yet only one atom thick, raising the question of whether an electrochemical reaction can be “felt’’ through it. In 2018, the labs of Dr. Mendoza-Cortes and Dr. Rodriguez-Lopez built model graphene–metal–molecule heterostructures and showed that a buried metal film can indeed modulate catalysis occurring on the opposite side of the graphene sheet.[61] Electrode architecture. Monolayer graphene was transferred onto patterned underlayers of Pt, Au or SiOx; selected regions were then coated with cobalt-porphyrin molecules that catalyse the oxygen-reduction reaction (ORR). Key observations. Porphyrin adlayers shift the ORR towards the 2-electron (H₂O₂) pathway, whereas bare graphene favours the slower 4-electron route. Despite being sealed beneath an impermeable carbon sheet, the metal underlayer lowers the ORR over-potential in the order Pt > Au > SiOx, indicating that d-electrons “leak’’ through graphene and interact with adsorbed reactants. The encapsulated metals are protected from cyanide poisoning and long-term cycling, confirming graphene’s role as a corrosion barrier that remains electronically transparent. Computational insight. Quantum mechanical calculations attribute the activity boost to tunnelling of metal d-orbitals through graphene, which stabilises reaction intermediates on the outer surface. The work establishes graphene-based heterostructures as a platform for decoupling chemical accessibility from electronic influence, with implications for durable fuel-cell cathodes, protective coatings and the fundamental study of two-dimensional materials.
See also: | Graphene | Oxygen reduction reaction | Electrocatalyst | Heterostructure |
Spintronics in 2D-bilayers
In 2023, the Mendoza-Cortes lab reported a computational study showing that thethe Mendoza-Cortes lab behaviour of two-dimensional covalent-organic frameworks (COFs) can be tuned over a wide range by inserting first-row transition-metal atoms between adjacent layers.[62] Framework design: The authors modelled a bilayer COF built from benzene, boroxine (B3O3) and triazine (C3N3) rings; motifs common in experimentally synthesised porous frameworks. Sixty-three distinct structures were generated by placing single or multiple 3d transition-metal ions (Sc–Zn) at different positions between the two layers; a pristine (metal-free) bilayer and a trilayer reference were also examined. Hybrid density-functional-theory calculations probed band structures, density of states and spin polarisation for each configuration. Key findings: Electronic tunability. Depending on the metal species, concentration and registry, the intercalated COFs span behaviour from wide-gap semiconductors to narrow-gap direct semiconductors in the visible range. Spintronics potential. Several Ni-, Co- and Mn-containing bilayers act as half-metals or bipolar magnetic semiconductors, materials that conduct electrons of one spin orientation while rejecting the opposite spin, desirable for spin-filter or spin-valve devices. Catalytic relevance. Structures exhibiting spin-polarised surface states near the Fermi level are predicted to be active for electrocatalytic water splitting, echoing trends seen in transition-metal oxides. The work highlights the versatility of layered COFs, traditionally valued for high porosity and low density, as a platform for next-generation organic spintronics and photo-electrocatalysis once their interlayer chemistry is engineered at the atomic scale.
See also: | Covalent organic framework | Two-dimensional materials | Spintronics | Transition metal |
Recreate (or switch off) the Dirac cone
Bilayer graphene (BLG) is normally a small-gap semiconductor, but its band structure can be reshaped by stacking order, electric fields or chemical doping. In 2018, the Mendoza-Cortés perfomed Quantum-mechanical calculations to show that sandwiching individual vanadium-group transition-metal atoms between the two graphene sheets can recreate (or switch off) the massless-fermion Dirac cone that makes monolayer graphene famous.[63] Key result. A single vanadium atom placed in the hexagonal hollow site (BLG-1V) preserves a Dirac crossing at the K-point, turning the bilayer back into a zero-gap semimetal. Changing the metal (Nb, Ta), the spin alignment or the intercalant concentration converts the same scaffold into a metal or a conventional semiconductor. Electronic mechanism. The cone originates from hybridisation between graphene 2pz orbitals and the metal 3dyz (or 4d/5d) orbitals, demonstrating that an out-of-plane dopant can control in-plane π-electrons without disrupting the lattice. Implication. Transition-metal intercalation offers a chemical “knob’’ for switching and tuning Dirac physics in BLG, with potential applications in valleytronics, spin-valves and reconfigurable 2-D electronics.
See also: | Graphene | Dirac cone | Intercalation (chemistry) | Two-dimensional materials |
Tunning and reshaping the Dirac cone
Because pristine bilayer graphene (BLG) is only a small-gap semiconductor, methods that shift or reshape its Dirac cone are central to future carbon-based electronics. In 2018, the Mendoza-Cortés lab performed quantum-mechanical calculations to screen twenty first-row transition-metal (TM) intercalates, ten atomically thin BLG structures and ten corresponding “bulk’’ (multilayer) analogues, and mapped how each metal controls the electronic landscape of graphene.[64] Material design. BLG was intercalated with every 3d element from Sc to Zn. Equilibrium geometries remain layered, with metals residing between the carbon sheets. Electronic trends. Hybridisation between graphene 2pz orbitals and TM 3dyz states generates the bands nearest the Fermi level. Depending on the metal and its concentration the systems span metallic, semi-metallic (Dirac) and semiconducting behaviour. Several intercalates (e.g., Ti, V, Mn) shift the Dirac crossing back to the K-point, effectively restoring massless carriers; others open or widen a gap, offering a chemical “dial’’ for device engineering. Dimensional comparison. Layered and bulk analogues show similar orbital interactions, indicating that the tuning strategy is scalable from two-dimensional films to three-dimensional graphitic solids. The study provides a computational roadmap for chemists seeking to build graphene-based spin valves, transistors and sensors in which electronic properties are set by the choice of an easily inserted transition-metal layer.
See also: | Fermion | Massless particle | Transition metal | Bilayer graphene |
Semiconductors with crystalline porosity
Covalent-organic frameworks (COFs) are crystalline, permanently porous solids built from light elements and robust covalent bonds, but their largely insulating backbones restrict use in electronics. In 2019 the Mendoza-Cortés lab proposed a general remedy: slip first-row transition-metal (TM) atoms into the pores to create electronically tunable "TM-intercalated COFs." Using density-functional theory they modelled 31 new materials, COF–TM-x (TM = Sc–Zn; x = 3–5 atoms per unit cell), based on boroxine- and triazine-linked frameworks.[65] Band-gap engineering. The calculated gaps span wide-gap semiconductors to narrow-gap (≈0.2 eV) materials, governed by the d-electron count of the intercalant and its concentration. Density of states. Sc-, Ti- and V-intercalated COFs introduce mid-gap states that lower the Fermi level toward the conduction band, while late TMs (Ni, Cu, Zn) leave the parent gap largely intact. Structure preservation. All intercalated models retain the parent COF topology, suggesting that high surface area and molecular sieving can coexist with tailored electronic behaviour. The study points to a new design axis, ion intercalation, for transforming insulating COFs into porous semiconductors suited to sensors, thin-film transistors or catalytic electrodes.
See also: | Covalent organic framework | Intercalation (chemistry) | Electronic band structure | Transition metal |
Programmable semiconductors
In 2017, the Mendoza-Cortes lab showed, through first-principles calculations, that slipping single iron atoms between the layers of a boroxine- or triazine-linked COF converts the normally insulating framework into a family of semiconductors with tunable band gaps.[66] Design strategy. Seven new structures were modelled by inserting Fe atoms at well-defined sites between adjacent COF sheets while preserving the parent hexagonal symmetry. Electronic tuning. Hybrid DFT predicts band-gap openings in the 0.4–1.2 eV range, depending on Fe concentration; the gaps arise from hybridisation between graphene-like 2pz orbitals of the COF backbone and Fe 3d states near the Fermi level. Semiconductor potential. Because the intercalants leave the rigid pore lattice intact, the authors propose the materials as porous semiconductors whose charge-transport and sorption properties can be adjusted independently, useful for sensors, catalysis and low-k dielectric films. The study establishes transition-metal intercalation as a general route to electronic functionality in COFs, complementing earlier efforts that relied on linker substitution or post-synthetic modification.
See also: | First-principles in physics | Covalent organic framework | Semiconductor | Density functional theory |
Molecular machines
In 2023, the lab of Prof. Mendoza-Cortés combined ab-initio molecular-dynamics (AIMD) with quantum–chemical analyses to uncover how a prototypical rotaxane, a mechanically interlocked molecule in which a macrocyclic "ring" encircles a dumbbell-shaped axle, responds to heat and light.[67] Atomistic simulations. AIMD trajectories (1.2 ps each) were run from 300 K up to 2 500 K using the atom-centred density-matrix-propagation method, revealing how hydrogen bonding, π–π stacking and Coulomb forces maintain structural integrity far above room temperature. Non-covalent mapping. Reduced-density-gradient (RDG) and density-overlap-region (DORI) plots pinpoint stabilising contacts between the axle's stopper groups and the macrocycle, while Hirshfeld surfaces quantify ring–axle dispersive interactions. Thermo-mechanical stability. A "thermal-fluctuation index" extracted from the trajectories shows that the rotaxane withstands large conformational excursions without dethreading even at ~900 K, highlighting its potential as a high-temperature molecular shuttle. Optical performance. Time-dependent DFT predicts a broad UV–visible absorption band; calculated first- and second-order hyperpolarizabilities indicate strong nonlinear-optical (NLO) responses, positioning the molecule as a candidate for photonic or solar-energy-harvesting applications. The study demonstrates how first-principles dynamics can guide the design of mechanically interlocked molecules with tailored thermal robustness and opto-electronic functionality, key attributes for molecular machines, data storage and organic NLO devices.
See also: | Rotaxane | Car–Parrinello molecular dynamics | Nonlinear optics | Molecular machine | Solar energy |
Conversion of heat into mechanical energy
Organic crystals rarely display fast, repeatable shape changes when heated and cooled, limiting their use as solid-state "muscles" in soft robotics or self-resetting fuses. In 2021 the lab of Dr. Mendoza-Cortes and co-workers described a molecular crystal that overcomes those hurdles: on heating above 45 °C it contracts to about 90 % of its length in a fraction of a second, and on cooling below 35 °C it instantaneously re-extends to its original size.[68] Thermo-elastic performance. The crystal delivers the same stroke for >200 heating–cooling cycles with negligible fatigue and remains stable for years at room temperature. Phase transition. X-ray diffraction reveals a single-crystal-to-single-crystal transformation that preserves long-range order, allowing rapid, repeatable motion with almost no hysteresis. Applications. Demonstrations include a resettable thermal fuse that switches an electrical circuit on and off in real time, illustrating potential for temperature-triggered actuators in micro-devices or safety systems. The work positions organic molecular crystals as practical, room-temperature converters of heat into mechanical work, complementing inorganic shape-memory alloys and polymer actuators.
See also: | Shape-memory polymer | Soft robotics | Actuator |
Designing defects of materials
Activating "inert" materials
Pristine hexagonal boron nitride (h-BN) is celebrated as a chemically inert, electrically insulating, two-dimensional crystal. That inertness, however, limits its usefulness as a catalyst support or functional surface. In 2021 the lab of Dr. Mendoza-Cortes and Dr. Terrones showed that room-temperature "inert" h-BN can be converted into a highly reactive platform for single-atom catalysis by a simple cryomilling step carried out in liquid nitrogen.[69] Defects designs by cryomilling. High-energy ball-milling at –196 °C introduces B- or N-vacancies (with and without oxygen decoration) to give "defective BN" (d-BN). These vacancies generate mid-gap electronic states, verified by photoluminescence and electron-spin-resonance spectroscopy. Spontaneous metal loading. The defect sites act as localized radicals that can reduce dissolved metal cations at room temperature, anchoring isolated single metal atoms (for example Pt, Pd) or small bimetallic clusters without an external reducing agent. Catalytic performance. Surface-bound single atoms on d-BN show markedly enhanced activity toward the hydrogen-evolution reaction, pointing to broader applications in energy conversion, sensing and quantum-information devices. Theoretical support. Density-functional-theory calculations reproduce the vacancy electronic structures, spin densities and Fermi-level shifts that drive spontaneous metal reduction. The work demonstrates that low-temperature mechanical processing can transform an otherwise inert 2-D insulator into a versatile, defect-rich support for next-generation single-atom catalysts.
See also: | Hexagonal boron nitride | Vacancy defect | Heterogeneous catalysis | Hydrogen evolution reaction | Electrocatalyst |
Arsenic-doped black phosphorus (b-AsP): Field-effect transistors
Black phosphorus (b-P) is a layered semiconductor whose band gap and carrier mobility can be tuned simply by changing the number of atomic layers, making it attractive for infrared opto-electronics and thin-film transistors. In 2019 the lab of Dr. Mendoza-Cortes and et al. examined how replacing one-quarter of the phosphorus atoms with arsenic (composition ≈ P0.75As0.25) modifies both the vibrational and electronic behaviour of few-layer flakes.[70] Raman spectroscopy. Density-functional-theory (DFT) calculations show that each of b-P's three characteristic phonon modes splits into a triplet of P–P, P–As and As–As vibrations; the measured Raman spectrum matches this prediction, confirming uniform alloying. Field-effect transistors. Few-layer b-AsP flakes on Si/SiO2 operate as p-type channels with a room-temperature on/off ratio ≈ 103 and intrinsic field-effect mobility ~ 300 cm2 V-1 s-1; mobility doubles to ~ 600 cm2 V-1 s-1 at 100 K, values as good as, or better than, pristine b-P. Gate-induced insulator–metal transition. At large gate voltages the sheet resistivity falls with decreasing temperature and saturates below ≈ 100 K, signalling a transition to a metallic state. Anisotropic behaviour. Like undoped b-P, transport remains highly direction-dependent, with markedly higher conductivity along the arm-chair crystallographic axis than along the zig-zag axis. Periodic hybrid-DFT band-structure calculations place the Fermi level just beneath the valence-band maximum, explaining the observed hole conduction. Because its band gap, carrier type and in-plane anisotropy can all be tuned by arsenic content and layer number, b-AsP is proposed as a versatile 2-D material for infrared photodetectors and anisotropic electronics.
See also: | Black phosphorus | Two-dimensional materials | Field-effect transistor | Raman spectroscopy | Anisotropy
Carbon-doped Silicon: Si-compatible spin-field-effect transistors
Silicene, the silicon analogue of graphene, offers compatibility with existing Si-based electronics but lacks an intrinsic band gap. In 2020 the lab of Dr. Mendoza-Cortes and Dr. Van Voorhis at MIT developed and deployed first-principles calculations to show that substituting silicon atoms with carbon atoms introduces controllable lattice disorder and electron–electron correlations that radically alter silicene's electronic and magnetic properties.[71] Disorder-tunable band gap. Varying the position and concentration of C dopants transforms metallic silicene into a semiconductor; the resulting band gap scales with the degree of structural disorder. Mott–Anderson transition. At critical dopant configurations the system undergoes a combined correlation- and disorder-driven transition from a delocalised to a localised electronic state, analogous to the Mott–Anderson transition predicted in bulk oxides. Room-temperature ferromagnetism. Spin-polarised calculations reveal robust ferromagnetism even in highly disordered monolayers, suggesting prospects for Si-compatible spin-field-effect transistors. The study highlights carbon doping as a practical route to engineer band gaps, magnetic order and correlation phenomena in silicene and potentially other two-dimensional materials, bridging fundamental physics and silicon-based nanoelectronics.
See also: | Silicene | Mott transition | Anderson localization | Two-dimensional materials | Spintronics |
Connection of Chemistry, Physics and Engineering
Restricted multilayer theory for gas adsorption
Traditional models of gas uptake in porous solids, Langmuir (single layer) and BET (infinite multilayer) isotherms, struggle to describe modern crystalline adsorbents such as metal–organic frameworks (MOFs), whose pore sizes cap the number of layers that can form. In 2018, the Mendoza-Cortés lab introduced a restricted multilayer theory (RMT) that bridges the two classics by accounting explicitly for pore geometry and host–guest interaction energy.[72] Key idea – The maximum number of adsorbed layers, nmax, is limited by the pore diameter; RMT derives a closed-form isotherm that reduces to Langmuir when nmax=1 and to BET as nmax→∞. Predictive power – With only geometric inputs and a single host–guest binding energy, the model reproduces experimental H₂ adsorption isotherms for seven benchmark MOFs over a wide pressure range. Design tool – RMT yields universal formulas for the optimal linker length (and thus pore size) that maximises storage capacity in a given MOF topology, guidance valuable for hydrogen-storage and gas-separation materials. Generality – Although demonstrated for H₂ in MOFs, the framework applies to any light gas in any crystalline porous material, providing a first-principles starting point for next-generation adsorbent design.
See also: | Adsorption | Langmuir adsorption model | BET theory | Metal–organic framework | Hydrogen storage |
Connection of Biology, Chemistry and Physics
Ubiquity of the Cubane motifs in proteins that catalyzes biochemical reactions
A 2022 work by the lab of Dr. Zdilla and Dr. Mendoza-Cortes charts how the heterocubane motif, a cube-shaped metal-chalcogen (or metal-oxo) cluster, recurs from fundamental biological enzymes to modern coordination-chemistry model complexes.[73] Biological importance. Cubane cores, most famously Fe4S4, Mn4O5Ca in photosystem II and the FeMo cofactor of nitrogenase, enable reactions ranging from single-electron transfer to the multi-electron water-splitting and N≡N bond cleavage that underpin photosynthesis and nitrogen fixation. Synthetic landscape. The review catalogues decades of laboratory syntheses that replace one or more iron atoms with Mn, Co, Ni, Mo or V, or swap oxide/sulfide ligands for phosphide, nitride or halide, revealing systematic trends in redox potential and magnetic coupling. Coordination-environment effects. Comparing "naked" cubanes with protein-bound clusters shows how ligand denticity, protonation and hydrogen-bond networks tune electronic structure; these insights guide the design of artificial water-oxidation and CO2-reduction catalysts. Spectroscopic & computational tools. Advances in Mössbauer, X-ray emission, ultrafast spectroscopy and density-functional theory now permit atom-by-atom assignment of oxidation states and spin populations, even in mixed-metal cubanes, bridging the gap between bioinorganic and synthetic communities. By stitching together biological examples with historically separate synthetic chemistry, the authors argue that heterocubanes form a unifying scaffold for understanding, and ultimately emulating, nature's most demanding redox transformations.
See also: | Iron–sulfur protein | Oxygen-evolving complex | Nitrogenase | Bioinorganic chemistry |
Connection of Biology, Chemistry, Physics and Medicine
Understanding the mistery of diabetes: Microfluidic monitoring of liver-cell glucose metabolism
Maintaining blood-glucose levels depends on how quickly the liver adjusts its uptake and release of glucose in response to hormones such as insulin, whose concentration can fluctuate on the scale of minutes. In 2019 the lab of Dr. Mendoza-Cortes and Dr. Roper designed, developed and engineered a modular microfluidic bioreactor that records these rapid metabolic changes in real time for up to 106 human HepG2 liver cells.[74] Technology: A perfused micro-chamber keeps the cells viable for ten days; outflow is mixed with reagents for an enzymatic glucose assay and segmented into droplets whose fluorescence gives a glucose read-out every ≈100 s across the 0–12 mM range. Key finding: When stepwise or pulsatile insulin profiles are perfused, the system detects correspondingly fast shifts in cellular glucose consumption, demonstrating liver-like responsiveness. Broader impact: Because the assay can be adapted to other metabolites (lactate, pyruvate, ketone bodies), the platform offers a route to correlate real-time pancreatic-hormone dynamics with hepatic metabolism, useful for diabetes and drug-discovery research.
See also: | Microfluidics | Glucose metabolism | Insulin | HepG2 | Diabetes |
References
- ^ "45th Annual Telly Awards – Winners Highlights". YouTube. The Institute for Cyber-Enabled Research. 23 May 2024. Retrieved 16 June 2025. "The 45th Annual Telly Awards – Winners Press Release". TellyAwards.com. The Telly Awards. 23 May 2024. Retrieved 16 June 2025.
- ^ a b "Jose L. Mendoza-Cortes - Michigan State University". www.egr.msu.edu. Retrieved 2023-10-12.
- ^ "Jose L. Mendoza-Cortes - Academic Tree". academictree.org. Retrieved 2023-12-31.
- ^ "IChO Results". www.icho-official.org. Retrieved 2023-10-12.
- ^ Mendoza-Cortes, Jose Luis (2012). Design of Molecules and Materials for Applications in Clean Energy, Catalysis and Molecular Machines Through Quantum Mechanics, Molecular Dynamics and Monte Carlo Simulations (phd thesis). California Institute of Technology.
- ^ El-Kaderi, Hani M.; Hunt, Joseph R.; Mendoza-Cortés, José L.; CôTé, Adrien P.; Taylor, Robert E.; O'Keeffe, Michael; Yaghi, Omar M. (2007-04-13). "Designed Synthesis of 3D Covalent Organic Frameworks". Science. 316 (5822): 268–272. Bibcode:2007Sci...316..268E. doi:10.1126/science.1139915. ISSN 0036-8075. PMID 17431178. S2CID 19555677.
- ^ Advancement, Research Corporation for Science. "Scialog – NES Fellows and Facilitators". Research Corporation for Science Advancement. Retrieved 2023-10-12.
- ^ "Contributors to the Emerging Investigators Issue 2018". Chemical Communications. 54 (50): 6442–6457. 2018-06-19. doi:10.1039/C8CC90258E. ISSN 1364-548X.
- ^ Patronis, Amy Farnum (2018-01-09). "GAP awards help FSU faculty propel research from lab to market". Florida State University News. Retrieved 2023-10-12.
- ^ "COM- 2019-Annual-Report" (PDF).
- ^ "IGEN Takes the APS Bridge Program to the Next Level". www.aps.org. Retrieved 2023-10-12.
- ^ "This Chemical Discovery Is A Big Step Forward For Clean Energy". forbes.com. Retrieved 2024-01-08.
- ^ "Power plant: High-tech photosynthesis". cnbc.com/. 27 November 2015. Retrieved 2024-01-08.
- ^ Berkowitz, Rachel (2017). "Energy Focus: New simulations suggest cost-effective materials design for H2 storage". MRS Bulletin. 42 (2): 87. doi:10.1557/mrs.2017.11. Retrieved 2024-01-08.
- ^ "Jose L. Mendoza-Cortes - Scholar Profile". scholar.google.com/. Retrieved 2023-12-31.
- ^ Morris, Ian M.; Klink, Kai; Singh, Jaideep T.; Mendoza-Cortes, José L.; Nicley, Shannon S.; Becker, Jonas N. (2024-01-22). "Rare isotope-containing diamond colour centres for fundamental symmetry tests". Philosophical Transactions of the Royal Society A. 382 (2265): 20230169. doi:10.1098/rsta.2023.0169. PMC 10693981. PMID 38043574.
- ^ a b Mendoza-Cortés, José L.; Dolores-Cuenca, Eric; Aguirre, Carlos; Tzeng, Yu-Ying; Xu, Guanglei; Crock, Nathan; Aduenko, Alex (2025). "Machine Learning and Quantum Computing Guide for Humans" (e-book).
- ^ Dolores-Cuenca, Eric; Mendoza-Cortes, Jose L. (2022). "A poset version of Ramanujan results on Eulerian numbers and zeta values". arXiv:2205.05208v3 [math.CO].
- ^ Dolores-Cuenca, Eric; Arciniega-Nevárez, José Antonio; Nguyen, Anh; Zou, Amanda Yitong; Van Popering, Luke; Crock, Nathan; Erlebacher, Gordon; Mendoza-Cortes, Jose L. (2022-03-02). "Polychrony as Chinampas". Algorithms. 16 (4): 193. arXiv:2103.15265. doi:10.3390/a16040193. ISSN 1999-4893.
- ^ Dolores-Cuenca, E.; Guzmán-Sáenz, A.; Kim, S.; López-Moreno, S.; Mendoza-Cortés, J. L. (2024). "Order Theory in the Context of Machine Learning". arXiv:2412.06097 [cs.CV].
- ^ Alday-Toledo, Leon; Bernal-Jaquez, Roberto; Zapotecas-Martinez, Saul; Mendoza-Cortes, Jose L. (2022). "Obtaining transferable chemical insight from solving machine-learning classification problems: Thermodynamical properties prediction, atomic composition as good as Coulomb matrix". arXiv:2212.01420 [physics.chem-ph].
- ^ Zhang, Xu; Nguyen, Hoang; Paci, Jeffrey T.; Sankaranarayanan, Subramanian K. R. S.; Mendoza-Cortés, José L.; Espinosa, Horacio D. (2021). "Multi-objective parametrization of interatomic potentials for large deformation pathways and fracture of two-dimensional materials". Nature Computational Materials. 7 (1): 113. doi:10.1038/s43588-020-00019-6 (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ a b Rodriguez, Austin; Smith, Justin S.; Mendoza-Cortes, José L. (2025). "Does Hessian Data Improve the Performance of Machine Learning Potentials?". arXiv:2503.07839 [physics.chem-ph].
- ^ Rodriguez, Austin; Smith, Justin S.; Mendoza-Cortés, José L. (2025). "OpenReACT-CHON-EFH : Open REaction Dataset of Atomic ConfiguraTions comprising C, H, O, N with Energies, Forces, and Hessians". 4. Figshare. Retrieved 30 May 2025.
- ^ Lucht, Kevin P.; Mahabir, A. D.; Alcantara, A.; Balatsky, A. V.; Mendoza-Cortes, Jose L.; Haraldsen, J. T. (2019). "Designing a path towards superconductivity through magnetic exchange in transition-metal intercalated bilayer graphene". arXiv:1903.10112 [cond-mat.mtrl-sci].
- ^ Jelic, Vedran; Adams, Stefanie; Maldonado-Lopez, Daniel; Buliyaminu, Ismail A.; Hassan, Mohamed; Mendoza-Cortes, Jose L.; Cocker, Tyler L. (2024). "Terahertz control of surface topology probed with subatomic resolution". arXiv:2411.07545 [cond-mat.mtrl-sci].
- ^ Zapata-Escobar, Andy D.; Pakhira, Srimanta; Barroso-Flores, Joaquín; Aucar, Gustavo A.; Mendoza-Cortés, José L. (2023-01-12). "Relativistic quantum calculations to understand the contribution of f-type atomic orbitals and chemical bonding of actinides with organic ligands". Physical Chemistry Chemical Physics. 25 (7): 5592–5601. arXiv:2108.06057. Bibcode:2023PCCP...25.5592Z. doi:10.1039/D2CP05399C. PMID 36727265.
- ^ Gannon, Ashley; Marxsen, Stephanie; Mueller, Kevin; Quaife, Bryan D.; Mendoza-Cortes, Jose L. (2018). "Using a High-Throughput Screening Algorithm and relativistic Density Functional Theory to Find Chelating Agents for Separation of Radioactive Waste". arXiv:1806.09733 [physics.chem-ph].
- ^ Galley, Shane S.; Pattenaude, Scott A.; Gaggioli, Carlo A.; Qiao, Yusen; Sperling, Joseph M.; Zeller, Matthias; Pakhira, Srimanta; Mendoza-Cortés, José L.; Schelter, Eric J.; Albrecht-Schmitt, Thomas E. (2019). "Synthesis and Characterization of Tris-chelate Complexes for Understanding f-Orbital Bonding in Later Actinides". Journal of the American Chemical Society. 141 (6): 2356–2366. Bibcode:2019JAChS.141.2356G. doi:10.1021/jacs.8b10251. PMID 30714372.
- ^ Zapata-Escobar, Andy D.; Maldonado, Alejandro F.; Mendoza-Cortés, José L.; Aucar, Gustavo A. (2023). "NMR Magnetic Shielding in Transition Metal Compounds Containing Cadmium, Platinum, and Mercury". Magnetochemistry. 9 (7): 165. doi:10.3390/magnetochemistry9070165.
- ^ Comaskey, William P.; Bodo, Filippo; Erba, Alessandro; Mendoza-Cortes, José L.; Desmarais, Jacques K. (2022-11-17). "Role of spin currents on electron–electron interaction in the quantum spin Hall phase". Physical Review B. 106 (20): L201109. arXiv:2208.13878. Bibcode:2022PhRvB.106t1109C. doi:10.1103/PhysRevB.106.L201109.
- ^ Pakhira, Srimanta; Mendoza-Cortés, José L. (2020). "Quantum Nature in the Interaction of Molecular Hydrogen with Porous Materials: Implications for Practical Hydrogen Storage". The Journal of Physical Chemistry C. 124 (11): 6454–6460. doi:10.1021/acs.jpcc.9b11556 (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Bryant, Jacob T.; Logan, Matthew W.; Chen, Zhihengyu; Djokic, Marcus; Cairnie, Daniel R.; Vazquez-Molina, Demetrius A.; Nijamudheen, A.; Langlois, Kyle R.; Markley, Michael J.; Pombar, Gisselle (2023). "Synergistic Steric and Electronic Effects on the Photoredox Catalysis by a Multivariate Library of Titania Metal–Organic Frameworks". Journal of the American Chemical Society. 145 (8): 4589–4600. doi:10.1021/jacs.2c12639 (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Lucht, Kevin P.; Mendoza-Cortés, José L. (2015). "Birnessite: A Layered Manganese Oxide to Capture Sunlight for Water-Splitting Catalysis". Journal of Physical Chemistry C. 119 (40): 22838–22846. doi:10.1021/acs.jpcc.5b05975 (inactive 12 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Stafford, Sean M.; Aduenko, Alexander; Djokic, Marcus; Lin, Yu-Hsiu; Mendoza-Cortés, José L. (2023). "Transforming materials discovery for artificial photosynthesis: High-throughput screening of earth-abundant semiconductors". Journal of Applied Physics. 134 (23): 235706. arXiv:2310.00118. Bibcode:2023JAP...134w5706S. doi:10.1063/5.0178907.
- ^ a b c Djokic, Marcus; Mendoza-Cortés, José L. (2023-11-28). "Multi-Binding Sites United in Covalent-Organic Frameworks (MSUCOF) for H2 Storage and Delivery at Room Temperature". Energy & Fuels. 38 (5): 4711–4720. doi:10.1021/acs.energyfuels.3c04075. S2CID 259203728.
- ^ "DOE Technical Targets for Onboard Hydrogen Storage for Light-Duty Vehicles". U.S. Department of Energy. Retrieved 21 May 2025.
- ^ Pramudya, Yohanes; Mendoza-Cortés, José L. (2016). "Design Principles for High H₂ Storage Using Chelation of Abundant Transition Metals in Covalent-Organic Frameworks for 0–700 bar at 298 K". Journal of the American Chemical Society. 138 (46): 15204. doi:10.1021/jacs.6b08803.
- ^ Berkowitz, Rachel (2017). "Energy Focus: New simulations suggest cost-effective materials design for H₂ storage". MRS Bulletin. 42 (2): 87. doi:10.1557/mrs.2017.9.
- ^ Lei, Yu; Pakhira, Srimanta; Fujisawa, Kazunori; Wang, Xuyang; Iyiola, Oluwagbenga O.; Perea-López, Néstor; Elías, Ana Laura; Rajukumar, Lakshmy Pulickal; Zhou, Chanjing; Kabius, Bernd (2017). "Low-Temperature Synthesis of Heterostructures of Transition-Metal Dichalcogenide Alloys (WₓMo₁–ₓS₂) and Graphene with Superior Catalytic Performance for Hydrogen Evolution". ACS Nano. 11 (5): 5103–5112. doi:10.1021/acsnano.7b00406 (inactive 12 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Niu, Wenhan; Pakhira, Srimanta; Cheng, Guangming; Zhao, Fang; Yao, Nan; Mendoza-Cortés, José L.; Koel, Bruce E. (2024). "Reaction-driven restructuring of defective PtSe2 into ultrastable catalyst for the oxygen reduction reaction". Nature Materials. 23 (12): 1704–1711. doi:10.1038/s41563-024-02020-w. PMID 39375480.
- ^ Ding, R.; Maldonado-Lopez, D.; Henebry, J. E.; Mendoza-Cortes, J. L.; Zdilla, M. J. (2025). "Enhanced activity in layered-metal-oxide-based oxygen evolution catalysts by layer-by-layer modulation of metal ion identity". arXiv:2504.20885 [cond-mat.mtrl-sci].
- ^ Liang, Kun; Pakhira, Srimanta; Yang, Zhenzhong; Nijamudheen, A.; Ju, Licheng; Wang, Maoyu; Aguirre-Velez, Carlos I.; Sterbinsky, George E.; Du, Yingge; Feng, Zhenxing (2019). "S-Doped MoP Nanoporous Layer toward High-Efficiency Hydrogen Evolution in pH-Universal Electrolyte". ACS Catalysis. 9 (1): 651–659. doi:10.1021/acscatal.8b03259 (inactive 12 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Maldonado-López, Daniel; Huang, Po-Wei; Sánchez-Liévano, Karla R.; Jana, Gourhari; Mendoza-Cortes, José L.; Knowles, Kathryn E.; Hatzell, Marta C. (2024). "Nitrogen-Containing Surface Ligands Lead to False Positives for Photofixation of N2 on Metal Oxide Nanocrystals: An Experimental and Theoretical Study". Advanced Functional Materials. 35 (3): 2413319. doi:10.1002/adfm.202413319.
- ^ Hill, Caleb M.; Mendoza-Cortés, José L.; Velázquez, Jesús M.; Whittaker-Brooks, Luisa (2022). "Multi-dimensional designer catalysts for negative emissions science: Bridging the gap between synthesis, simulations, and analysis". iScience. 25 (1): 103627. doi:10.1016/j.isci.2021.103627.
- ^ Garza, Alejandro J.; Pakhira, Srimanta; Bell, Alexis T.; Mendoza-Cortés, José L.; Head-Gordon, Martin (2018). "Reaction mechanism of the selective reduction of CO₂ to CO by a tetraaza [Co II N₄H]²⁺ complex in the presence of protons". Physical Chemistry Chemical Physics. 20 (37): 24058–24064. doi:10.1039/C8CP04783K (inactive 12 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Safaie, Niloofar; Rodriguez, Austin; Jana, Gourhari; Smak, Jessica; Mendoza-Cortés, José L.; Ferrier Jr., Robert C. (2023). "Unveiling the Mechanisms of Epoxide Polymerization with N-Al Adduct Catalysts: A Comprehensive Experimental and Theoretical Investigation". Polymer Chemistry. 14 (27): 3213–3224. doi:10.1039/d3py00555d (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Ferrier Jr, Robert C.; Pakhira, Srimanta; Palmon, Sarah E.; Rodriguez, Christina G.; Goldfeld, David J.; Iyiola, Oluwagbenga O.; Chwatko, Malgorzata; Mendoza-Cortés, José L.; Lynd, Nathaniel A. (2018). "Demystifying the Mechanism of Regio- and Isoselective Epoxide Polymerization Using the Vandenberg Catalyst". Macromolecules. 51 (5): 1777–1786. doi:10.1021/acs.macromol.7b02447.
- ^ Choudhury, Snehashis; Tu, Zhengyuan; Nijamudheen, A.; Zachman, Michael J.; Stalin, Sanjuna; Deng, Yue; Zhao, Qing; Vu, Duylinh; Kourkoutis, Lena F.; Mendoza-Cortés, José L.; Archer, Lynden A. (2019-07-12). "Stabilizing polymer electrolytes in high-voltage lithium batteries". Nature Communications. 10 (1): 3091. Bibcode:2019NatCo..10.3091C. doi:10.1038/s41467-019-11015-0. PMC 6626095. PMID 31300653.
- ^ Sarbapalli, Dipobrato; Lin, Yu-Hsiu; Stafford, Sean M.; Son, Jangyup; Mishra, Abhiroop; Hui, Jingshu; Nijamudheen, A.; Romo, Adolfo I. B.; Gossage, Zachary T.; van der Zande, Arend M. (2022). "A surface modification strategy towards reversible Na-ion intercalation on graphitic carbon using fluorinated few-layer graphene". Journal of the Electrochemical Society. 169 (10): 106522. doi:10.1149/1945-7111/ac9f52 (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Hui, Jingshu; Schorr, Noah B.; Pakhira, Srimanta; Qu, Zihan; Mendoza-Cortés, José L.; Rodríguez-López, Joaquín (2018-10-08). "Achieving fast and efficient K⁺ intercalation on ultrathin graphene electrodes modified by a Li⁺-based solid-electrolyte interphase". Journal of the American Chemical Society. 140 (42): 13599–13603. doi:10.1021/jacs.8b08907. PMID 30299954. S2CID 52947242.
- ^ Hui, Jingshu; Nijamudheen, A.; Sarbapalli, Dipobrato; Xia, Chang; Qu, Zihan; Mendoza-Cortés, José L.; Rodríguez-López, Joaquín (2021). "Nernstian Li+ intercalation into few-layer graphene and its use for the determination of K+ co-intercalation processes". Chemical Science. 12 (2): 559–568. doi:10.1039/d0sc04919d (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Lin, Yu-Hsiu; Mendoza-Cortés, José L. (2024). "Beyond lithium-ion batteries: Are effective electrodes possible for alkaline and other alkali elements? Exploring ion intercalation in surface-modified few-layer graphene and examining layer quantity and stages". Journal of Applied Physics. 136 (7): 075001. doi:10.1063/5.0187651. PMID 38411468.
- ^ Nijamudheen, A.; Sarbapalli, Dipobrato; Hui, Jingshu; Rodríguez-López, Joaquín; Mendoza-Cortés, José L. (2020). "Impact of surface modification on the lithium, sodium, and potassium intercalation efficiency and capacity of few-layer graphene electrodes". ACS Applied Materials & Interfaces. 12 (17): 19393–19401. doi:10.1021/acsami.0c04705 (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Vázquez-Molina, Demetrius A.; Pope, Giovanna M.; Ezazi, Andrew A.; Mendoza-Cortés, José L.; Harper, James K.; Uribe-Romo, Fernando J. (2018). "Framework vs. side-chain amphidynamic behaviour in oligo-(ethylene oxide)-functionalised covalent-organic frameworks". Chemical Communications. 54 (50): 6947–6950. doi:10.1039/C8CC03245F (inactive 12 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Hu, Yiming; Dunlap, Nathan; Long, Hai; Chen, Hongxuan; Wayment, Lacey J.; Ortiz, Michael; Jin, Yinghua; Nijamudheen, Abdulrahiman; Mendoza-Cortés, José L.; Lee, Se-Hee (2021). "Helical covalent polymers with unidirectional ion channels as single lithium-ion conducting electrolytes". CCS Chemistry. 3 (12): 2762–2770. doi:10.31635/ccschem.021.202100588 (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Niu, Wenhan; Pakhira, Srimanta; Marcus, Kyle; Li, Zhao; Mendoza-Cortés, José L.; Yang, Yang (2018). "Apically Dominant Mechanism for Improving Catalytic Activities of N-Doped Carbon Nanotube Arrays in Rechargeable Zinc–Air Battery". Advanced Energy Materials. 8 (20): 1800480. doi:10.1002/aenm.201800480.
- ^ Lin, Yu-Chuan; Torsi, Riccardo; Younas, Rehan; Hinkle, Christopher L.; Rigosi, Albert F.; Hill, Heather M.; Zhang, Kunyan; Huang, Shengxi; Mendoza-Cortes, Jose L.; al, et (2023). "Recent Advances in 2D Material Theory, Synthesis, Properties, and Applications". ACS Nano. 17 (11): 9694–9747. doi:10.1021/acsnano.2c12759. PMC 10324635. PMID 37219929.
- ^ Lin, Yu-Hsiu; Comaskey, William P.; Mendoza-Cortes, José L. (2025). "How Can We Engineer Electronic Transitions Through Twisting and Stacking in TMDC Bilayers and Heterostructures? A First-Principles Approach". Nanoscale Advances. 7 (7): 2047–2056. Bibcode:2025NanoA...7.2047L. doi:10.1039/d5na00112a. PMC 11833679. PMID 39974342.
- ^ Pablo-Pedro, Ricardo; Lopez-Rios, Héctor; Mendoza-Cortés, José L.; Kong, Jing; Fomine, Serguei; Van Voorhis, Troy; Dresselhaus, Mildred S. (2018-05-08). "Exploring low internal reorganisation energies for silicene nanoclusters". Physical Review Applied. 9 (5): 054012. doi:10.1103/PhysRevApplied.9.054012.
- ^ Hui, Jingshu; Pakhira, Srimanta; Bhargava, Richa; Barton, Zachary J.; Zhou, Xuan; Chinderle, Adam J.; Mendoza-Cortés, José L.; Rodríguez-López, Joaquín (2018). "Modulating Electrocatalysis on Graphene Heterostructures: Physically Impermeable yet Electronically Transparent Electrodes". ACS Nano. 12 (3): 2980–2990. doi:10.1021/acsnano.7b08418.
- ^ Maldonado-Lopez, Daniel; Mendoza-Cortes, José L. (2023). "Exquisite control of electronic and spintronic properties on highly porous covalent-organic frameworks: transition-metal intercalation in bilayers". Physica Scripta. 98 (10): 105926. doi:10.1088/1402-4896/acf4e9 (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Pakhira, Srimanta; Lucht, Kevin P.; Mendoza-Cortés, José L. (2018). "Dirac cone in two-dimensional bilayer graphene by intercalation with V, Nb and Ta transition metals". The Journal of Chemical Physics. 148 (6): 064707. doi:10.1063/1.5012994.
- ^ Pakhira, Srimanta; Mendoza-Cortés, José L. (2018). "Tuning the Dirac Cone of Bilayer and Bulk-Structure Graphene by Intercalating First-Row Transition Metals Using First-Principles Calculations". The Journal of Physical Chemistry C. 122 (9): 4768–4782. doi:10.1021/acs.jpcc.7b12048 (inactive 12 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Pakhira, Srimanta; Mendoza-Cortés, José L. (2019). "Intercalation of First-Row Transition Metals inside Covalent-Organic Frameworks: A Strategy to Fine-Tune the Electronic Properties of Porous Crystalline Materials". Physical Chemistry Chemical Physics. 21 (17): 8785–8796. doi:10.1039/C9CP00513A (inactive 1 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Pakhira, Srimanta; Lucht, Kevin P.; Mendoza-Cortés, José L. (2017). "Iron intercalation in covalent-organic frameworks: A promising approach for semiconductors". The Journal of Physical Chemistry C. 121 (39): 21160–21170. doi:10.1021/acs.jpcc.7b06228.
- ^ Jana, Gourhari; Mendoza-Cortes, José L. (2023). "Thermodynamics, kinetics, and optical properties of rotaxane: a first-principles molecular dynamics study". The Journal of Physical Chemistry A. 127 (12): 2671–2687. doi:10.1021/acs.jpca.3c00210. PMID 36944165.
- ^ Dharmarwardana, Madushani; Pakhira, Srimanta; Welch, Raymond P.; Caicedo-Narvaez, Carlos; Luzuriaga, Michael A.; Arimilli, Bhargav S.; McCandless, Gregory T.; Fahimi, Babak; Mendoza-Cortés, José L.; Gassensmith, Jeremiah J. (2021). "Rapidly Reversible Organic Crystalline Switch for Conversion of Heat into Mechanical Energy". Journal of the American Chemical Society. 143 (15): 5951–5957. doi:10.1021/jacs.1c00650 (inactive 1 July 2025). PMID 33822596.
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Lei, Yu; Pakhira, Srimanta; Fujisawa, Kazunori; Liu, He; Guerrero-Bermea, Cynthia; Zhang, Tianyi; Dasgupta, Archi; Martinez, Luis M.; Singamaneni, Srinivasa Rao; Wang, Ke (2021). "Low-temperature activation of inert hexagonal boron nitride for metal deposition and single-atom catalysis". Materials Today. 51: 108–116. doi:10.1016/j.mattod.2021.10.013.
- ^ Pradhan, Nihar R.; Garcia, Carlos; Lucking, Michael C.; Pakhira, Srimanta; Martinez, Juan; Rosenmann, Daniel; Divan, Ralu; Sumant, Anirudha V.; Terrones, Humberto; Mendoza-Cortés, José L. (2019). "Raman and electrical transport properties of few-layered arsenic-doped black phosphorus". Nanoscale. 11 (39): 18449–18463. doi:10.1039/c9nr04598h. OSTI 1568937. PMID 31576874.
- ^ Pablo-Pedro, Ricardo; Magaña-Fuentes, Miguel A.; Videa, Marcelo; Kong, Jing; Li, Mingda; Mendoza-Cortés, José L.; Van Voorhis, Troy (2020). "Understanding disorder in 2D materials: the case of carbon doping of silicene". Nano Letters. 20 (9): 6336–6343. arXiv:1912.00333. Bibcode:2020NanoL..20.6336P. doi:10.1021/acs.nanolett.0c01775. PMID 32787169.
- ^ Aduenko, Alexander A.; Murray, Andy; Mendoza-Cortés, José L. (2018). "General Theory of Adsorption in Porous Materials: Restricted Multilayer Theory". ACS Applied Materials & Interfaces. 10 (15): 13244–13251. doi:10.1021/acsami.7b19306 (inactive 12 July 2025).
{{cite journal}}
: CS1 maint: DOI inactive as of July 2025 (link) - ^ Bigness, Alec; Vaddypally, Shivaiah; Zdilla, Michael J.; Mendoza-Cortés, José L. (2022). "Ubiquity of cubanes in bioinorganic relevant compounds". Coordination Chemistry Reviews. 450 214168. doi:10.1016/j.ccr.2021.214168.
- ^ Adams, Anna G.; Bulusu, Radha Krishna Murthy; Mukhitov, Nikita; Mendoza-Cortés, José L.; Roper, Michael G. (2019-04-16). "Online measurement of glucose consumption from HepG2 cells using an integrated bioreactor and enzymatic assay". Analytical Chemistry. 91 (8): 5184–5190. doi:10.1021/acs.analchem.8b05798. PMC 6472493. PMID 30884946.