Dan C. Sorescu’s research while affiliated with University of Pittsburgh and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (154)


Defect Thermodynamics and Transport Properties of Proton Conducting Perovskite Electrode and Electrolyte Materials Evaluated Based on Density Functional Theory Modeling
  • Conference Paper

January 2025

·

11 Reads

Yueh-Lin Lee

·

·

Dan Sorescu

·

[...]

·

Harry Abernathy

Workflow of the developed defect model solver that includes the hydride defect formation reaction. The defect model equations are summarized in Table II. CLa and CM are concentrations of La and M (M = non-transition metal) substitutions at the A-site and B-site of BaFeO3−δ, respectively. QM is the M metal charge relative to Fe³⁺. ΔHdef and ΔSdef are the defect reaction enthalpy and entropy expressed as polynomial functions (up to sixth order) of oxygen nonstoichiometry (δ). The targeted temperature range in the defect model solver routine is within 700–1200 K, and the P(O2) limits are constrained by the oxygen nonstoichiometry range of δ = 0–0.8 following the scheme proposed by Poulsen.³⁷ The P(H2O) pressure is set to be within 10–5 to 1 atm range.
Stability of dry (empty bar) and hydrated (blue bar) BaFeO3−δ at various δ and for different magnetic ordering arrangements: ferromagnetic (FM), A-type antiferromagnetic (AAF), C-type antiferromagnetic (CAF), and G-type antiferromagnetic (GAF).¹³ The structural configurations are provided in the schematics on the right. The letter before the underscore symbol indicates the spin direction with respect to crystallographic orientations a, b, and c. All the images of the atomistic configurations are created using VESTA.⁴⁹ The green and brown spheres are Ba and Fe atoms, respectively, and oxygen atoms are located at the corners of the polyhedrals. It is noted that the number of vacant lattice oxygen sites is used to describe δ values shown in Fig. 2. For example, an oxygen vacancy in a 2 × 2 × 2 Ba8Fe8O24 supercell corresponds to δ = 0.125. For the hydrated structure with δ = 0, it is considered that a H2O molecule is dissociated and filled into the oxygen vacancy of the dry Ba8Fe8O23 with δ = 0.125 such that the only oxygen vacancy site is filled with the oxygen from the H2O molecule accompanied by two proton formation, i.e., H2O + VO·· + OOx ↔ 2 OHO· . As there is no vacant O lattice site left in the hydrated structure Ba8Fe8O24H2, δ is then assigned to 0.
Stability of the simulated La0.125Ba0.875FeO2.875 (δ = 0.125) configurations with an incorporated hydrogen in the form of proton OHO· (H⁺, blue empty circles, with an OH bond length of ∼1Ǻ) and hydride HO· (H⁻, orange empty diamonds, with formation of an Fe-H bond) in (a) G-type antiferromagnetic (GAF), (b) C-type antiferromagnetic (CAF), (c) A-type antiferromagnetic (AAF), and (d) ferromagnetic (FM) structures. The relative stability of the most stable hydride ( HO· ) defect configuration relative to the most stable proton ( OHO· ) incorporated configuration (which is set to be the reference configuration) is provided by the labeled number in Figs. 3a–3d. Schematics of representative proton ( OHO· ) containing and hydride ( HO· ) configurations are provided in Figs. 3e and 3f, respectively. Light green, dark green, brown, red, and light blue spheres represent Ba, La, Fe, O, and H atoms, respectively. The relative stabilities of these representative H incorporated configurations are indicated with filled symbols in Figs. 3a–3d.
Stability of the simulated La0.125Ba0.875FeO2.625 (δ = 0.375) configurations with an incorporated hydrogen in the form of proton OHO· (H⁺, blue empty circles, with an OH bond length of ∼1Ǻ) and hydride HO· (H⁻, orange empty diamonds, with formation of an Fe-H bond) in (a) G-type antiferromagnetic (GAF), (b) C-type antiferromagnetic (CAF), (c) A-type antiferromagnetic (AAF), and (d) ferromagnetic (FM) structures. The relative stability of the most stable hydride ( HO· ) defect configuration relative to the most stable proton ( OHO· ) incorporated configuration (which is set to be the reference configuration) is provided by the labeled number in Figs. 4a–4d. Schematics of representative proton ( OHO· ) and hydride ( HO· ) configurations are provided in Figs. 4e and 4f, respectively. Light green, dark green, brown, red, and light blue spheres represent Ba, La, Fe, O, and H atoms, respectively. The relative stabilities of these representative H incorporated configurations are indicated with the filled symbols in Figs. 4a–4d.
Stability of the simulated Ba0.875La0.125FeO2.5 (δ = 0.5) configurations with an incorporated hydrogen in the form of proton OHO· (H⁺, blue empty circles, with an OH bond length of ∼1Ǻ) and hydride HO· (H⁻, orange empty diamonds, with formation of an Fe-H bond) in (a) G-type antiferromagnetic (GAF), (b) C-type antiferromagnetic (CAF), (c) A-type antiferromagnetic (AAF), and (d) ferromagnetic (FM) structures. The relative stability of the most stable hydride ( HO· ) defect configuration relative to the most stable proton ( OHO· ) incorporated configuration (which is set to be the reference configuration) is provided by the labeled number in Figs. 5a–5d. Schematics of representative proton ( OHO· ) and hydride ( HO· ) configurations are provided in Figs. 5e and 5f, respectively Light green, dark green, brown, red, and light blue spheres represent Ba, La, Fe, O, and H atoms, respectively. The relative stabilities of these representative H incorporated configurations is indicated with filled symbols in Figs. 5a–5d.

+5

Thermodynamic Modeling of Point Defects in Triple Conducting Perovskite Ba0.95La0.05FeO3−δ with Incorporation of the Hydride Defect Formation Reaction for Solid Oxide Cells
  • Article
  • Publisher preview available

December 2024

·

42 Reads

Distinct from the proton defect, the hydride defect species may be present in certain perovskite materials in reducing environments such as in fuel electrodes of solid oxide cells (SOCs) or on the reducing side of ceramic membranes. A generalized defect thermodynamic model was developed for the triple-conducting perovskites (La,Ba)Fe1−xMxO3−δ (M = Y and Zr) to allow inclusion of the hydride defect formation reaction in addition to the other three main defect reactions, namely, the oxygen vacancy formation, hydration, and charge disproportionation reactions. This comprehensive defect model also allows the incorporation of polynomial functional forms of oxygen nonstoichiometry δ to describe the defect reaction energies and entropies and to enable refinements of the defect reaction equilibrium constants in the defect thermodynamic analysis. As a first step, the developed model is applied to the Ba0.95La0.05FeO3−δ material as an illustrative system to obtain its Brouwer diagrams with both the proton and hydride defects in relevant SOC conditions, particularly for more reducing environments. The results provide direct guidance on the influence of electronic and ionic defect concentrations upon thermodynamic properties and ultimately on the performance of Ba0.95La0.05FeO3−δ and potentially other (La,Ba)Fe1−xMxO3−δ perovskite materials involved in SOC applications.

View access options


Influence of gadolinium doping on structural, optical, and electronic properties of polymeric graphitic carbon nitride

July 2024

·

42 Reads

Polymeric graphitic carbon nitride (gCN) materials have received great attention in the fields of photo and electrocatalysis due to their distinct properties in metal-free systems with high physicochemical stability. Nevertheless, the activity of undoped gCN is limited due to its relatively low specific surface area, low conductivity, and poor dispersibility. Doping Gd atoms in a gCN matrix is an efficient strategy to fine-tune its catalytic activity and its electronic structure. Herein, the influence of various wt% of gadolinium (Gd) doped in melon-type carbon nitride was systematically investigated. Gadolinium-doped graphitic carbon nitride (GdgCN) was synthesized by adding gadolinium nitrate to dicyandiamide during polymerization. The X-ray diffraction (XRD) and transmission electron microscopy (TEM) results revealed that the crystallinity and the morphological properties are influenced by the % of Gd doping. Furthermore, X-ray photoelectron spectroscopy (XPS) studies revealed that the gadolinium ions bonded with nitrogen atoms. Complementary density functional theory (DFT) calculations illustrate possible bonding configurations of Gd ions both in bulk material and on ultrathin melon layers and provide evidence for the corresponding bandgap modifications induced by gadolinium doping.


Growth of Well-Defined Model Catalysts for Electrochemistry: From Surface Science Studies to Electrocatalytic CO2 Conversion

May 2024

·

4 Reads

We combined ultrahigh vacuum (UHV) surface science techniques, electrochemical measurements, and computational modeling to investigate electrocatalytic systems that are crucial components of the carbon management effort, including oxygen evolution reactions (OER) and CO2 reduction reactions (CO2RR). Well-defined Model catalysts were grown on substrates in the UHV chamber and characterized with X-ray photoelectron spectroscopy (XPS) and scanning tunneling microscopy (STM). Selected model electrocatalysts were then tested in electrochemical cells to establish the structure-property relationships in OER and CO2RR. Our results showed that the edge sites of Fe2O3 grown on Au(111) were the most active toward OER and incorporation of Ni at the edge sites (NiFeOx) further boosted their OER activity. We also resolved the size-dependent electrocatalytic CO2-to-CO conversion of the Ag nanoparticle electrocatalysts with average particle diameter between 2 to 6 nm: smaller diameter (< 3 nm) particles favored H2 evolution reaction (HER) due to a high population of Ag edge sites, whereas larger diameter particles favored CO2RR as the population of Ag(100) surface sites grew. We further discovered that electronic interactions between small diameter Ag particles and highly defective carbon supports could break the size-dependent CO2RR selectivity, resulting in highly selective (CO Faradaic Efficiency > 90%) and active Ag nanoparticle electrocatalysts with sizes < 2 nm diameter.


Detection of opioid drugs using a AuNP‐decorated sc‐SWCNT‐based FET device. a) Molecular structures of fentanyl, codeine, hydrocodone, and morphine. b) Schematic illustration of a AuNP‐decorated sc‐SWCNT‐based FET device. c) AFM image of bare sc‐SWCNTs deposited between the source and drain electrodes. d) AFM image of AuNP‐decorated sc‐SWCNTs.
Fentanyl sensing performance of the Au‐SWCNT FET device. a) FET transfer characteristics of the Au‐SWCNT FET device upon adding increasing concentrations of fentanyl from 10 ng mL⁻¹ to 500 µg mL⁻¹. b) Calibration plot for fentanyl detection. The concentration of fentanyl indicated on x‐axis is plotted on a logarithmic scale. The linear range is indicated using a solid line. c) Comparison of sensor responses for fentanyl, codeine, hydrocodone, and morphine. All error bars were calculated from multiple devices. The number of devices (n) used for calculation is indicated in parentheses in the legend.
Results of DFT calculations for the adsorption of opioid molecules on Au‐SWCNTs. a) Comparison of the calculated adsorption energies (kcal mol⁻¹) on Au(111), Au(ad)/Au(111), and AuNP/SWCNT systems. b) Results of DFT calculations illustrating the most stable adsorption configurations on AuNP/SWCNT for: I. fentanyl; II. codeine; III. hydrocodone; IV. morphine Charge difference maps for fentanyl adsorbed on Au/SWCNT at c) the interface between the fentanyl and AuNP‐SWCNT system, respectively, at d) the interface of fentanyl‐AuNP system with SWCNT. The indicated yellow and blue isosurfaces correspond to values of ± 0.01 e⁻ Å⁻³. e) The partial charge density isosurface (0.01 e⁻ Å⁻³) of the electronic levels located within 0.2 eV above the Fermi level. f) Total density of states (DOS) for fentanyl adsorbed on AuNP/SWCNT and the local density of states (LDOS) projections for SWCNT (green), AuNP (red), and fentanyl molecule (blue).
Fentanyl sensing performance of the fentanyl antibody‐functionalized Au‐SWCNT (fentanyl‐ab@Au‐SWCNT) FET device. a) FET transfer characteristics of the fentanyl‐ab@Au‐SWCNT FET device upon adding increasing concentrations of fentanyl from 1 ag mL⁻¹ to 10 ng mL⁻¹. The inset shows the sensor configuration. b) Calibration plot for fentanyl detection. The concentration of fentanyl indicated on x‐axis is plotted on a logarithmic scale. The linear range is denoted using a solid line. c) Comparison of sensor responses for fentanyl, codeine, hydrocodone, and morphine. All error bars were calculated from multiple devices. The number of devices (n) used for calculation is indicated in the parentheses in the legend.
Machine Learning Discrimination and Ultrasensitive Detection of Fentanyl Using Gold Nanoparticle‐Decorated Carbon Nanotube‐Based Field‐Effect Transistor Sensors

April 2024

·

20 Reads

·

1 Citation

The opioid overdose crisis is a global health challenge. Fentanyl, an exceedingly potent synthetic opioid, has emerged as a leading contributor to the surge in opioid‐related overdose deaths. The surge in overdose fatalities, particularly due to illicitly manufactured fentanyl and its contamination of street drugs, emphasizes the urgency for drug‐testing technologies that can quickly and accurately identify fentanyl from other drugs and quantify trace amounts of fentanyl. In this paper, gold nanoparticle (AuNP)‐decorated single‐walled carbon nanotube (SWCNT)‐based field‐effect transistors (FETs) are utilized for machine learning‐assisted identification of fentanyl from codeine, hydrocodone, and morphine. The unique sensing performance of fentanyl led to use machine learning approaches for accurate identification of fentanyl. Employing linear discriminant analysis (LDA) with a leave‐one‐out cross‐validation approach, a validation accuracy of 91.2% is achieved. Meanwhile, density functional theory (DFT) calculations reveal the factors that contributed to the enhanced sensitivity of the Au‐SWCNT FET sensor toward fentanyl as well as the underlying sensing mechanism. Finally, fentanyl antibodies are introduced to the Au‐SWCNT FET sensor as specific receptors, expanding the linear range of the sensor in the lower concentration range, and enabling ultrasensitive detection of fentanyl with a limit of detection at 10.8 fg mL⁻¹.



(Invited) Coupling of Electrochemistry with Surface Science and Computational Modeling to Understand Nanoparticle Electrocatalysts

December 2023

·

14 Reads

ECS Meeting Abstracts

We investigated nanoparticle electrocatalysts in oxygen evolution reaction (OER) and CO 2 reduction reaction (CO 2 RR) via an approach coupling electrochemistry with ultra-high vacuum (UHV) surface science and computational modeling. Well-defined nanoparticle electrocatalysts such as Fe 2 O 3 , NiFeO x , and Ag, were prepared in the UHV chamber via physical vapor deposition (PVD) and characterized with surface science techniques including X-ray photoelectron spectroscopy (XPS) and scanning tunneling microscopy (STM), followed by electrochemistry measurements to establish the structure-property relationships in OER and CO 2 RR. Calculations based on density functional theory (DFT) and microkinetic modeling (MKM) were then implemented to gain fundamental insight into these systems. We observed the edge-enhanced OER activity of Au-supported, ultra-thin Fe 2 O 3 electrocatalysts and revealed incorporation of Ni at the edge sites (NiFeO x ) further boosting their OER activity. We also resolved the size-dependent transition between CO 2 RR and H 2 evolution reaction (HER) selectivity in sub-5 nm Ag electrocatalysts and identified an effective minimum size limit of ∼4 nm for active and selective Ag catalysts: particle diameters below this range experienced increased H 2 evolutions due to a high population of Ag edge sites; and larger diameter particles favored CO 2 RR as the population of Ag(100) surface sites grew but will begin to lose catalyst utilization based on total metal masses.


Defect Thermodynamic Modeling of Triple Conducting Perovskites (La,Ba)Fe 1-x M x O 3-δ for Proton-Conducting Solid-Oxide Cells

August 2023

·

14 Reads

ECS Meeting Abstracts

Density functional theory (DFT) based modeling was performed to investigate the defect energetics of the proton conducting oxides BaFe 1-x M x O 3-δ (with M=Zr, Y, and Zn at the B-site). High-throughput defect modeling was carried out to obtain the defect energetics ensembles as a function of B-site x (0 – 0.25) and oxygen δ (0 – 0.5) compositions for assessment of oxygen vacancy formation (E vac ) and hydration reaction (E hyd ) energies in complex defect-dopant configurations. The modeling effort aims to elucidate the complex coupling among oxygen vacancies, the protons and metal dopants in various concentrations and their combined effect upon defect energetics in BaFe 1-x M x O 3-δ perovskite. This information complements the current experimental measurements with essential data related to the energetics of various defect configurations that can play a role in solid-oxide-cell (SOC) applications. Preliminary results among more than 3,000 configurations investigated are summarized in Figure 1. All calculations were performed with the Perdew-Burke-Ernzerhof (PBE) functional in charge neutral 2 × 2 × 2 perovskite supercells. An effective Hubbard U eff correction of 4 eV was applied to the Fe 3d shell, with all the Fe atoms being in a high-spin state and having a ferromagnetic ordering. Stability analysis of each of the BaFe 1-x M x O 3-δ systems investigated at specified x and δ composition values reveals that the oxygen vacancy (V O ) is energetically unfavorable when placed adjacent to the M dopant. Overall, the Fe-V O -Fe configurations are identified to be more stable than the Fe-V O -M and M-V O -M configurations, while one with a M-V O -M cluster arrangement is the least stable. Such a short-range ordering leads to the blocking of the oxygen vacancy migration toward the oxygen sites around the M dopant, whereas the proton hopping between the oxygens around the M dopants is energetically not hindered. Based on the stability of various defect-dopant configurations of the BaFe 0.75 M 0.25 O 2.875 systems with or without hydration, it is revealed that the stability range of analogous configurations with varying dopant types can be an indicator of the defect-dopant ordering strength and can be correlated with the ionic radii of the dopants in the BaFe 1-x M x O 3-δ systems. Taking the energy of the most stable BaFe 1-x M x O 3-δ structures at various x and δ values analyzed as representative for the respective compositions, the corresponding derivations of the E vac and E hyd energies as a function of δ (and x) are summarized in Figures 1(a) and 1(b), respectively. Overall, the results indicate a stronger oxygen non-stoichiometry (δ) dependence in the E vac values of the BaFeO 3-δ and BaFe 1-x M x O 3-δ systems (M=Zr, Y, and Zn at x=0.125 and x=0.25), whereas the E hyd data exhibit a much weaker δ dependence. This contrasting behavior in the δ dependences between E vac and E hyd values may be attributed to the charge transfer characteristics of the oxygen vacancy formation reaction, which is associated with the Fe 3d and the O 2p energy levels of the materials at various δ values while the charge transfer involved in the hydration reaction is relatively minimal. The effect of dopant concentration (x) further shows an opposite trend in the E vac values vs. x with the dopant types, with an increase of E vac upon increasing the Zr doping concentration and a decrease of E vac upon increasing the Y and Zn doping content. Due to the strong oxygen non-stoichiometry dependence in the E vac values, the results indicate an increased proton affinity of the BaFe 1-x M x O 3-δ oxides. Further assessments of the E vac and E hyd data on δ values and their impact upon the defect equilibria of the BaFe 1-x M x O 3-δ systems under the SOC operating conditions will be discussed. Disclaimer This project was funded by the U.S. Department of Energy, National Energy Technology Laboratory, in part, through a site support contract. Neither the United States Government nor any agency thereof, nor any of their employees, nor the support contractor, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof. Figure 1


Defect Thermodynamic Modeling of Triple Conducting Perovskites (La,Ba)Fe 1-x M x O 3-δ for Proton-Conducting Solid-Oxide Cells

May 2023

·

12 Reads

·

1 Citation

ECS Transactions

Both the experimental and first-principles modeling results revealed the dependence of defect energetics on oxygen non-stoichiometry and magnetic coupling of Fe in the Fe-based perovskite oxides. A generalized defect thermodynamic model of the proton-conducting (La,Ba)Fe 1-x M x O 3-δ perovskite oxide is developed to allow inclusion of nonlinear δ dependent terms in three key defect reaction energies, namely, the oxygen vacancy formation, hydration, and charge disproportionation reactions. A transition from a large polaron description at lower δ values to a small polaron expression at higher δ is also considered in our analysis. Based on the functional forms of defect energetics on δ as guided by first principles modeling and literature data, the Brouwer diagrams of BaFe 0.9 Y 0.1 O 3-δ are assessed to provide information on electronic and ionic defect concentration (including the proton species) as a function of O 2 and H 2 O pressure at different temperatures for solid-oxide cell applications.


Citations (77)


... In the PXRD of g-C 3 N 4, in addition to the two distinct peaks of g-C 3 N 4 , other peaks are also detected, which are recognized as carbon nitride compounds resembling amorphous graphite. These compounds may have originated from the incomplete polymerization of dicyandiamide (Jiang et al. 2017;Li et al. 2020;Kesavan et al. 2023). ...

Reference:

Recyclable g-C3N4 and K-doped g-C3N4 pellets for the photocatalytic production of H2O2 under direct sunlight
Optimizing Dicyandiamide Pretreatment Conditions for Enhanced Structure and Electronic Properties of Polymeric Graphitic Carbon Nitride
  • Citing Article
  • January 2023

Journal of Materials Chemistry C

... [21][22][23] Therefore, in this work, the hydride defect formation reaction is incorporated in the (La,Ba)Fe 1−x M x O 3−δ (M is a nontransition metal cation such as Zr and Y) defect model. 24 Such defect species were previously speculated to exist in proton conducting oxides operated in reducing conditions. 21 Furthermore, controlling the material synthesis to form a single perovskite oxyhydride phase that contains substantial lattice hydride species has been shown to be a feasible approach. ...

Defect Thermodynamic Modeling of Triple Conducting Perovskites (La,Ba)Fe 1-x M x O 3-δ for Proton-Conducting Solid-Oxide Cells
  • Citing Article
  • May 2023

ECS Transactions

... 20 These results further indicate the possibility of hydrogen incorporation from the surfaces into the perovskite lattice to form hydride defect species under reducing conditions. [21][22][23] Therefore, in this work, the hydride defect formation reaction is incorporated in the (La,Ba)Fe 1−x M x O 3−δ (M is a nontransition metal cation such as Zr and Y) defect model. 24 Such defect species were previously speculated to exist in proton conducting oxides operated in reducing conditions. ...

Defect Thermodynamics and Transport Properties of Proton Conducting Oxide BaZr1−xYxO3−δ (x ≤ 0.1) Guided by Density Functional Theory Modeling
  • Citing Article
  • October 2022

JOM: the journal of the Minerals, Metals & Materials Society

... 12 For example, zeolites show microporosity and very high surface-to-volume ratios, 9 reducing their water uptake. Metal-organic frameworks (MOFs) based on zirconium, [13][14][15][16][17][18] aluminium, [19][20][21] zinc, 10,22 chromium, [23][24][25] magnesium, 26 copper, 27 and iron 28 have also been studied. This class of porous, hybrid metal-organic materials, frequently possess tuneable porosity, significant surface areas and ease of functionalisation. ...

Crystallographic Mapping and Tuning of Water Adsorption in Metal-Organic Frameworks Featuring Distinct Open Metal Sites
  • Citing Article
  • October 2022

Journal of the American Chemical Society

... Specific heat is typically expressed in the form of a cubic function of temperature. 62 However, to remain consistent with MD calculations, a constant classical value of specific heat was used in the calculations performed herein, ...

Toward Addressing the Challenge to Predict the Heat Capacities of RDX and HMX Energetic Materials
  • Citing Article
  • March 2022

Propellants Explosives Pyrotechnics

... They found a reduction in adsorption-free energy with an increase in tensile strain. The effect of 5 % tensile strain along the (Zg,x) direction reduced adsorption free energy from 1.698 eV(un-strained) to 1.550 eV, and from 1.698 eV(un-strained) to 1.355 eV along the (Zg,d) direction 70 . Other Janus materials such as MoXY (X, Y = O, S, Se, and Te) 71 , PtSSe 72 , and XM 2 Y (X, Y = S, Se, Te; M = Ga, In) 73 have also been theoretically investigated for photocatalytic water splitting as their bandgaps belong to visible range. ...

Real-Time Modulation of Hydrogen Evolution Activity of Graphene Electrodes Using Mechanical Strain
  • Citing Article
  • February 2022

ACS Applied Materials & Interfaces

... 39,40 GO structure resembles that of a polymeric phenolic compound. 41 Therefore, its oxidation (vide infra) could produce lignin-like degradation by-products with similar toxicity to the fungus. An accumulation of phenolic compounds derived from progressive GO degradation would explain why, for the same incubation time, the higher the initial GO concentration in liquid media, the lower the "growth-stimulating effect". ...

Composition and Structure of Fluorescent Graphene Quantum Dots Generated by Enzymatic Degradation of Graphene Oxide
  • Citing Article
  • June 2021

The Journal of Physical Chemistry C

... UiO-66-NH 2 nanoparticles were prepared according to the previous literature [31]. To synthesize UiO-66-NH 2 /SA, 1.5 g of UiO-66-NH 2 was dispersed in 50 mL of water, followed by the addition of 1 g of sodium alginate after ultrasonication. ...

Heterogeneous Growth of UiO-66-NH 2 on Oxidized Single-Walled Carbon Nanotubes to Form “Beads-on-a-String” Composites
  • Citing Article
  • March 2021

ACS Applied Materials & Interfaces

... To address this gap, numerous atomic-scale studies of the hydrogen behavior in the bulk of monoclinic zirconia have been conducted [22][23][24][25][26]. The studies focusing on the hydrogen behavior in tetragonal zirconia [25,27,28] or monoclinic HfO 2 [24,29] are also interesting points of comparison in this context. All these studies help to understand how hydrogen interacts with the point defects of the bulk under various environments, such as reductive or oxidizing conditions. ...

Density functional theory modeling of cation diffusion in tetragonal bulk Zr O 2 : Effects of humidity and hydrogen defect complexes on cation transport

Physical Review Research