Next Article in Journal
Anti-Inflammatory Effects of 6-Methylcoumarin in LPS-Stimulated RAW 264.7 Macrophages via Regulation of MAPK and NF-κB Signaling Pathways
Previous Article in Journal
Simultaneous Enantiomeric Separation of Carfentrazone-Ethyl Herbicide and Its Hydrolysis Metabolite Carfentrazone by Cyclodextrin Electrokinetic Chromatography. Analysis of Agrochemical Products and a Degradation Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbon Nanostructures Doped with Transition Metals for Pollutant Gas Adsorption Systems

by
J. M. Ramirez-de-Arellano
1,
M. Canales
2 and
L. F. Magaña
3,*
1
Escuela de Ingeniería y Ciencias, Tecnologico de Monterrey, Av. Eugenio Garza Sada 2501, Monterrey 64849, Mexico
2
Universidad Autónoma Metropolitana Unidad Azcapotzalco, Av. San Pablo Xalpa No. 180, Colonia Reynosa Tamaulipas, Delegación Azcapotzalco, Ciudad de México 02200, Mexico
3
Instituto de Física, Universidad Nacional Autónoma de Mexico, Apartado Postal 20-364, Ciudad de México 01000, Mexico
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(17), 5346; https://doi.org/10.3390/molecules26175346
Submission received: 28 July 2021 / Revised: 28 August 2021 / Accepted: 31 August 2021 / Published: 2 September 2021
(This article belongs to the Section Applied Chemistry)

Abstract

:
The adsorption of molecules usually increases capacity and/or strength with the doping of surfaces with transition metals; furthermore, carbon nanostructures, i.e., graphene, carbon nanotubes, fullerenes, graphdiyne, etc., have a large specific area for gas adsorption. This review focuses on the reports (experimental or theoretical) of systems using these structures decorated with transition metals for mainly pollutant molecules’ adsorption. Furthermore, we aim to present the expanding application of nanomaterials on environmental problems, mainly over the last 10 years. We found a wide range of pollutant molecules investigated for adsorption in carbon nanostructures, including greenhouse gases, anticancer drugs, and chemical warfare agents, among many more.

1. Introduction

Since their discovery in 1985, the study of fullerenes and their exciting properties led to the subsequent discovery of many other carbon nanoforms: nanotubes, carbon onions, graphene, graphdiyne, carbon nanotori, etc., including hybrid nanostructures [1,2,3,4,5,6], see Figure 1 and Figure 2. As a result, there are numerous practical applications of carbon nanostructures, like gas sensors, adsorption of pollutant molecules, electric batteries, hydrogen adsorption, electronic devices, etc.
The capture of gases like CO, CO2, N2O, CH4, and many others is urgent. CO comes mainly from the non-complete combustion of fuels that contain carbon, and it is toxic and an asphyxiant. On the other hand, plenty of industrial processes in our world generate greenhouse gases like CO2, N2O, and CH4. Thus, humankind’s climate change and ecological challenges have increased the interest in carbon nanostructures for fast green applications [7].
We should mention that the specific surface area of graphene is 2630 m2/g, more significant than that of zeolites (between 500 and 800 m2/g) although smaller than that of metal-organic frameworks (or MOFs), which is around 7000 m2/g.
Thus, carbon nanostructures’ sizeable specific surface area is an excellent appeal for gas sensing and gas adsorption systems. Furthermore, the adequate doping of these structures may increase their adsorption capability and tune their selective adsorption. In particular, the reactivity of transition metals makes them ideal dopant candidates. Investigations based on the density functional theory (DFT) have been crucial to explore the adsorption properties of the different carbon nanostructures. Some researchers studied the variability in results obtained from diverse DFT implementations. For instance, predictions made using the Augmented Plane Waves plus local orbitals method (APW+lo) and the Projector Augmented Wave method (PAW) are identical for practical purposes [8]. Most of the computational studies reviewed in this paper also report the density of states (DOS) spectrum to complement their investigations on carbon nanostructures. They also include Mulliken or Lowdin charge analysis, electron density differences, HOMO and LUMO analysis, etc. In general, DFT is very efficient in computing interaction energies, while the Grand Canonical Monte Carlo (GCMC) calculations are adequate to predict adsorption capacities.
This work aims to review the most important and recent developments related to the adsorption of pollutants or hazardous gases by using carbon nanostructures. Among the most investigated pollutant gases, we must mention carbon monoxide (CO), carbon dioxide (CO2), methanol (CH3OH), methane (CH4), nitrogen monoxide (NO), nitrogen dioxide (NO2), ozone (O3), and formaldehyde (CH2O) with many others. However, we found some works on hydrogen adsorption on carbon nanostructures, which is essential to serve as a non-pollutant fuel. Although, we should mention that the number of works on this subject in the last 10 years is considerably less than previously.
We covered a 10-year range approximately, not in an exhaustive way, but describing the state-of-the-art up to date. When results were considered relevant enough, or when we believed that they would add important information for the context of the topic covered, we also included works dating from earlier years.
In the second section, we review the adsorption of molecules on undoped or nonmetal functionalized carbon nanostructures. We considered the subsections of nanotubes, graphene, fullerenes, graphdiyne, and hybrid systems. The third section reviews the same carbon structures as in the second section but doped with transition metals. Our conclusions are in the fourth section.

2. Adsorption of Molecules on Pristine or Nonmetal Functionalized Systems

2.1. Nanotubes

Various experimental studies focus on the gas adsorption capabilities of single-walled carbon nanotubes (SWCNTs) and multiple walled carbon nanotubes (MWCNTs), showing a good agreement with the Langmuir adsorption model. The nature of the gas sensing mechanism in nanotubes would be due to charge transfer between them and the molecules adsorbed, or due to effects of the carbon nanotube network, for devices made of bundles [9].
Several works using first-principles calculations have also shed light on this topic. Numerical DFT simulations using the local density approximation (LDA) for the exchange-correlation potential show that most gas molecules adsorb weakly on pristine SWCNTs. Still, certain gases like NO2 and O2 are more sensitive to be adsorbed due to the electronic properties of the former [10]. Although LDA is usually inaccurate to describe long-range interactions, the molecule-tube interaction seems to be between the LDA and the generalized gradient approximation (GGA). Hydrogen adsorption is another broad field of interest due to its potential applications as a non-pollutant fuel. Experimental studies suggest that fold-structure adsorbents convey better H2 adsorption than pore or flat structures [11]. It is common to characterize the carbon material by N2 adsorption at a given temperature and correlate that to their hydrogen storage capacity. SWNTs perform below activated carbon samples for hydrogen uptake, but the process is reversible in both cases. The storage capacity of the latter is about 4.5 wt% at 77 K [12]. Functionalizing MWCNTs using acids (F-MWCNTs) or by KOH activation (A-MWCNTs) increases the H2 gas sensing response [13,14].
The adsorption uptake of CO2 in pristine SWCNTs is more significant than that of CO, CH4, N2, and H2, a result that is confirmed both experimentally and by Monte Carlo simulations [15]. Experimental results show that carbon nanotubes, when modified by 3-aminopropyl-triethoxysilane (APTS), adsorb CO2 better. This effect decreases with temperature and increases with water content in the air. MWCNTs modified by APTS have good adsorption performance at 20 °C. The interaction between molecules and CNTs-APTS implies physisorption and chemisorption that shows stability for a cyclic operation. This operation is enhanced using saturated water vapor in the gas stream, giving CO2 deconcentration of around 67% [16,17,18]. Carbon nanotubes films have also been studied as CO2 sensors, both experimentally and via grand canonical Monte Carlo (GCMC) simulations [19].
Some other ways of increasing the adsorption capacity of CO2 are by functionalizing carbon nanotubes with nitrogen groups (N-MWCNTs); by fabricating polyethylenimine-polyethylene glycol/MWCNTs bilayer-structure devices—obtaining CO2 desorption without heating; or by a two-step modification of MWNTs using a mixture of diluted nitric and sulfuric acid and then 1,3-diamino propane [20,21,22].
The diameter and chirality of CNTs affect the adsorption. Flexible SWCNTs would adsorb up to 35 wt% of CO2, as studied using molecular dynamics (MD) calculations, including ad hoc potentials built to model the intramolecular interactions within the SWCNT [23].
The selectivity of CO2/CH4—i.e., good adsorption of CO2 combined with the exclusion of CH4—also increases by this functionalization, an effect that reaches an optimal value at lower pressures and room temperature, which is essential for industrial gas separation. Oxygen-rich surface functional groups attached to the surface of MWCNTs also increase the CO2 adsorption capacity [20,24]. MWCNTs physisorb methane (CH4) and natural gas showing fast sorption kinetics, making MWCNTs a promising porous media for natural gas storage [25]. Applying high pressures to double-walled carbon nanotube arrays increases the CO2 and N2 gas adsorption, an effect that is also obtained by functionalizing the tubes via oxygen plasma treatment [26]. GCMC simulations also indicate that water molecules can increase the selectivity of CO2 when CO2/CH4 and H2O/CO2/CH4 mixtures are considered in CNTs and silicon carbide-derived carbon (SiC-DC) [27].
Sulfur hexafluoride (SF6) is another greenhouse gas, considered several orders of magnitude more nocive than CO2, measured over long periods. MWNTs modified via H2SO4/H2O2 oxidation, or KOH activation, are suitable SF6 adsorbents [28].
The adsorption of acetone (C3H6O) is another issue extensively studied. This substance is liquid at room temperature, and it has a low boiling point of 329.20 K or 56.05 °C [29]. SWCNTs can strongly adsorb acetone molecules and then desorb them with increasing temperature within the interval of 400–900 K. For large-diameter SWCNTs, adsorption can happen in the interstitial channels, the bundles. The elastic deformation of the tubes can also increase the adsorption of acetone molecules [30]. Similar studies have been performed combining experimental and DFT methods to study the ethanol (C2H5OH) sensing properties of CNTs [31]. Polyaniline functionalized multiwalled carbon nanotubes (PANI/MWCNTs) have also been experimentally shown to detect ammonia (NH3) gas [32].
There are other ways of increasing the sensing performance of CNTs, besides deformation. For instance, sorted semiconducting nanotubes are more sensitive to pollutant gases like NO2 and NH3 than unsorted tubes [33]. For nitric oxide (NO) adsorption, in situ ultraviolet (UV) light illumination of pristine carbon nanotubes can increase their performance. The cleaning of the surface of the tubes by UV light would improve their gas sensing capacity [34]. Chemical polymerization with pyrrole (PPy) to obtain SWNTs/PPy nanocomposites can also enhance their sensitivity 10 times [35].
We found investigations of recent years on the adsorption of other gas molecules, like Xe, natural gas, or the nitrate ion NO3. The adsorption of Xe by single-walled carbon nanotubes (SWCNTs) has been studied experimentally and using ab initio calculations. Closed SWCNTs can be opened by oxidation at their ends and wall defect sites using ozone. This opening enhances the adsorption of Xe, as the nanotube etching increases wall openings at an optimal radius of 5–7 Å. After these optimal values, the adsorption rate drops [36]. SWCNTs can also adsorb a nitrate ion NO3 in its gas phase [37].
Carbon nanotubes can also adsorb formaldehyde (HCHO) molecules in their interior and exterior walls, with preferential adsorption on the latter. The adsorbed HCHO molecules alter the electronic structure of the CNNTs, reducing the HOMO/LUMO gap from its original value of 4.02 to 2.44 eV, according to ab initio calculations [38]. These tubes can adsorb mixtures of SO2/N2 at the equimolar ratio, with MD calculations showing an increase in SO2 adsorption as the CNT diameter increases [39]. Finally, experimental studies have demonstrated the feasibility of MWCNTs as adsorbents of several volatile organic compounds (VOCs) such as benzene, methanol, ethanol, acetone, etc. [40].
Table 1 summarizes the adsorption energies—for the studies that reported them—and the characteristics of investigations covered in this section. When the study considers different variations or adsorption conditions for the same adsorbate, an energy range is given instead of a particular value.

2.2. Graphene

Typical calculations of molecules adsorption on graphene consider either a 4 × 4 or a 5 × 5 unit cell to avoid spurious interactions, as many codes work with periodic boundary conditions. The three adsorption sites considered in these studies are (a) at the bond between two C atoms or a B-site; (b) at the top of a C atom or T-site; (c) over the center of a hollow hexagon of C atoms, or H-site. Most DFT calculations work with a graphene monolayer, but some studies consider two or more sheets.
The adsorption of atomic or molecular hydrogen on graphene has attracted much interest in the last decade. Graphene can adsorb H atoms with binding energies per atom of about 0.8–1.9 eV [42]. There are predictions of theoretical calculations that hydrogenated graphene sheet would be a semiconductor. Deuterium (D) has been incorporated into graphene via thermal annealing in a process above 400 °C, which is not entirely reversible [43] to verify those predictions. Defects in graphene can also improve H2 adsorption. Defective graphene V222 is obtained by removing a C atom from the graphene sheet and then di-hydrogenating the three C atoms on the vacancy edges. MD calculations using a V222 structure show an H2 adsorption–desorption process that is more reversible than pristine graphene [44].
It is frequent to find graphene used as a substrate to adsorb H2O, NH3, CO, NO2, and NO, as explored via DFT ab initio calculations and experimentally. It is well established that pristine graphene adsorbs H2O molecules, with the latter acting as an acceptor. NO2 can also be physisorbed on pristine graphene, and it can induce more significant doping than NO [45,46]. Pristine graphene can physysorb NO molecules with electrons transferred from the first to the second, with the B-site being the most stable adsorption site [47]. It can also weakly adsorb hydrogen sulfide (H2S) and methane (CH4) [48]. NonTM-functionalized graphene can also detect ammonia (NH3). A way to increase the sensing performance of graphene is to produce pristine graphene noncovalently functionalized, usually by a biocompatible stabilizer such as flavin mononucleotide sodium salt [49].
From experimental and theoretical investigations, we find that graphene can adsorb organic molecules such as acetone, acetonitrile, dichloromethane, ethanol, ethyl acetate, hexane, and toluene. Graphene modeled as coronene would interact with said molecules primarily by London dispersive forces [50]. Table 2 summarizes the main results reviewed in this section.

2.3. Fullerenes

Fullerenes can be obtained directly from graphene [61]. They are the subject of a great deal of experimental and theoretical research due to their exciting adsorption properties given by their characteristics. From first-principles studies we know that the size of the fullerene impacts its stability and its hydrogen storage capacity. The cohesive and formation energies of fullerenes decrease with increasing size [62].
Furthermore, fullerenes adsorb gases like N2, Ar, and CO2, and for the latter, an exceptionally large C460 fullerene would adsorb up to 28 wt% at room temperature [62]. A P-doped fullerene can adsorb the CO2 molecules; such doping increases the selectivity of CO2 concerning an N2/CO2 mixture, as shown by DFT studies. Additionally, an electric field applied to the systems affects the interaction between the P-doped-CO2 and the CO2 molecule, going from physisorption to chemisorption [63].
Monte Carlo simulations show that C60 fullerenes can also adsorb ethylene (C2H4) at 150 K [64]. Experimental results show that self-assembled C60 crystals have excellent sensing properties to detect toxic aromatic solvent vapors such as aniline, toluene, benzene, ethanol, hexane, cyclohexane, and methanol [65].
Additionally, DFT studies showed that fullerene-based devices reduced nitrous oxide (N2O) and carbon monoxide (CO) pollution. In that case, Si-coordinated nitrogen-doped C60 fullerenes, labeled as Si@C54N4, catalyzes the N2O reduction and CO oxidation in the presence of an O2 molecule [66]. In [67], the authors found that metal oxides (MOxs) such as Cu2O, ZnO, and NiO, when adsorbed on C60 fullerenes, can adsorb nitrogen dioxide (NO2) as well as CO much better than C60 alone. This is caused by a more significant charge transfer, more considerable adsorption energies, and more extensive enthalpy changes when compared to those of the pristine C60 fullerene.
Such pristine C60 fullerenes can adsorb H2, and their interaction has been studied by several means, including a modification of the L-J potential and the continuum approach [68]. When C60 fullerenes oxidize at 400 °C and a pressure of 2 bar, they enhance their surface and increase H2 adsorption. Experimental studies showed that this process increases the H2 adsorption by a factor of three, at 77 K and a pressure of 120 bar [69]. When C60 or C70 fullerenes are doping helium droplets, they can physisorb both hydrogen H2 and deuterium D2 [70]. Furthermore, in [62], the authors showed that the previously mentioned C460 fullerene would reach H2 storage of 7.60 wt% at 300 K.
There are theoretical investigations to compare the efficiency of the porous structure of fullerenes with that of similar hexagonal boron nitride (hBN) nanocages. In [71], the authors showed that a B16N16 fullerene-like form would have low conductivity and reactivity compared with the fullerene C32 and a B-doped fullerene labeled as B8C24. They concluded that the B16N16 system was the most suitable H2 adsorbent of these three cases.
Other theoretical calculations using DFT show the adsorption of nitrogen gas (N2) molecules by different fullerenes such as sphere-like C82 or the tubelike C110. The interaction between N2 and such fullerenes can lead to (N2)n molecular clusters [72]. Similar studies show that NO and NO2 molecules can be better adsorbed by a C60 fullerene when the latter is B-, N-, or BN-codoped. The best adsorption performance is obtained with the BN-codoped C60 fullerene, labeled as C58BN, and obtained after doping by substitution of two carbon atoms by nitrogen and boron atoms [73].
There are clusters formed with rare gas atoms with C60XN where X = Ne, Ar, Kr, or Xe. A study using an empirical approach to the potential energy surface explored the X-C60 and X-X interactions, showing that there is generally an energetically favorable cluster formation, with N ranging from 32 for Xe up to 60 for Ne [74]. By itself, a C60 fullerene can also adsorb He molecules [68].
The detection and sensing of chemical warfare agents (CWAs) is a research line where fullerenes play a relevant role. The tabun nerve agent (C5H11N2O2P), a highly toxic and dangerous gas, is an example of this, as DFT calculations have shown that a C20 fullerene can adsorb it. Doping the fullerene with either boron or nitrogen affects its interaction with Tabun, showing that the B-doped fullerene, or C19B, is better to detect the hazardous gas, due to a larger change of the system’s electrical conductivity [75].
Cyanogen (C2N2) is another toxic gas that fullerenes could detect. DFT/B3LYP calculations show that OH-functionalizing or Ge-doping of C60 fullerene result in a larger adsorption capacity than a pristine C60 fullerene, although the increase is relatively small [76]. Experimental investigations show that C60 fullerenes can adsorb the amphetamine (AA) drug. Additionally, DFT calculations show that the AA-C60 interaction increases the fullerene by substituting C atoms with either Si or Ge atoms. In both cases, the AA drug tends to be adsorbed at the (6, 6) bonds of the C60 structure, i.e., the bonds joining two hexagons in the fullerene structure [77].
At a larger scale, fullerenes help to detect and adsorb industrial waste. For instance, a functionalized magnetic fullerene nanocomposite (FMFNc) obtained employing a thermal decomposition of polyethylene terephthalate (PET) bottles (environmentally convenient decomposition) is an excellent candidate to purify dye-contaminated wastewater [78]. Table 3 shows the main quantitative results of this section.

2.4. Graphdiyne

Graphdiyne—in some works abbreviated as GDY—is a recently developed 2D material that has raised interest because it is a metal-free catalyst for low-temperature CO oxidation. It is helpful to absorb CO molecules. According to spin-polarized DFT calculations, graphdiyne can also absorb O2 and does it more quickly than CO adsorption [81]. However, pristine graphdiyne does not capture N2 efficiently [82].
In [83], the authors used DFT simulations and found that CO and methanol (CH3OH) are physisorbed by the pristine GDY surface, while a Ca-decorated graphdiyne (Ca-GDY) can further increase their adsorption. Ca-GDY can adsorb up to 29.81 wt% of CH3OH and 27.10 wt% of CO.
A GDY monolayer also adsorbs atomic oxygen and hydrogen at different sites. Hydrogen chemisorbs above a carbon atom and oxygen at the hollow place. According to such studies, oxygenated graphdiyne could be useful for spintronic devices [84].
Graphdiyne is helpful to detect amino acids, such as glycine (C₂H₅NO₂), glutamic acid (C5H9NO4), histidine (C6H9N3O2), and phenylalanine (C9H11NO2). MD simulations at room temperature suggest that graphdiyne adsorbs such amino acids with an energy more significant than that on graphene, with dispersion interactions predominating in the process [85]. Graphdiyne can also adsorb dimethylamine (DMA) and—to a lesser level—trimethylamine (TMA) vapor molecules and desorb them in a short time, making it suitable for sensing devices [86].
Other studies show that graphdiyne is a helpful drug delivery agent for medical applications. DFT and Quantum-Monte-Carlo simulations suggest that GDY adsorb the drugs sorafenib (C21H16ClF3N4O3) and regorafenib (C21H15ClF4N4O3) then release them upon protonation [87].
Graphdiyne sheets adsorb ammonia molecules (NH3) according to ab initio calculations, showing a better adsorption capacity than graphdiyne nanotubes [88]. Formaldehyde (HCHO) and formic acid (HCOOH) in vapor form can also be adsorbed on the top, ring and bridge sites of a graphdiyne nanosheet [89].
Functionalizing hydrogen-substituted GDY nanostructures with pyridinic nitrogen enhances its electrocatalytic performance for oxygen reduction. Hydrogen-substituted graphdiyne (HsGDY) results from bonding hydrogen to three of the C atoms in a benzene ring, facilitating nitrogen doping [90]. HsGDY and B-decorated HsGDY has also been able to physisorb several gas molecules such as NO, NO2, NH3 and N2. The optimal adsorption positions for these cases were located in the vicinity of the GDY benzene rings. The B-decorated HsGDY had a better adsorption performance, in particular for NO and NO2 [91]. Boron decoration on a regular graphdiyne layer (BGDY) has also been considered as a possible electronic sensor for anticancer drugs like temozolomide (TMZ). The B decoration was also shown to increase the TMZ adsorption capabilities of GDY, related to an increase in its electrical conductivity of around 40%, according to ab initio calculations [92]. The 5-fluorouracil (5FU) anti-cancer drug—which shows no interaction with pristine GDY—can also be adsorbed by a BGDY composite, increasing the latter’s electrical conductivity by around 25%, a feature that would make BGDY suitable for 5FU sensing [93].
Besides boron, other nonmetal atoms like Si, P, S, As, Se, and Te are used through DFT calculations to decorate graphdiyne surfaces to increase their ability to dissociate molecular oxygen. The As- Se- and Te-decorated GDY monolayers physisorbed the O2 molecule, while the B, N, O, Si, P, and S-decorated GDY chemisorbed it [94].
Li, an alkali metal, is another exciting option. The authors in [95,96] showed that this metal catalyzes the adsorption properties of graphene or carbon nanotubes. Studies have compared DFT-PBE and hybrid DFT+LC-⍵PBE calculations to explore the Li adsorption on graphdiyne. PBE tends to overestimate the adsorption energy, suggesting that hybrid DFT should be preferred when studying GDY electronic properties [97]. Similar studies, including Li-decorated GDY, or Li@GDY, show that it could capture CO2 molecules while investigating the possible conversion of CO2 into beneficial—or at least less nocive—chemicals via carbon dioxide electrochemical reduction reaction (CO2RR) [98].
In [99], the authors found that CO2RR improved with nitrogen doping on graphdiyne (NGDY), increasing the selectivity regarding the doping with CH3OH and CH4. In particular, the lowest limiting potential of CH3OH on NGDY gets reduced. Further composites such as boron- and nitrogen-doped GDY anchoring a single Cu atom (labeled as Cu@N- and Cu@B-doped GDY) have also been systematically explored via ab initio calculations. The result is that Cu@N-doped GDY monolayers are more efficient than the boron-doped ones for CO2 reduction and highly catalytic activity toward CO2RR [100].
However, there are cases where boron-doped GDY shows a better performance than nitrogen-doped GDY, as demonstrated by spin-polarized DFT studies. For instance, B-GDY shows excellent sensitivity and selectivity toward NO, NO2, and ammonia (NH3) [101]. Recently, a DFT study combining boron and nitrogen doping—a BN co-doping labeled as BN@GDY—of defective graphdiyne showed that it could increase the catalytic efficiency compared with boron doping alone. The increment in catalytic efficiency would be by a change in the B hybridization from sp2 to sp3 caused by the introduction of N atoms [102].
Improving catalysis can be seen the other way around: instead of decorating graphene sheets, they could be the decoration. For instance, when placed on the Pt(111) surface, a graphdiyne layer gives better results at increasing the latter two-dimensional confined catalysis performance compared with graphene or hexagonal boron nitride [103].
Finally, an increasing interest in the effective detection of chemical warfare agents (CWAs) has led to the investigation of graphdiyne as a possible sensor. Using ⍵B97XD DFT and Quantum theory of atoms in molecule (QTAIM) analysis, studies in [104] found that G-type nerve agents like GA(tabun), GB(sarin), GD(soman), and GF(cyclosarin) can be physisorbed at the triangular section of a GDY nanosheet. The GDY surface shows a short recovery time at room temperature, suggesting it can be a good option for sensor devices. Highly nocive A-series CWAs can also be physisorbed by GDY nanoflakes via noncovalent adsorption, as per DFT-⍵B97XD calculations that consider long-range interactions [105]. Lewisite molecules are another type of toxic CWAs that DFT calculations show can be adsorbed by graphdiyne nanoflakes. In this case the study shows that GDY physisorbs Lewisite L1, L2 and L3 with the highest GDY sensitivity being found to be towards L2 [106]. Table 4 shows the main quantitative results of this section.

2.5. Hybrid Systems

We define a hybrid system as the surface created using different carbon nanostructures, i.e., the combination of graphene or graphdiyne with nanotubes or fullerenes or semi fullerenes or buckybowls. The result is a set of structures with a larger specific area for gas adsorption purposes. For instance, theoretical investigations use DFT focused on the hydrogen adsorption on a system made of graphene decorated with C176. The results show that the adsorption energies ranged between 0.069 and 0.115 eV. Up to four hydrogen molecules could get into the C176 [109].
Another frequently investigated structure for gas adsorption is pillared graphene, which comprises graphene sheets joined by carbon nanotubes. In [110], the authors investigated the hydrogen storage capacity in pillared graphene using MD calculations to study the effects of the environment’s pressure, temperature, and geometric structure. They found that pillared graphene has a higher performance than CNTs as a hydrogen storage material, a fact that due to its larger surface area.
Using a hybrid molecular dynamics–Monte Carlo simulation method, the authors in [111] investigated the methane (CH4) adsorption on Pillared graphene. They found that in this case, the CNT length has the most significant effect on the adsorption amount of methane among different geometrical parameters such as CNT diameter, graphene sheet layer spacing, and the number of CNTs. The adsorption ability of pillared graphene is greater than that of graphene sheets. Such a structure has higher mechanical stability than graphene.
A variation of the pillared graphene system considers fullerenes instead of nanotubes placed between the graphene sheets. In [112], the authors considered this variation and studied methane adsorption with adjustable micro and mesoporous morphology. Their simulations via grand canonical Monte Carlo simulations revealed that this system is suitable for methane storage. Other studies performing similar simulations have explored the same kind of nanostructure for hydrogen adsorption [113]. They considered three different fullerenes such as C180, C320, and C540. Hence, this would be a potentially helpful system for hydrogen storage.
Other studies involving grand canonical Monte Carlo simulations considered fullerenes placed between graphene sheets and lithium doping to investigate methane (CH4) adsorption. These structures are labeled as Sandwiched Graphene-Fullerene Composite (SGFC) and have a promising potential for methane storage applications. This Li-doped nanostructure could be considered suitable host materials for lightweight methane storage devices [112]. The same SGFC, both undoped and Li-doped, has been explored in terms of its hydrogen adsorption capacity. Employing grand canonical Monte Carlo calculations, three fullerene types as the sandwich core have also been studied: C180, C320, and C540. It was observed that a Li-doped nanostructure with a doping ratio of Li:C = 1:8 can overpass the gravimetric capacity of 5%, while an undoped one can reach the value of 3.83% at 77 K and 1 bar. The Li-doped SGFC capacity is larger than that observed in graphene-based systems [113].
Carbon nano onions (CNO) are another interesting mixed nanostructure discovered in 1992. These systems are multi-shell fullerenes with concentric shells of carbon atoms with an interlayer distance similar to graphite [114]. Experimental investigations on the synthesis of hollow carbon nano-onions and their use for adsorbing hydrogen molecules showed that the hollow CNOs exhibit good performance as materials for electrochemical hydrogen storage. Furthermore, using DFT calculations, the authors in [115] studied the properties of stoichiometric and defective nitrogen carbon nano onions, or NCNOs. Exploring the stability and thermodynamic properties of mono and divacancy-containing CNOs showed that the latter is more stable and can increase the bandgap notably.
To finish this section, we will focus on buckybowls. A buckybowl is not a hybrid system, but an interesting one nevertheless. These structures come from the fragmentation of fullerenes. Buckybowls are adsorbents of CO2, CH4, and C2H2. Studies using hybrid DFT calculations found that the curvature of the buckybowl surface plays a significant role in the adsorption of these gas molecules. Buckybowls are high-performance materials for capturing greenhouse gases, molecular recognition, and nanotechnology applications in general [116]. Hybrid nanostructures can also see their adsorption capacity increase by doping them, usually with lithium and nitrogen as dopants. Table 5 shows the main quantitative results of this section.

3. Adsorption on Systems Doped with Transition Metals

3.1. Transition Metals

The definition of transition metals usually refers to those elements in groups 3–12 on the periodic table, with partially filled d-orbitals. The frequent name for the lanthanide and actinide series is inner transition metals. However, it is more frequent to call transition metals the elements of groups 4–11. This fact relies on the typical chemical behavior of the metals in these groups. Note that the filling of the d-orbitals increases going from left to right in the periodic table. It is usual to name them the “d-block elements.” Thus, these metals have several properties in common. First, they have low ionization energies, and several of them form paramagnetic compounds. Second, these metals are reactive but not as much as the alkalis. Finally, they conduct heat and electricity very well and have high melting and boiling temperatures.

3.2. Nanotubes Doped with Transition Metals

Doping nanotubes to increase their adsorption capacity is frequent. In particular, the decoration involving transition metals, oxides, or polymers, can increase their sensitivity [119]. It is also frequent to use transition metals Pd, Pt, Ag, Au, and Rh for doping carbon nanotubes that in turn can be used as gas sensors [31,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136] for detection of gases of C3H6O (acetone), C6H6 (benzene), H2, CH4, NH3, NO2, H2O, CO, H2S, HCN, HCl, Cl2. Specifically, Pd used for the adsorption of H2, Cl, CH4, Cl2, acetone, benzene; Pt used for the adsorption of C7H8 (toluene), H2, NO2, H2O, NH2; Au for the adsorption of C2H6O (ethanol), NH3, NO2, and Rh used for the adsorption of HCl, CO, and CH4. Furthermore, doping with Fe makes adsorption of the hydrogen molecule stronger than on an undoped nanotube. The adsorption of Fe on the nanotube is exothermic [137].
Experimental studies found that the H2 storage capacity of pristine MWCNTs increases up to 15 times at room temperature and 2.0 MPa by decorating them with ultrafine Ti nanoparticles. This improvement does not affect the cyclic hydrogenation–dehydrogenation operation of the tubes [138]. The intercalation of nickel on MWCNTs increases the H2 adsorption, probably due to the strong interaction between the Ni atoms and the MWNTs [139].
DFT calculations of Pt-decorated nanotubes show different effects in the adsorption capacities of the latter. For instance, the nanotube conductivity can vary with the adsorption of H2S, CO, and SO2 [140]. Another DFT calculation shows that Au-decorated nanotubes have a high sensitivity to adsorption of SO2 and H2S [141]. These Au-decorated nanotubes help sense propanone or acetone (C3H6O) gas, with the Au decoration increasing the device sensitivity up to around 3% at room temperature [142].
Rh-doped single-walled carbon nanotubes are advantageous for the adsorption of CO, C2H4 (ethylene), and SO2 molecules while proving to be insensitive to CO2 and CH4 [143]. DFT calculations using dispersion corrections in [144] showed that the doping of a (14,0) carbon nanotube by substituting a carbon atom with a Ru atom conveys better adsorption of SO2 and H2S molecules by the Ru-doped SWNT.
Pd and Pt decorations on a nanotube are practical to adsorb CO and NO molecules. Pd decoration is more suitable to adsorb NO, while Pt works better for CO adsorption. Ab initio calculations showed that this adsorption would change the magnetic properties of the Pd- and Pt-decorated SWCNTs, which would be helpful for sensors [145].
Pd as a dopant facilitates the adsorption of N2O on the otherwise inert nanotube [146]. More DFT calculations predict that Rh-doped nanotubes present a good adsorption ability of O2 and O3 and show relatively significant conductivity changes [147]. There is experimental evidence that Au nanoparticles adsorbed on nanotubes are helpful for acetone gas sensing at room temperature [148].
Table 6 presents a list of transition metals used for doping carbon nanotubes and the gases adsorbed in the doped nanotube. Notice that the most used is Au, and the least used are Ag and Fe. From the 32 elements in the periodic table (groups 4–11), we found mainly six transition metals used as dopants. Table 7 shows the main quantitative results of this section.

3.3. Graphene Doped with Transition Metals

We label TM-doped graphene the graphene doped with transition metals in this review. As with the other nanostructures considered in this review, hydrogen adsorption is a heavily explored topic, and transition metal-doped graphene has been no exception. For instance, a TiO2 decoration of nanoparticles on graphene oxide (GO + Ti) can adsorb twice as much H2 as pristine graphene at room temperature and 5 MPa, showing reversibility of up to 80% [149]. Copper-decorated N-doped defective graphene nanoribbons can improve their reactivity towards H2 adsorption. In [150], a DFT study considered single vacancy defects on graphene doped with N atoms, forming a pyridine-like structure. The authors labeled these structures as SV+1N, SV+2N, or SV+3N, depending on the number of nitrogen atoms. They added Cu decoration on each resulting system and concluded that the Cu-decorated SV+3N showed the best H2 adsorption performance, with a promising reversible cycle.
TM-doped graphene has been considered a possible candidate to sense different polluting gases, which usually interact weakly with pristine graphene. We have the chemical warfare agents (CWAs) among the pollutant gases under investigation, mentioned in Section 2.3. Transition metals such as Zr, Mo, Ti, Mn, Fe, and Co dope graphene and increase adsorption capabilities. For instance, Mn-doped graphene shows the most substantial chemical adsorption of phosgene (COCl2), a highly toxic CWA [59].
Other DFT studies suggest that Mn-doped graphene substantially improves the adsorption of C2H2, CH4, and CO, which are characteristic dissolved gases in transformer oil. In [54], the authors found that this substrate would be appropriate for sensing only the first two gases mentioned. The adsorption of CH4 is weak due to an electronic hybridization.
TM-doped graphene with TM = Ni, Cu, Zn, or Ni-doped vacancy defect graphene (DG) and graphene oxide (G-OH) adsorb hydrogen sulfide (H2S) that is another toxic and colorless gas. DFT studies show that Cu-doped graphene and DG display the best performance to adsorb H2S among the cases considered. Ni and Zn-doped graphene show weaker adsorption energies. Other studies show that Pt-decorated graphene is an H2S detector. The most stable configuration for adsorption corresponds to the H atoms of the H2S pointing towards the TM-doped graphene [46,48,58]. The bilayer graphene (BG) adsorbs H2S, and this adsorption increases by doping (BG) with transition metals as Fe, Ni, Mn, Cr, Co, and V. Notice that the TM-doping can occur at more sites concerning a single graphene layer, for example, in the interlayer region.
DFT calculations also show that an external electric field would further increase the adsorption of H2S onto such TM/BG systems. As most TM/BG-H2S systems have semiconductor behavior, this could also be a good candidate for sensing devices [53]. Additionally, the H2S sensing performance of zigzag graphene nanoribbons (ZGNR) is also increased mainly by combined doping of Cu and Zn atoms, the resulting system labeled as Cu/Zn-ZGNR, with a high response value of around 49% [51].
TM-doped graphene (with TM = Ti, Mn, Fe, Co, Ni, and Ag) adsorbs arsine (AsH3), another toxic gas. As in previous cases, TM doping increases the chemical interaction with the gas compared with pristine graphene. In [57], the authors found that Ni-doped graphene has the best AsH3/CO selectivity, an important feature when removing arsine from CO gas streams. The resulting hybridization of states between AsH3, CO, and TM orbitals confirmed the chemical interaction.
Recent theoretical investigations utilizing DFT considered Ag- Pt- and Au-doped graphene as a possible sensing material of nitrogen oxides such as NO and NO2. Au-doped graphene shows the best adsorption performance of NO/NO2 molecules of the three transition metals considered. Ag/Pt/Au-doping generally gives better sensing results than non-TM doping like B, N, Al, or S [56]. In [151], from DFT simulations, the authors found that Fe-doped graphene shows a high NO2 adsorption rate due to the d-orbital impurity states from the metal on the graphene surface. They also found that an increase in the strain applied to the graphene sheet decreases its adsorption capacity. Fe-doped armchair graphene nanoribbons (Fe-AGNR) are candidates for NO and NO2 sensing devices due to their electronic and transport properties [152].
Group 10 transition metals such as Ni, Pd and Pt can also decorate graphene to increase its interactions with NO2 [46] significantly. Of the same group 10, palladium is another exciting option for graphene doping. It has been shown via DFT calculations that Pd-doped graphene can adsorb CO and NO molecules, with the most stable site being a bridge site (B-site). Pd doping vastly increases the interaction between the substrate and CO/NO [55].
Platinum clusters like the Pt13 structure, when supported on pristine graphene, can also enhance the adsorption capacity of the latter. A systematic DFT study showed that such a composite could adsorb CO2, NO2, and SO2 more effectively than defective graphene (DG) supported Pt13 [153].
The latter gas, sulfur dioxide, has also been the subject of experimental and theoretical studies involving TM-decorated graphene. SO2 can be adsorbed more efficiently by metal-oxide such as ZnO, BeO, and Ni-decorated graphene than pristine graphene. ZnO- and BeO-decorated graphene physisorb SO2, while Ni-decorated graphene chemisorbs it [52].
TM-doped graphene can also enhance the detection of formaldehyde (HCHO) gas. A first-principles study considered a two-probe sensor device built with TM-doped graphene, where TM = Co, Ni, Cu, Zn, Pd, and Ag. The Cu- and Ag-doped graphene devices showed the best performance at low voltages, with a short response time and high HCHO sensing [154]. Table 8 shows the main quantitative results of this section.

3.4. Fullerenes Doped with Transition Metals

The TM-doping or decoration increases the adsorption capacities of fullerenes. This increase by doping is present in the other carbon nanostructures reviewed in this work. In [79], the authors performed a systematic DFT study focused on TM decoration of C60 fullerene, with TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn. The changes in the electronic structure implied that the TM-decorated C60 fullerene is more sensitive to CO and NO molecules than the pristine one.
Doping fullerenes with titanium increases their H2 storage capacity. A way of doing it is by substituting carbon atoms with titanium atoms and doping a C60 fullerene with an increasing number of Ti atoms, from one up to six. With six Ti atoms doping C60 fullerenes in this fashion—labeled as Ti6C48—the hydrogen storage can be up to 7.7 wt% [156]. In [157], the authors performing DFT simulations found that Co atoms forming compact clusters on the surface of a C60 fullerene labeled as C60Con, chemisorb an H2 molecule, with different features for different values of n (n = 1 to 8). They also found that up to 13 H2 molecules could be absorbed in this way.
Theoretical calculations performed in [158] showed that a C70 fullerene doped with TM ions (Cr2+ and Co2+) adsorbed nitrogen dioxide (NO2) on the exterior surface. The doping mechanism is a porphyrin-induced process labeled as PIC70F. In this manner, this fullerene would be helpful for selective detection of NO2 in the presence of sulfur dioxide, also showing a short recovery time.
Another way of doping fullerenes is from the inside, an example of which would be the endohedral metallofullerenes (EMFs). Such structures are obtained by putting a lowest energy structure of a metal cluster inside of the fullerene and then relaxing the system. When C60 fullerenes are considered, the EMFs obtained in this manner are labeled as Mn@C60, and they can increase the oxygen reduction reaction (ORR) activities of the former. A comprehensive DFT study showed that EMFs such as Mn5@C60, Cu4@C60, Co2@C60, and Ni4@C60 have better ORR activities, which could be helpful in developing sensing devices able to detect polluting gases [159]. Table 9 shows the main quantitative results of this section.

3.5. Graphdiyne Doped with Transition Metals

Transition metals (TMs) can enhance the adsorption properties of graphdiyne (GDY). Several DFT studies have explored the adsorption of metals like Au, Cu, Ni, and Zn—although Zn is not a transition metal—on graphdiyne. GDY strongly chemisorbs nickel, and it also chemisorbs silver and copper. On the other hand, graphdiyne physisorbs a zinc atom. Overall, this material is better adsorbent for those metals compared with graphene [107]. When the GDY surface first adsorbs one or two nitrogen atoms, the resulting N-GDY composite increases the stability of further transition metal decorations. In [160], the authors investigated the adsorption of several TMs (Cr, Mn, Fe, Co, Ni, and Cu) on such N-GDY substrates. They found that Fe@2N-GDY (a Fe atom decorating a GDY doped with 2 N atoms) has the best catalytic activity with the lowest CO oxidization energy barrier.
As with the other nanostructures covered in this review, graphdiyne has been explored for its H2 adsorption capabilities, aiming at reversible hydrogen storage. DFT studies have considered decorating graphdiyne nanosheets with light metals such as Li, Na, K, Ca, Sc and Ti. Some works have explored up to 11% of metal functionalization in the GDY surface, which can then anchor multiple H2 molecules [161].
Water splitting is another exciting application of TM-decorated GDY nanosheets. For instance, theoretical simulations and experimental studies show that Ru-doped GDY (Ru/GDY) has a high catalytic performance for oxygen evolution reactions [162].
The electrochemical reduction of N2 at room temperature is relevant for the production of ammonia (NH3). A systematic study considering transition-metal-embedded GDY, or TM@GDY, explored the effect of Sc, Fe,Cr, Mn, Mo,Co, Ni, Cu, Zn, Ru, Rh, Pd, and Ag in the N2 reduction reaction. According to spin-polarized DFT simulations, the most stable and best catalytic activity was the Mo-embedded graphdiyne monolayer [163]. All TMs increase GDY sheets’ N2 reduction reaction (NRR) character. However, other studies indicate that the single-atom catalyst V@GDY monolayer has the best NRR performance [82].
Anchoring two transition metals to GDY improves the NRR activity. In a study involving DFT calculations, the authors considered several TMs (Fe, Co, Ni, Cu, and Mo). They found that the best NRR performance was with the Co-Ni heteronuclear complex (CoNi@GDY) and the Mo-Mo homonuclear complex (Mo2@GDY) [164].
Scandium and titanium can also decorate graphdiyne to enhance sensing of formaldehyde (HCHO), a typical air pollutant. Sc or Ti take a stable place on the corner sites of graphdiyne sheets, and both can better adsorb HCHO compared with pristine graphdiyne and graphene. This effect would be due to the electronegativity of HCHO, and the best result is with Ti-decorated graphdiyne, according to DFT calculations [165].
In [108], the authors investigated the palladium clusters supported on graphdiyne surfaces utilizing DFT calculations. They found that the electronic HOMO-LUMO gap changes with the size of said clusters, which chemisorbed with relatively high adsorption energies of around 3–4 eV in magnitude. Palladium nanoparticles can also be experimentally anchored to the GDY surface, forming a stable nanocatalyst labeled PdNPs/GDY. Such composite can decompose H2O2 to produce O2, an essential reaction in antitumor treatments combined with doxorubicin, a chemotherapeutic agent [166].
Doping GDY with Fe atoms improves the electrochemical reduction of CO2 by GDY sheets. An ab initio study found that doping the GDY surface with a Fe dimer or a Fe trimer optimizes the CO2 adsorption and selectivity. The number of Fe atoms considered can generally tune the GDY catalytic activity significantly [167]. CO oxidation would also be accomplished employing several single-atom catalysts (SACs) such as Ni-GDY and Cu-GDY. DFT calculations have shown that both Ni and Cu can be anchored at the corner of the acetylenic ring of graphdiyne, the resulting system being able to adsorb CO [168]. Table 10 shows the main quantitative results of this section.

3.6. Hybrid Systems Doped with Transition Metals

Several studies have also focused on the TM-doping of hybrid carbon nanostructures starting. The most convenient places to dope a buckybowl have been studied by computational means, covering the binding possibilities of the transition-metal ion Cu+, and cations as Li+, Na+, K+, to buckybowls on convenient locations where the gradient electron density favors the adsorption [169]. In other studies, authors have investigated the catalytic effect of Ni, Fe, and a Fe-Ni alloy on synthesizing metal-containing carbon nano-onions (CNOs) and studied their electrochemical hydrogen storage properties. They found that the electrochemical hydrogen storage capacity of the CNOs is in the order of Ni@CNOs > Fe0.64Ni0.36@CNOs > Fe3C@CNOs. The Ni@CNOs have a maximum hydrogen storage capacity of 1.42%. Large amounts of defects, good electrical conductivity, and electrocatalytic activity of the Ni particles are responsible for their excellent electrochemical performance [170].
In [171], the authors considered the doping of a semi-fullerene C30 with titanium to adsorb molecules of CO and CO2. With an exploration involving density functional theory (DFT) and first-principles molecular dynamics (FPMD) at 300 K and atmospheric pressure, they found that the most stable adsorption of the titanium atom on C30 occurs on the concave surface of the molecule. Besides, the considered molecules are chemisorbed, with no dissociation. The adsorption energies depend on the initial orientation of the molecules concerning TiC30. Similarly, in [172], density functional theory (DFT) was also used to study the adsorption of an H2 molecule in a system formed by a graphene layer and a Ti-doped semi-fullerene. The authors found that the semi-fullerene is bound to the graphene layer, with one of the hexagonal faces of the former being oriented into the latter. Besides, the semi-fullerene chemisorbs the titanium atom. Finally, the authors studied the interaction between the hydrogen molecule and the combined system, finding that the system can adsorb the H2 molecule. Table 11 shows the main quantitative results of this section.

4. Conclusions

We aimed to present the most important and recent advances in the study of polluting gas adsorption employing carbon nanostructures. The porous nature of the latter has made them a natural choice for developing sensing devices throughout the years, and physical features like diameter, curvature or size, influence their adsorption capabilities. The carbon nanostructures reviewed in general physically adsorb many of the pollutant gas molecules considered, and the interaction can usually increase by different means of nonmetal functionalizations or doping. However, transition-metal doping and decoration give better results overall, increasing their sensing properties, involving chemisorption in most cases studied. There have also been many systems showing a good adsorption–desorption cyclic performance, which makes them good candidates for sensing devices.
Among the transition metals used as dopants in the reviewed works, we found Ag, Au, Pt, Pd, Fe, Rh, Zr, Mo, Ti, Mn, Co, Ni, Cu, Sc, and V. Besides, we found the non-transition metals as dopants too: Zn, Si, P, S, As, Se, Te, Li, N, and B.
On the other hand, among the pollutant molecules investigated for adsorption in carbon nanostructures that we found in this review, we have NO2, NH3, NH2, SO2, H2S, C2H6O, C3H6O, C7H8, NO, Cl2, CH4, N2O, CO, C2H4, CO2, O3, CH3OH, H2CO. H2CO2, C2H2, SF6, C2N2, C6H14, C7H8, C2H3N, CH2Cl2 CH3COOC2H5, (CH3)2NH, C3H9N, AsH3. Moreover, the anticancer drugs 5-fluorouracil and temozolomide. We also found amphetamines. Finally, we also found the chemical warfare agents (CWAs) tabun, sarin, soman, cyclosarin, phosgene, and Lewisite molecules.
Carbon nanotubes and graphene are probably the most investigated systems, as shown by the abundance of cites. However, the detailed study of its adsorption properties is ongoing.
To end, we should mention that the study of the adsorption properties of carbon nanostructures is far from being completed. Relatively recent structures like graphdiyne have plenty of potential applications waiting for the investigation, and the building of hybrid systems has much to offer for materials scientists in the future.

Author Contributions

Conceptualization, L.F.M.; writing—original draft preparation, L.F.M., J.M.R.-d.-A. and M.C.; writing—review and editing, L.F.M. and J.M.R.-d.-A.; graphic materials, M.C.; funding acquisitions, L.F.M. All authors have read and agreed to the published version of the manuscript.

Funding

We thank Dirección General de Asuntos del Personal Académico de la Universidad Nacional Autónoma de México, partial financial support by Grant IN113220.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Dirección General de Asuntos del Personal Académico de la Universidad Nacional Autónoma de México, partial financial support by Grant IN113220. We also appreciate UNAM-Miztli-Super-Computing Center technical assistance by the project LANCAD-UNAM-DGTIC-030.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kroto, H.W.; Heath, J.R.; O’Brien, S.C.; Curl, R.F.; Smalley, R.E. C60: Buckminsterfullerene. Nature 1985, 318, 162–163. [Google Scholar] [CrossRef]
  2. Nasibulin, A.G.; Pikhitsa, P.V.; Jiang, H.; Brown, D.P.; Krasheninnikov, A.V.; Anisimov, A.S.; Queipo, P.; Moisala, A.; Gonzalez, D.; Lientschnig, G.; et al. A Novel Hybrid Carbon Material. Nat. Nanotechnol. 2007, 2, 156–161. [Google Scholar] [CrossRef] [PubMed]
  3. Moisala, A.; Nasibulin, A.G.; Shandakov, S.D.; Jiang, H.; Kauppinen, E.I. On-Line Detection of Single-Walled Carbon Nanotube Formation during Aerosol Synthesis Methods. Carbon 2005, 43, 2066–2074. [Google Scholar] [CrossRef]
  4. Delgado, J.L.; Herranz, M.; Martín, N. The Nano-Forms of Carbon. J. Mater. Chem. 2008, 18, 1417. [Google Scholar] [CrossRef]
  5. Falcao, E.H.; Wudl, F. Carbon Allotropes: Beyond Graphite and Diamond. J. Chem. Technol. Biotechnol. 2007, 82, 524–531. [Google Scholar] [CrossRef]
  6. Langenhorst, F.; Campione, M. Ideal and Real Structures of Different Forms of Carbon, with Some Remarks on Their Geological Significance. J. Geol. Soc. 2019, 176, 337–347. [Google Scholar] [CrossRef]
  7. Pacheco, M.; Pacheco, J.; Valdivia, R.; Santana, A.; Tu, X.; Mendoza, D.; Frias, H.; Medina, L.; Macias, J. Green Applications of Carbon Nanostructures Produced by Plasma Techniques. MRS Adv. 2017, 2, 2647–2659. [Google Scholar] [CrossRef] [Green Version]
  8. Lejaeghere, K.; Van Speybroeck, V.; Van Oost, G.; Cottenier, S. Error Estimates for Solid-State Density-Functional Theory Predictions: An Overview by Means of the Ground-State Elemental Crystals. Crit. Rev. Solid State Mater. Sci. 2014, 39, 1–24. [Google Scholar] [CrossRef] [Green Version]
  9. Boyd, A.; Dube, I.; Fedorov, G.; Paranjape, M.; Barbara, P. Gas Sensing Mechanism of Carbon Nanotubes: From Single Tubes to High-Density Networks. Carbon 2014, 69, 417–423. [Google Scholar] [CrossRef]
  10. Zhao, J.; Buldum, A.; Han, J.; Lu, J.P. Gas Molecule Adsorption in Carbon Nanotubes and Nanotube Bundles. Nanotechnology 2002, 13, 195–200. [Google Scholar] [CrossRef]
  11. Feng, Y.; Wang, J.; Liu, Y.; Zheng, Q. Adsorption Equilibrium of Hydrogen Adsorption on Activated Carbon, Multi-Walled Carbon Nanotubes and Graphene Sheets. Cryogenics 2019, 101, 36–42. [Google Scholar] [CrossRef]
  12. Panella, B.; Hirscher, M.; Roth, S. Hydrogen Adsorption in Different Carbon Nanostructures. Carbon 2005, 43, 2209–2214. [Google Scholar] [CrossRef]
  13. Dhall, S.; Jaggi, N.; Nathawat, R. Functionalized Multiwalled Carbon Nanotubes Based Hydrogen Gas Sensor. Sens. Actuators A Phys. 2013, 201, 321–327. [Google Scholar] [CrossRef]
  14. Elyassi, M.; Rashidi, A.; Hantehzadeh, M.R.; Elahi, S.M. Hydrogen Storage Behaviors by Adsorption on Multi-Walled Carbon Nanotubes. J. Inorg. Organomet. Polym. 2017, 27, 285–295. [Google Scholar] [CrossRef]
  15. Lithoxoos, G.P.; Labropoulos, A.; Peristeras, L.D.; Kanellopoulos, N.; Samios, J.; Economou, I.G. Adsorption of N2, CH4, CO and CO2 Gases in Single Walled Carbon Nanotubes: A Combined Experimental and Monte Carlo Molecular Simulation Study. J. Supercrit. Fluids 2010, 55, 510–523. [Google Scholar] [CrossRef]
  16. Su, F.; Lu, C.; Cnen, W.; Bai, H.; Hwang, J.F. Capture of CO2 from Flue Gas via Multiwalled Carbon Nanotubes. Sci. Total Environ. 2009, 407, 3017–3023. [Google Scholar] [CrossRef] [PubMed]
  17. Hsu, S.-C.; Lu, C.; Su, F.; Zeng, W.; Chen, W. Thermodynamics and Regeneration Studies of CO2 Adsorption on Multiwalled Carbon Nanotubes. Chem. Eng. Sci. 2010, 65, 1354–1361. [Google Scholar] [CrossRef]
  18. Su, F.; Lu, C.; Chung, A.-J.; Liao, C.-H. CO2 Capture with Amine-Loaded Carbon Nanotubes via a Dual-Column Temperature/Vacuum Swing Adsorption. Appl. Energy 2014, 113, 706–712. [Google Scholar] [CrossRef]
  19. Inoue, S.; Tomita, Y.; Kokabu, T.; Matsumura, Y. Principles of Detection Mechanism for Adsorbed Gases Using Carbon Nanotube Nanomat. Chem. Phys. Lett. 2018, 709, 77–81. [Google Scholar] [CrossRef]
  20. Fatemi, S.; Vesali-Naseh, M.; Cyrus, M.; Hashemi, J. Improving CO2/CH4 Adsorptive Selectivity of Carbon Nanotubes by Functionalization with Nitrogen-Containing Groups. Chem. Eng. Res. Des. 2011, 89, 1669–1675. [Google Scholar] [CrossRef]
  21. Lin, C.-C.; Gupta, S.; Chang, C.; Lee, C.-Y.; Tai, N.-H. Polyethylenimine-Polyethylene Glycol/Multi-Walled Carbon Nanotubes Bilayer Structure for Carbon Dioxide Gas Sensing at Room Temperature. Mater. Lett. 2021, 297, 129941. [Google Scholar] [CrossRef]
  22. Rahimi, K.; Riahi, S.; Abbasi, M.; Fakhroueian, Z. Modification of Multi-Walled Carbon Nanotubes by 1,3-Diaminopropane to Increase CO2 Adsorption Capacity. J. Environ. Manag. 2019, 242, 81–89. [Google Scholar] [CrossRef]
  23. Faginas-Lago, N.; Apriliyanto, Y.B.; Lombardi, A. Confinement of CO2 inside Carbon Nanotubes. Eur. Phys. J. D 2021, 75, 161. [Google Scholar] [CrossRef]
  24. Mukhtar, A.; Mellon, N.; Saqib, S.; Khawar, A.; Rafiq, S.; Ullah, S.; Al-Sehemi, A.G.; Babar, M.; Bustam, M.A.; Khan, W.A.; et al. CO2/CH4 Adsorption over Functionalized Multi-Walled Carbon Nanotubes; an Experimental Study, Isotherms Analysis, Mechanism, and Thermodynamics. Microporous Mesoporous Mater. 2020, 294, 109883. [Google Scholar] [CrossRef]
  25. Delavar, M.; Asghar Ghoreyshi, A.; Jahanshahi, M.; Khalili, S.; Nabian, N. Equilibria and Kinetics of Natural Gas Adsorption on Multi-Walled Carbon Nanotube Material. RSC Adv. 2012, 2, 4490. [Google Scholar] [CrossRef]
  26. Babu, D.J.; Lange, M.; Cherkashinin, G.; Issanin, A.; Staudt, R.; Schneider, J.J. Gas Adsorption Studies of CO2 and N2 in Spatially Aligned Double-Walled Carbon Nanotube Arrays. Carbon 2013, 61, 616–623. [Google Scholar] [CrossRef]
  27. Liu, L.; Nicholson, D.; Bhatia, S.K. Impact of H2O on CO2 Separation from Natural Gas: Comparison of Carbon Nanotubes and Disordered Carbon. J. Phys. Chem. C 2015, 119, 407–419. [Google Scholar] [CrossRef]
  28. Chiang, Y.-C.; Wu, P.-Y. Adsorption Equilibrium of Sulfur Hexafluoride on Multi-Walled Carbon Nanotubes. J. Hazard. Mater. 2010, 178, 729–738. [Google Scholar] [CrossRef]
  29. Lide, D.R. (Ed.) CRC Handbook of Chemistry and Physics: A Ready-Reference Book of Chemical and Physical Data, 81st ed.; CRC Press: Boca Raton, FL, USA, 2000; ISBN 978-0-8493-0481-1. [Google Scholar]
  30. Kazachkin, D.V.; Nishimura, Y.; Irle, S.; Feng, X.; Vidic, R.; Borguet, E. Temperature and Pressure Dependence of Molecular Adsorption on Single Wall Carbon Nanotubes and the Existence of an “Adsorption/Desorption Pressure Gap”. Carbon 2010, 48, 1867–1875. [Google Scholar] [CrossRef]
  31. Young, S.-J.; Lin, Z.-D. Ethanol Gas Sensors Composed of Carbon Nanotubes with Au Nanoparticles Adsorbed onto a Flexible PI Substrate. ECS J. Solid State Sci. Technol. 2017, 6, M130–M132. [Google Scholar] [CrossRef]
  32. Abdulla, S.; Mathew, T.L.; Pullithadathil, B. Highly Sensitive, Room Temperature Gas Sensor Based on Polyaniline-Multiwalled Carbon Nanotubes (PANI/MWCNTs) Nanocomposite for Trace-Level Ammonia Detection. Sens. Actuators B Chem. 2015, 221, 1523–1534. [Google Scholar] [CrossRef]
  33. Battie, Y.; Ducloux, O.; Thobois, P.; Dorval, N.; Lauret, J.S.; Attal-Trétout, B.; Loiseau, A. Gas Sensors Based on Thick Films of Semi-Conducting Single Walled Carbon Nanotubes. Carbon 2011, 49, 3544–3552. [Google Scholar] [CrossRef]
  34. Chen, G.; Paronyan, T.M.; Pigos, E.M.; Harutyunyan, A.R. Enhanced Gas Sensing in Pristine Carbon Nanotubes under Continuous Ultraviolet Light Illumination. Sci. Rep. 2012, 2, 343. [Google Scholar] [CrossRef]
  35. Huyen, D.N.; Tung, N.T.; Vinh, T.D.; Thien, N.D. Synergistic Effects in the Gas Sensitivity of Polypyrrole/Single Wall Carbon Nanotube Composites. Sensors 2012, 12, 7965–7974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Kuznetsova, A.; Yates, J.T.; Simonyan, V.V.; Johnson, J.K.; Huffman, C.B.; Smalley, R.E. Optimization of Xe Adsorption Kinetics in Single Walled Carbon Nanotubes. J. Chem. Phys. 2001, 115, 6691–6698. [Google Scholar] [CrossRef] [Green Version]
  37. Beheshtian, J.; Peyghan, A.A.; Bagheri, Z. Nitrate Adsorption by Carbon Nanotubes in the Vacuum and Aqueous Phase. Mon. Chem. 2012, 143, 1623–1626. [Google Scholar] [CrossRef]
  38. Beheshtian, J.; Peyghan, A.A.; Bagheri, Z. Formaldehyde Adsorption on the Interior and Exterior Surfaces of CN Nanotubes. Struct. Chem. 2013, 24, 1331–1337. [Google Scholar] [CrossRef]
  39. Hu, Z.; Xie, H.; Wang, Q.; Chen, S. Adsorption and Diffusion of Sulfur Dioxide and Nitrogen in Single-Wall Carbon Nanotubes. J. Mol. Graph. Model. 2019, 88, 62–70. [Google Scholar] [CrossRef] [PubMed]
  40. Li, M.; Wu, S.C.; Shih, Y. Characterization of Volatile Organic Compound Adsorption on Multiwall Carbon Nanotubes under Different Levels of Relative Humidity Using Linear Solvation Energy Relationship. J. Hazard. Mater. 2016, 315, 35–41. [Google Scholar] [CrossRef]
  41. Wang, Y.; Yeow, J.T.W. A Review of Carbon Nanotubes-Based Gas Sensors. J. Sens. 2009, 2009, 493904. [Google Scholar] [CrossRef]
  42. Casolo, S.; Løvvik, O.M.; Martinazzo, R.; Tantardini, G.F. Understanding Adsorption of Hydrogen Atoms on Graphene. J. Chem. Phys. 2009, 130, 054704. [Google Scholar] [CrossRef]
  43. Feijó, T.O.; Rolim, G.K.; Corrêa, S.A.; Radtke, C.; Soares, G.V. Thermally Driven Hydrogen Interaction with Single-Layer Graphene on SiO2/Si Substrates Quantified by Isotopic Labeling. J. Appl. Phys. 2020, 128, 225702. [Google Scholar] [CrossRef]
  44. Sunnardianto, G.K.; Bokas, G.; Hussein, A.; Walters, C.; Moultos, O.A.; Dey, P. Efficient Hydrogen Storage in Defective Graphene and Its Mechanical Stability: A Combined Density Functional Theory and Molecular Dynamics Simulation Study. Int. J. Hydrog. Energy 2021, 46, 5485–5494. [Google Scholar] [CrossRef]
  45. Leenaerts, O.; Partoens, B.; Peeters, F.M. Adsorption of H2O, NH3, CO, NO2, and NO on Graphene: A First-Principles Study. Phys. Rev. B 2008, 77, 125416. [Google Scholar] [CrossRef] [Green Version]
  46. Bo, Z.; Guo, X.; Wei, X.; Yang, H.; Yan, J.; Cen, K. Density Functional Theory Calculations of NO2 and H2S Adsorption on the Group 10 Transition Metal (Ni, Pd and Pt) Decorated Graphene. Phys. E Low-Dimens. Syst. Nanostruct. 2019, 109, 156–163. [Google Scholar] [CrossRef]
  47. Gao, H.; Liu, Z. DFT Study of NO Adsorption on Pristine Graphene. RSC Adv. 2017, 7, 13082–13091. [Google Scholar] [CrossRef] [Green Version]
  48. Gao, X.; Zhou, Q.; Wang, J.; Xu, L.; Zeng, W. Performance of Intrinsic and Modified Graphene for the Adsorption of H2S and CH4: A DFT Study. Nanomaterials 2020, 10, 299. [Google Scholar] [CrossRef] [Green Version]
  49. Huang, S.; Panes-Ruiz, L.A.; Croy, A.; Löffler, M.; Khavrus, V.; Bezugly, V.; Cuniberti, G. Highly Sensitive Room Temperature Ammonia Gas Sensor Using Pristine Graphene: The Role of Biocompatible Stabilizer. Carbon 2021, 173, 262–270. [Google Scholar] [CrossRef]
  50. Lazar, P.; Karlický, F.; Jurečka, P.; Kocman, M.; Otyepková, E.; Šafářová, K.; Otyepka, M. Adsorption of Small Organic Molecules on Graphene. J. Am. Chem. Soc. 2013, 135, 6372–6377. [Google Scholar] [CrossRef]
  51. Salih, E.; Ayesh, A.I. Co-Doped Zigzag Graphene Nanoribbon Based Gas Sensor for Sensitive Detection of H2S: DFT Study. Superlattices Microstruct. 2021, 155, 106900. [Google Scholar] [CrossRef]
  52. Karami, Z.; Hamed Mashhadzadeh, A.; Habibzadeh, S.; Ganjali, M.R.; Ghardi, E.M.; Hasnaoui, A.; Vatanpour, V.; Sharma, G.; Esmaeili, A.; Stadler, F.J.; et al. Atomic Simulation of Adsorption of SO2 Pollutant by Metal (Zn, Be)-Oxide and Ni-Decorated Graphene: A First-Principles Study. J. Mol. Model. 2021, 27, 70. [Google Scholar] [CrossRef] [PubMed]
  53. Xie, Y.; Cao, S.; Wu, X.; Yu, B.-Y.; Chen, L.-Y.; Zhang, J.-M. Density Functional Theory Study of Hydrogen Sulfide Adsorption onto Transition Metal-Doped Bilayer Graphene Using External Electric Fields. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 124, 114252. [Google Scholar] [CrossRef]
  54. Gui, Y.; Peng, X.; Liu, K.; Ding, Z. Adsorption of C2H2, CH4 and CO on Mn-Doped Graphene: Atomic, Electronic, and Gas-Sensing Properties. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 119, 113959. [Google Scholar] [CrossRef]
  55. Shukri, M.S.M.; Saimin, M.N.S.; Yaakob, M.K.; Yahya, M.Z.A.; Taib, M.F.M. Structural and Electronic Properties of CO and NO Gas Molecules on Pd-Doped Vacancy Graphene: A First Principles Study. Appl. Surf. Sci. 2019, 494, 817–828. [Google Scholar] [CrossRef]
  56. Jia, X.; An, L. The Adsorption of Nitrogen Oxides on Noble Metal-Doped Graphene: The First-Principles Study. Mod. Phys. Lett. B 2019, 33, 1950044. [Google Scholar] [CrossRef]
  57. Li, Y.; Sun, X.; Zhou, L.; Ning, P.; Tang, L. Density Functional Theory Analysis of Selective Adsorption of AsH3 on Transition Metal-Doped Graphene. J. Mol. Model. 2019, 25, 145. [Google Scholar] [CrossRef]
  58. Khodadadi, Z. Evaluation of H2S Sensing Characteristics of Metals–Doped Graphene and Metals-Decorated Graphene: Insights from DFT Study. Phys. E Low-Dimens. Syst. Nanostruct. 2018, 99, 261–268. [Google Scholar] [CrossRef]
  59. Zhang, T.; Sun, H.; Wang, F.; Zhang, W.; Tang, S.; Ma, J.; Gong, H.; Zhang, J. Adsorption of Phosgene Molecule on the Transition Metal-Doped Graphene: First Principles Calculations. Appl. Surf. Sci. 2017, 425, 340–350. [Google Scholar] [CrossRef]
  60. Promthong, N.; Tabtimsai, C.; Rakrai, W.; Wanno, B. Transition Metal-Doped Graphene Nanoflakes for CO and CO2 Storage and Sensing Applications: A DFT Study. Struct. Chem. 2020, 31, 2237–2247. [Google Scholar] [CrossRef]
  61. Chuvilin, A.; Kaiser, U.; Bichoutskaia, E.; Besley, N.A.; Khlobystov, A.N. Direct Transformation of Graphene to Fullerene. Nat. Chem. 2010, 2, 450–453. [Google Scholar] [CrossRef]
  62. Manna, A.K.; Pati, S.K. Stability and Electronic Structure of Carbon Capsules with Superior Gas Storage Properties: A Theoretical Study. Chem. Phys. 2013, 426, 23–30. [Google Scholar] [CrossRef]
  63. Khan, A.A.; Ahmad, I.; Ahmad, R. Influence of Electric Field on CO2 Removal by P-Doped C60-Fullerene: A DFT Study. Chem. Phys. Lett. 2020, 742, 137155. [Google Scholar] [CrossRef]
  64. Tascón, J.M.D.; Bottani, E.J. Ethylene Physisorption on C60 Fullerene. Carbon 2004, 42, 1333–1337. [Google Scholar] [CrossRef]
  65. Furuuchi, N.; Shrestha, R.; Yamashita, Y.; Hirao, T.; Ariga, K.; Shrestha, L. Self-Assembled Fullerene Crystals as Excellent Aromatic Vapor Sensors. Sensors 2019, 19, 267. [Google Scholar] [CrossRef] [Green Version]
  66. Khan, A.A.; Ahmad, R.; Ahmad, I. Removal of Nitrous and Carbon Mono Oxide from Flue Gases by Si-Coordinated Nitrogen Doped C60-Fullerene: A DFT Approach. Mol. Catal. 2021, 509, 111674. [Google Scholar] [CrossRef]
  67. Haghgoo, S.; Nekoei, A.-R. Metal Oxide Adsorption on Fullerene C60 and Its Potential for Adsorption of Pollutant Gases; Density Functional Theory Studies. RSC Adv. 2021, 11, 17377–17390. [Google Scholar] [CrossRef]
  68. Bubenchikov, M.A.; Bubenchikov, A.M.; Usenko, O.V.; Tsyrenova, V.B.; Budaev, S.O. Ability of Fullerene to Accumulate Hydrogen. EPJ Web Conf. 2016, 110, 01077. [Google Scholar] [CrossRef] [Green Version]
  69. Saha, D.; Deng, S. Hydrogen Adsorption on Partially Truncated and Open Cage C60 Fullerene. Carbon 2010, 48, 3471–3476. [Google Scholar] [CrossRef]
  70. Kaiser, A.; Leidlmair, C.; Bartl, P.; Zöttl, S.; Denifl, S.; Mauracher, A.; Probst, M.; Scheier, P.; Echt, O. Adsorption of Hydrogen on Neutral and Charged Fullerene: Experiment and Theory. J. Chem. Phys. 2013, 138, 074311. [Google Scholar] [CrossRef] [Green Version]
  71. Kalateh, K.; Cordshooli, G.A.; Kheirollahpoor, S. Hydrogen Adsorption,Structural, Electronic, and Spectroscopic Properties of C32, B16N16, and B8C24 by DFT Calculations. Fuller. Nanotub. Carbon Nanostruct. 2017, 25, 459–465. [Google Scholar] [CrossRef]
  72. Barajas-Barraza, R.E.; Guirado-López, R.A. Endohedral Nitrogen Storage in Carbon Fullerene Structures: Physisorption to Chemisorption Transition with Increasing Gas Pressure. J. Chem. Phys. 2009, 130, 234706. [Google Scholar] [CrossRef]
  73. Esrafili, M.D.; Janebi, H. B-, N-Doped and BN Codoped C60 Heterofullerenes for Environmental Monitoring of NO and NO2: A DFT Study. Mol. Phys. 2020, 118, e1631495. [Google Scholar] [CrossRef]
  74. Acosta-Gutiérrez, S.; Bretón, J.; Llorente, J.M.G.; Hernández-Rojas, J. Optimal Covering of C60 Fullerene by Rare Gases. J. Chem. Phys. 2012, 137, 074306. [Google Scholar] [CrossRef]
  75. Siadati, S.A.; Vessally, E.; Hosseinian, A.; Edjlali, L. Possibility of Sensing, Adsorbing, and Destructing the Tabun-2D-Skeletal (Tabun Nerve Agent) by C20 Fullerene and Its Boron and Nitrogen Doped Derivatives. Synth. Met. 2016, 220, 606–611. [Google Scholar] [CrossRef]
  76. Najafi, M. Density Functional Study of Cyanogen (C2N2) Sensing Using OH Functionalized Fullerene (C60) and Germanium-Fullerene (Ge60). Vacuum 2016, 134, 88–91. [Google Scholar] [CrossRef]
  77. Bashiri, S.; Vessally, E.; Bekhradnia, A.; Hosseinian, A.; Edjlali, L. Utility of Extrinsic [60] Fullerenes as Work Function Type Sensors for Amphetamine Drug Detection: DFT Studies. Vacuum 2017, 136, 156–162. [Google Scholar] [CrossRef]
  78. Elessawy, N.A.; El-Sayed, E.M.; Ali, S.; Elkady, M.F.; Elnouby, M.; Hamad, H.A. One-Pot Green Synthesis of Magnetic Fullerene Nanocomposite for Adsorption Characteristics. J. Water Process Eng. 2020, 34, 101047. [Google Scholar] [CrossRef]
  79. El Mahdy, A.M. Density Functional Investigation of CO and NO Adsorption on TM-Decorated C60 Fullerene. Appl. Surf. Sci. 2016, 383, 353–366. [Google Scholar] [CrossRef]
  80. Khan, A.A.; Ahmad, R.; Ahmad, I.; Su, X. Selective Adsorption of CO2 from Gas Mixture by P-Decorated C24N24 Fullerene Assisted by an Electric Field: A DFT Approach. J. Mol. Graph. Model. 2021, 103, 107806. [Google Scholar] [CrossRef]
  81. Wu, P.; Du, P.; Zhang, H.; Cai, C. Graphdiyne as a Metal-Free Catalyst for Low-Temperature CO Oxidation. Phys. Chem. Chem. Phys. 2014, 16, 5640–5648. [Google Scholar] [CrossRef] [PubMed]
  82. Feng, Z.; Tang, Y.; Chen, W.; Li, Y.; Li, R.; Ma, Y.; Dai, X. Graphdiyne Coordinated Transition Metals as Single-Atom Catalysts for Nitrogen Fixation. Phys. Chem. Chem. Phys. 2020, 22, 9216–9224. [Google Scholar] [CrossRef]
  83. Ebadi, M.; Reisi-Vanani, A. Methanol and Carbon Monoxide Sensing and Capturing by Pristine and Ca-Decorated Graphdiyne: A DFT-D2 Study. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 125, 114425. [Google Scholar] [CrossRef]
  84. Yang, Z.; Zhang, Y.; Guo, M.; Yun, J. Adsorption of Hydrogen and Oxygen on Graphdiyne and Its BN Analog Sheets: A Density Functional Theory Study. Comput. Mater. Sci. 2019, 160, 197–206. [Google Scholar] [CrossRef]
  85. Chen, X.; Gao, P.; Guo, L.; Zhang, S. Graphdiyne as a Promising Material for Detecting Amino Acids. Sci. Rep. 2015, 5, 16720. [Google Scholar] [CrossRef] [Green Version]
  86. Nagarajan, V.; Srimathi, U.; Chandiramouli, R. First-Principles Insights on Detection of Dimethyl Amine and Trimethyl Amine Vapors Using Graphdiyne Nanosheets. Comput. Theor. Chem. 2018, 1123, 119–127. [Google Scholar] [CrossRef]
  87. Srimathi, U.; Nagarajan, V.; Chandiramouli, R. Investigation on Graphdiyne Nanosheet in Adsorption of Sorafenib and Regorafenib Drugs: A DFT Approach. J. Mol. Liq. 2019, 277, 776–785. [Google Scholar] [CrossRef]
  88. Nagarajan, V.; Chandiramouli, R. Investigation of NH3 Adsorption Behavior on Graphdiyne Nanosheet and Nanotubes: A First-Principles Study. J. Mol. Liq. 2018, 249, 24–32. [Google Scholar] [CrossRef]
  89. Bhuvaneswari, R.; Princy Maria, J.; Nagarajan, V.; Chandiramouli, R. Graphdiyne Nanosheets as a Sensing Medium for Formaldehyde and Formic Acid—A First-Principles Outlook. Comput. Theor. Chem. 2020, 1176, 112751. [Google Scholar] [CrossRef]
  90. Lv, Q.; Si, W.; He, J.; Sun, L.; Zhang, C.; Wang, N.; Yang, Z.; Li, X.; Wang, X.; Deng, W.; et al. Selectively Nitrogen-Doped Carbon Materials as Superior Metal-Free Catalysts for Oxygen Reduction. Nat. Commun. 2018, 9, 3376. [Google Scholar] [CrossRef] [Green Version]
  91. Song, M.; Chen, Y.; Liu, X.; Xu, W.; Zhao, Y.; Zhang, M.; Zhang, C. A First-Principles Study of Gas Molecule Adsorption on Hydrogen-Substituted Graphdiyne. Phys. Lett. A 2020, 384, 126332. [Google Scholar] [CrossRef]
  92. Xu, P.; Na, N.; Mohamadi, A. Investigation the Application of Pristine Graphdiyne (GDY) and Boron-Doped Graphdiyne (BGDY) as an Electronic Sensor for Detection of Anticancer Drug. Comput. Theor. Chem. 2020, 1190, 112996. [Google Scholar] [CrossRef]
  93. Yuan, J.; Mohamadi, A. Study the Adsorption Process of 5-Fluorouracil Drug on the Pristine and Doped Graphdiyne Nanosheet. J. Mol. Model. 2021, 27, 32. [Google Scholar] [CrossRef]
  94. Feng, Z.; Ma, Y.; Li, Y.; Li, R.; Tang, Y.; Dai, X. Oxygen Molecule Dissociation on Heteroatom Doped Graphdiyne. Appl. Surf. Sci. 2019, 494, 421–429. [Google Scholar] [CrossRef]
  95. Rangel, E.; Ramirez-de-Arellano, J.M.; Magana, L.F. Variation of Hydrogen Adsorption with Increasing Li Doping on Carbon Nanotubes: Variation of Hydrogen Adsorption with Increasing Li Doping on CNTs. Phys. Status Solidi B 2011, 248, 1420–1424. [Google Scholar] [CrossRef]
  96. Rangel, E.; Ramírez-Arellano, J.M.; Carrillo, I.; Magana, L.F. Hydrogen Adsorption around Lithium Atoms Anchored on Graphene Vacancies. Int. J. Hydrog. Energy 2011, 36, 13657–13662. [Google Scholar] [CrossRef]
  97. Kim, J.; Kang, S.; Lim, J.; Kim, W.Y. Study of Li Adsorption on Graphdiyne Using Hybrid DFT Calculations. ACS Appl. Mater. Interfaces 2019, 11, 2677–2683. [Google Scholar] [CrossRef] [PubMed]
  98. Feng, Z.; Su, G.; Ding, H.; Ma, Y.; Li, Y.; Tang, Y.; Dai, X. Atomic Alkali Metal Anchoring on Graphdiyne as Single-Atom Catalysts for Capture and Conversion of CO2 to HCOOH. Mol. Catal. 2020, 494, 111142. [Google Scholar] [CrossRef]
  99. Liu, T.; Wang, Q.; Wang, G.; Bao, X. Electrochemical CO2 Reduction on Graphdiyne: A DFT Study. Green Chem. 2021, 23, 1212–1219. [Google Scholar] [CrossRef]
  100. Feng, Z.; Tang, Y.; Ma, Y.; Li, Y.; Dai, Y.; Ding, H.; Su, G.; Dai, X. Theoretical Investigation of CO2 Electroreduction on N (B)-Doped Graphdiyne Mononlayer Supported Single Copper Atom. Appl. Surf. Sci. 2021, 538, 148145. [Google Scholar] [CrossRef]
  101. Wu, Y.; Chen, X.; Weng, K.; Jiang, J.; Ong, W.; Zhang, P.; Zhao, X.; Li, N. Highly Sensitive and Selective Gas Sensor Using Heteroatom Doping Graphdiyne: A DFT Study. Adv. Electron. Mater. 2021, 7, 2001244. [Google Scholar] [CrossRef]
  102. Cao, J.; Li, N.; Zeng, X. Exploring the Synergistic Effect of B–N Doped Defective Graphdiyne for N2 Fixation. New J. Chem. 2021, 45, 6327–6335. [Google Scholar] [CrossRef]
  103. Chen, X.; Lin, Z.-Z. Single-Layer Graphdiyne-Covered Pt(111) Surface: Improved Catalysis Confined under Two-Dimensional Overlayer. J. Nanopart. Res. 2018, 20, 136. [Google Scholar] [CrossRef]
  104. Khan, S.; Yar, M.; Kosar, N.; Ayub, K.; Arshad, M.; Zahid, M.N.; Mahmood, T. First-Principles Study for Exploring the Adsorption Behavior of G-Series Nerve Agents on Graphdyine Surface. Comput. Theor. Chem. 2020, 1191, 113043. [Google Scholar] [CrossRef]
  105. Sajid, H.; Khan, S.; Ayub, K.; Mahmood, T. Effective Adsorption of A-Series Chemical Warfare Agents on Graphdiyne Nanoflake: A DFT Study. J. Mol. Model. 2021, 27, 117. [Google Scholar] [CrossRef] [PubMed]
  106. Khan, S.; Sajid, H.; Ayub, K.; Mahmood, T. Sensing of Toxic Lewisite (L1, L2, and L3) Molecules by Graphdiyne Nanoflake Using Density Functional Theory Calculations and Quantum Theory of Atoms in Molecule Analysis. J. Phys. Org. Chem. 2021, 34, e4181. [Google Scholar] [CrossRef]
  107. Mashhadzadeh, A.H.; Vahedi, A.M.; Ardjmand, M.; Ahangari, M.G. Investigation of Heavy Metal Atoms Adsorption onto Graphene and Graphdiyne Surface: A Density Functional Theory Study. Superlattices Microstruct. 2016, 100, 1094–1102. [Google Scholar] [CrossRef]
  108. Seif, A.; López, M.J.; Granja-DelRío, A.; Azizi, K.; Alonso, J.A. Adsorption and Growth of Palladium Clusters on Graphdiyne. Phys. Chem. Chem. Phys. 2017, 19, 19094–19102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Wang, X.; Ma, C.; Chen, K.; Li, H.; Wang, P. Interaction between Nanobuds and Hydrogen Molecules: A First-Principles Study. Phys. Lett. A 2009, 374, 87–90. [Google Scholar] [CrossRef]
  110. Wu, C.-D.; Fang, T.-H.; Lo, J.-Y. Effects of Pressure, Temperature, and Geometric Structure of Pillared Graphene on Hydrogen Storage Capacity. Int. J. Hydrog. Energy 2012, 37, 14211–14216. [Google Scholar] [CrossRef]
  111. Hassani, A.; Hamed Mosavian, M.T.; Ahmadpour, A.; Farhadian, N. Hybrid Molecular Simulation of Methane Storage inside Pillared Graphene. J. Chem. Phys. 2015, 142, 234704. [Google Scholar] [CrossRef]
  112. Baykasoglu, C.; Mert, H.; Deniz, C.U. Grand Canonical Monte Carlo Simulations of Methane Adsorption in Fullerene Pillared Graphene Nanocomposites. J. Mol. Graph. Model. 2021, 106, 107909. [Google Scholar] [CrossRef] [PubMed]
  113. Ozturk, Z.; Baykasoglu, C.; Kirca, M. Sandwiched Graphene-Fullerene Composite: A Novel 3-D Nanostructured Material for Hydrogen Storage. Int. J. Hydrog. Energy 2016, 41, 6403–6411. [Google Scholar] [CrossRef]
  114. Ugarte, D. Curling and Closure of Graphitic Networks under Electron-Beam Irradiation. Nature 1992, 359, 707–709. [Google Scholar] [CrossRef]
  115. Goclon, J.; Bankiewicz, B.; Kolek, P.; Winkler, K. Role of Nitrogen Doping in Stoichiometric and Defective Carbon Nano-Onions: Structural Diversity from DFT Calculations. Carbon 2021, 176, 198–208. [Google Scholar] [CrossRef]
  116. Hussain, M.A.; Vijay, D.; Sastry, G.N. Buckybowls as Adsorbents for CO2, CH4, and C2H2: Binding and Structural Insights from Computational Study. J. Comput. Chem. 2016, 37, 366–377. [Google Scholar] [CrossRef]
  117. Zhang, C.; Li, J.; Liu, E.; He, C.; Shi, C.; Du, X.; Hauge, R.H.; Zhao, N. Synthesis of Hollow Carbon Nano-Onions and Their Use for Electrochemical Hydrogen Storage. Carbon 2012, 50, 3513–3521. [Google Scholar] [CrossRef]
  118. Nikmaram, F.R.; Khoddamzadeh, A. Chemical Shielding of Doped Nitrogen on C20 Cage and Bowl Fullerenes. J. Struct. Chem. 2017, 58, 173–177. [Google Scholar] [CrossRef]
  119. Zhang, W.-D.; Zhang, W.-H. Carbon Nanotubes as Active Components for Gas Sensors. J. Sens. 2009, 2009, 160698. [Google Scholar] [CrossRef] [Green Version]
  120. Kong, J.; Chapline, M.G.; Dai, H. Functionalized Carbon Nanotubes for Molecular Hydrogen Sensors. Adv. Mater. 2001, 13, 1384–1386. [Google Scholar] [CrossRef]
  121. Sayago, I.; Terrado, E.; Aleixandre, M.; Horrillo, M.C.; Fernández, M.J.; Lozano, J.; Lafuente, E.; Maser, W.K.; Benito, A.M.; Martinez, M.T.; et al. Novel Selective Sensors Based on Carbon Nanotube Films for Hydrogen Detection. Sens. Actuators B Chem. 2007, 122, 75–80. [Google Scholar] [CrossRef]
  122. Mubeen, S.; Zhang, T.; Yoo, B.; Deshusses, M.A.; Myung, N.V. Palladium Nanoparticles Decorated Single-Walled Carbon Nanotube Hydrogen Sensor. J. Phys. Chem. C 2007, 111, 6321–6327. [Google Scholar] [CrossRef]
  123. Sun, Y.; Wang, H.H. Electrodeposition of Pd Nanoparticles on Single-Walled Carbon Nanotubes for Flexible Hydrogen Sensors. Appl. Phys. Lett. 2007, 90, 213107. [Google Scholar] [CrossRef]
  124. Sippel-Oakley, J.; Wang, H.-T.; Kang, B.S.; Wu, Z.; Ren, F.; Rinzler, A.G.; Pearton, S.J. Carbon Nanotube Films for Room Temperature Hydrogen Sensing. Nanotechnology 2005, 16, 2218–2221. [Google Scholar] [CrossRef] [PubMed]
  125. Ding, D.; Chen, Z.; Rajaputra, S.; Singh, V. Hydrogen Sensors Based on Aligned Carbon Nanotubes in an Anodic Aluminum Oxide Template with Palladium as a Top Electrode. Sens. Actuators B Chem. 2007, 124, 12–17. [Google Scholar] [CrossRef]
  126. Lu, Y.; Li, J.; Han, J.; Ng, H.-T.; Binder, C.; Partridge, C.; Meyyappan, M. Room Temperature Methane Detection Using Palladium Loaded Single-Walled Carbon Nanotube Sensors. Chem. Phys. Lett. 2004, 391, 344–348. [Google Scholar] [CrossRef]
  127. Li, Y.; Wang, H.; Chen, Y.; Yang, M. A Multi-Walled Carbon Nanotube/Palladium Nanocomposite Prepared by a Facile Method for the Detection of Methane at Room Temperature. Sens. Actuators B Chem. 2008, 132, 155–158. [Google Scholar] [CrossRef]
  128. Kumar, M.K.; Ramaprabhu, S. Nanostructured Pt Functionlized Multiwalled Carbon Nanotube Based Hydrogen Sensor. J. Phys. Chem. B 2006, 110, 11291–11298. [Google Scholar] [CrossRef]
  129. Krishnakumar, M.; Ramaprabhu, S. Palladium Dispersed Multiwalled Carbon Nanotube Based Hydrogen Sensor for Fuel Cell Applications. Int. J. Hydrog. Energy 2007, 32, 2518–2526. [Google Scholar] [CrossRef]
  130. Krishna Kumar, M.; Leela Mohana Reddy, A.; Ramaprabhu, S. Exfoliated Single-Walled Carbon Nanotube-Based Hydrogen Sensor. Sens. Actuators B Chem. 2008, 130, 653–660. [Google Scholar] [CrossRef]
  131. Kamarchuk, G.V.; Kolobov, I.G.; Khotkevich, A.V.; Yanson, I.K.; Pospelov, A.P.; Levitsky, I.A.; Euler, W.B. New Chemical Sensors Based on Point Heterocontact between Single Wall Carbon Nanotubes and Gold Wires. Sens. Actuators B Chem. 2008, 134, 1022–1026. [Google Scholar] [CrossRef]
  132. Penza, M.; Cassano, G.; Rossi, R.; Alvisi, M.; Rizzo, A.; Signore, M.A.; Dikonimos, T.; Serra, E.; Giorgi, R. Enhancement of Sensitivity in Gas Chemiresistors Based on Carbon Nanotube Surface Functionalized with Noble Metal (Au, Pt) Nanoclusters. Appl. Phys. Lett. 2007, 90, 173123. [Google Scholar] [CrossRef]
  133. Espinosa, E.H.; Ionescu, R.; Bittencourt, C.; Felten, A.; Erni, R.; Van Tendeloo, G.; Pireaux, J.-J.; Llobet, E. Metal-Decorated Multi-Wall Carbon Nanotubes for Low Temperature Gas Sensing. Thin Solid Film. 2007, 515, 8322–8327. [Google Scholar] [CrossRef]
  134. Star, A.; Joshi, V.; Skarupo, S.; Thomas, D.; Gabriel, J.-C.P. Gas Sensor Array Based on Metal-Decorated Carbon Nanotubes. J. Phys. Chem. B 2006, 110, 21014–21020. [Google Scholar] [CrossRef] [PubMed]
  135. Lu, Y.; Partridge, C.; Meyyappan, M.; Li, J. A Carbon Nanotube Sensor Array for Sensitive Gas Discrimination Using Principal Component Analysis. J. Electroanal. Chem. 2006, 593, 105–110. [Google Scholar] [CrossRef]
  136. Kwon, Y.J.; Na, H.G.; Kang, S.Y.; Choi, S.-W.; Kim, S.S.; Kim, H.W. Selective Detection of Low Concentration Toluene Gas Using Pt-Decorated Carbon Nanotubes Sensors. Sens. Actuators B Chem. 2016, 227, 157–168. [Google Scholar] [CrossRef]
  137. Tabtimsai, C.; Keawwangchai, S.; Nunthaboot, N.; Ruangpornvisuti, V.; Wanno, B. Density Functional Investigation of Hydrogen Gas Adsorption on Fe−doped Pristine and Stone−Wales Defected Single−walled Carbon Nanotubes. J. Mol. Model. 2012, 18, 3941–3949. [Google Scholar] [CrossRef]
  138. Rather, S. Hydrogen Uptake of Ti-Decorated Multiwalled Carbon Nanotube Composites. Int. J. Hydrog. Energy 2021, 46, 17793–17801. [Google Scholar] [CrossRef]
  139. Dixit, S.; Patodia, T.; Sharma, K.B.; Katyayan, S.; Dixit, A.; Jain, S.K.; Agarwal, G.; Tripathi, B. Adsorption Characteristics of MWNTs via Intercalation of Nickel. Mater. Today Proc. 2021, 38, 1233–1236. [Google Scholar] [CrossRef]
  140. Zhang, X.; Dai, Z.; Wei, L.; Liang, N.; Wu, X. Theoretical Calculation of the Gas-Sensing Properties of Pt-Decorated Carbon Nanotubes. Sensors 2013, 13, 15159–15171. [Google Scholar] [CrossRef]
  141. Zhang, X.; Dai, Z.; Chen, Q.; Tang, J. A DFT Study of SO2 and H2S Gas Adsorption on Au-Doped Single-Walled Carbon Nanotubes. Phys. Scr. 2014, 89, 065803. [Google Scholar] [CrossRef]
  142. Lam, A.D.K.-T.; Lin, Z.-D.; Lu, H.-Y.; Young, S.-J. Carbon Nanotubes with Adsorbed Au Nanoparticles for Sensing Propanone Gas. Microsyst. Technol. 2019, 1–4. [Google Scholar] [CrossRef]
  143. Li, S.; Jiang, J. Adsorption Behavior Analyses of Several Small Gas Molecules onto Rh-Doped Single-Walled Carbon Nanotubes. Appl. Phys. A 2017, 123, 669. [Google Scholar] [CrossRef]
  144. Kuganathan, N.; Chroneos, A. Ru-Doped Single Walled Carbon Nanotubes as Sensors for SO2 and H2S Detection. Chemosensors 2021, 9, 120. [Google Scholar] [CrossRef]
  145. Li, K.; Wang, W.; Cao, D. Metal (Pd, Pt)-Decorated Carbon Nanotubes for CO and NO Sensing. Sens. Actuators B Chem. 2011, 159, 171–177. [Google Scholar] [CrossRef]
  146. Yoosefian, M. Powerful Greenhouse Gas Nitrous Oxide Adsorption onto Intrinsic and Pd Doped Single Walled Carbon Nanotube. Appl. Surf. Sci. 2017, 392, 225–230. [Google Scholar] [CrossRef]
  147. Cui, H.; Zhang, X.; Yao, Q.; Miao, Y.; Tang, J. Rh-Doped Carbon Nanotubes as a Superior Media for the Adsorption of O2 and O3 Molecules: A Density Functional Theory Study. Carbon Lett. 2018, 28, 55–59. [Google Scholar] [CrossRef]
  148. Young, S.J.; Lin, Z.D. Acetone Gas Sensors Composed of Carbon Nanotubes with Adsorbed Au Nanoparticles on Plastic Substrate. Microsyst. Technol. 2018, 24, 3973–3976. [Google Scholar] [CrossRef]
  149. Pei, P.; Whitwick, M.B.; Kureshi, S.; Cannon, M.; Quan, G.; Kjeang, E. Hydrogen Storage Mechanism in Transition Metal Decorated Graphene Oxide: The Symbiotic Effect of Oxygen Groups and High Layer Spacing. Int. J. Hydrog. Energy 2020, 45, 6713–6726. [Google Scholar] [CrossRef]
  150. Singla, M.; Jaggi, N. Enhanced Hydrogen Sensing Properties in Copper Decorated Nitrogen Doped Defective Graphene Nanoribbons: DFT Study. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 131, 114756. [Google Scholar] [CrossRef]
  151. Ni, J.; Quintana, M.; Song, S. Adsorption of Small Gas Molecules on Transition Metal (Fe, Ni and Co, Cu) Doped Graphene: A Systematic DFT Study. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 116, 113768. [Google Scholar] [CrossRef]
  152. Zitoune, H.; Adessi, C.; Benchallal, L.; Samah, M. Quantum Transport Properties of Gas Molecules Adsorbed on Fe Doped Armchair Graphene Nanoribbons: A First Principle Study. J. Phys. Chem. Solids 2021, 153, 109996. [Google Scholar] [CrossRef]
  153. Kuang, A.; Mo, M.; Kuang, M.; Wang, B.; Tian, C.; Yuan, H.; Wang, G.; Chen, H. The Comparative Study of XO2 (X = C, N, S) Gases Adsorption and Dissociation on Pristine and Defective Graphene Supported Pt13. Mater. Chem. Phys. 2020, 247, 122712. [Google Scholar] [CrossRef]
  154. Yang, L.; Xiao, W.; Wang, J.; Li, X.; Wang, L. Formaldehyde Gas Sensing Properties of Transition Metal-Doped Graphene: A First-Principles Study. J. Mater. Sci. 2021, 56, 12256–12269. [Google Scholar] [CrossRef]
  155. Cai, Y.; Luo, X. First-Principles Investigation of Carbon Dioxide Adsorption on MN4 Doped Graphene. AIP Adv. 2020, 10, 125013. [Google Scholar] [CrossRef]
  156. Guo, J.; Liu, Z.; Liu, S.; Zhao, X.; Huang, K. High-Capacity Hydrogen Storage Medium: Ti Doped Fullerene. Appl. Phys. Lett. 2011, 98, 023107. [Google Scholar] [CrossRef]
  157. Germán, E.; Alonso, J.A.; Janssens, E.; López, M.J. C60Con Complexes as Hydrogen Adsorbing Materials. Int. J. Hydrog. Energy 2021, 46, 20594–20606. [Google Scholar] [CrossRef]
  158. Arshadi, S.; Anisheh, F. Theoretical Study of Cr and Co- Porphyrin-Induced C70 Fullerene: A Request for a Novel Sensor of Sulfur and Nitrogen Dioxide. J. Sulfur Chem. 2017, 38, 357–371. [Google Scholar] [CrossRef]
  159. Chen, X.; Zhang, H.; Lai, N. Endohedral Metallofullerenes Mn@C60 (M = Mn, Co, Ni, Cu; n = 2–5) as Electrocatalysts for Oxygen Reduction Reaction: A First-Principles Study. J. Mater. Sci. 2020, 55, 11382–11390. [Google Scholar] [CrossRef]
  160. Zou, L.; Zhu, Y.; Cen, W.; Jiang, X.; Chu, W. N-Doping in Graphdiyne on Embedding of Metals and Its Effect in Catalysis. Appl. Surf. Sci. 2021, 557, 149815. [Google Scholar] [CrossRef]
  161. Panigrahi, P.; Dhinakaran, A.K.; Naqvi, S.R.; Gollu, S.R.; Ahuja, R.; Hussain, T. Light Metal Decorated Graphdiyne Nanosheets for Reversible Hydrogen Storage. Nanotechnology 2018, 29, 355401. [Google Scholar] [CrossRef]
  162. Yu, H.; Hui, L.; Xue, Y.; Liu, Y.; Fang, Y.; Xing, C.; Zhang, C.; Zhang, D.; Chen, X.; Du, Y.; et al. 2D Graphdiyne Loading Ruthenium Atoms for High Efficiency Water Splitting. Nano Energy 2020, 72, 104667. [Google Scholar] [CrossRef]
  163. Zhai, X.; Yan, H.; Ge, G.; Yang, J.; Chen, F.; Liu, X.; Yang, D.; Li, L.; Zhang, J. The Single-Mo-Atom-Embedded-Graphdiyne Monolayer with Ultra-Low Onset Potential as High Efficient Electrocatalyst for N2 Reduction Reaction. Appl. Surf. Sci. 2020, 506, 144941. [Google Scholar] [CrossRef]
  164. Jasin Arachchige, L.; Xu, Y.; Dai, Z.; Zhang, X.L.; Wang, F.; Sun, C. Double Transition Metal Atoms Anchored on Graphdiyne as Promising Catalyst for Electrochemical Nitrogen Reduction Reaction. J. Mater. Sci. Technol. 2021, 77, 244–251. [Google Scholar] [CrossRef]
  165. Chen, X.; Gao, P.; Guo, L.; Wen, Y.; Fang, D.; Gong, B.; Zhang, Y.; Zhang, S. High-Efficient Physical Adsorption and Detection of Formaldehyde Using Sc- and Ti-Decorated Graphdiyne. Phys. Lett. A 2017, 381, 879–885. [Google Scholar] [CrossRef]
  166. Liu, J.; Wang, L.; Shen, X.; Gao, X.; Chen, Y.; Liu, H.; Liu, Y.; Yin, D.; Liu, Y.; Xu, W.; et al. Graphdiyne-Templated Palladium-Nanoparticle Assembly as a Robust Oxygen Generator to Attenuate Tumor Hypoxia. Nano Today 2020, 34, 100907. [Google Scholar] [CrossRef]
  167. He, T.; Zhang, L.; Kour, G.; Du, A. Electrochemical Reduction of Carbon Dioxide on Precise Number of Fe Atoms Anchored Graphdiyne. J. CO2 Util. 2020, 37, 272–277. [Google Scholar] [CrossRef]
  168. Liu, X.; Tang, W.; Liu, S.; Chen, X.; Li, Y.; Hu, X.; Qiao, L.; Zeng, Y. CO Oxidation on Ni and Cu Embedded Graphdiyne as Efficient Noble Metal-Free Catalysts: A First-Principles Density-Functional Theory Investigation. Appl. Surf. Sci. 2021, 539, 148287. [Google Scholar] [CrossRef]
  169. Vijay, D.; Sakurai, H.; Subramanian, V.; Sastry, G.N. Where to Bind in Buckybowls? The Dilemma of a Metal Ion. Phys. Chem. Chem. Phys. 2012, 14, 3057. [Google Scholar] [CrossRef] [PubMed]
  170. Zhang, C.; Li, J.; Shi, C.; He, C.; Liu, E.; Zhao, N. Effect of Ni, Fe and Fe-Ni Alloy Catalysts on the Synthesis of Metal Contained Carbon Nano-Onions and Studies of Their Electrochemical Hydrogen Storage Properties. J. Energy Chem. 2014, 23, 324–330. [Google Scholar] [CrossRef]
  171. Canales, M.; Ramírez-de-Arellano, J.M.; Magana, L.F. Interaction of a Ti-Doped Semi-Fullerene (TiC30) with Molecules of CO and CO2. J. Mol. Model. 2016, 22, 223. [Google Scholar] [CrossRef]
  172. Canales-Lizaola, M.; Arellano, J.S.; Magaña, L.F. Hydrogen Molecule Adsorption on a Ti-Doped Graphene+ Semi-Fullerene Surface. J. Phys. Conf. Ser. 2019, 1221, 012081. [Google Scholar] [CrossRef]
Figure 1. These are some of the carbon nanostructures reported experimentally. In (a,b), we have C60; we present the C180 in (c); C20 in (d); the buckybowl C30 in (e,f); graphene in (g); graphdiyne in (h); a single-walled carbon nanotube (SWCNT) in (i).
Figure 1. These are some of the carbon nanostructures reported experimentally. In (a,b), we have C60; we present the C180 in (c); C20 in (d); the buckybowl C30 in (e,f); graphene in (g); graphdiyne in (h); a single-walled carbon nanotube (SWCNT) in (i).
Molecules 26 05346 g001
Figure 2. We present some of the hybrid carbon nanostructures reported experimentally. In (ad), we have a C30 (a buckybowl) adsorbed on graphene; in (e), we show a fullerene C60, adsorbed on graphene; we show in (f), a bi-layer graphene; in (g,h), we present pillared graphene with nanotubes (SWCNTs); in (i) we show pillared graphene with hydrogen; we present in (j,k), two examples of carbon nano onions (CNO).
Figure 2. We present some of the hybrid carbon nanostructures reported experimentally. In (ad), we have a C30 (a buckybowl) adsorbed on graphene; in (e), we show a fullerene C60, adsorbed on graphene; we show in (f), a bi-layer graphene; in (g,h), we present pillared graphene with nanotubes (SWCNTs); in (i) we show pillared graphene with hydrogen; we present in (j,k), two examples of carbon nano onions (CNO).
Molecules 26 05346 g002
Table 1. Summary of adsorption energies Eads (in eV) for different non-metallic doped nanotubes systems and adsorbates. The type of study is also specified.
Table 1. Summary of adsorption energies Eads (in eV) for different non-metallic doped nanotubes systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
CNTs [41]DFTNO2
CH4
CO2
−0.427
−0.122
−0.109
SWCNTs [39]GCMC, MDSO2−0.464
MWCNTs [11]ExperimentalH2[−0.26, −0.046]
MWNTs [28]ExperimentalSF6[−0.529, −1.285]
SWCNTs [37]DFTNO3−1.30
CNNT [38]DFTH2CO−0.321
SWNTs and bundles [10]DFTNO2
O2
H2O
NH3
N2
CO2
CH4
Ar
−0.427
−0.306
−0.128
−0.162
−0.123
−0.109
−0.122
−0.082
SWCNTs [12]ExperimentalH2−0.056
SWCNTs and SiC-DC [27]GCMCCO2
CH4
−0.005
−0.003
SWCNTs [30]Exp. and DFTC3H6O[−0.255, −0.771]
CNTs [31]ExperimentalC2H5OHNot reported
CNTs films [19]Exp. and GCMCCO2Not reported
MWCNTs [40]ExperimentalVOCNot reported
SWCNTs [36]Exp., GCMC, MDXeNot reported
F-MWCNTs [24]ExperimentalCO2[−0.084, −0.036]
F-MWCNTs [13]ExperimentalH2Not reported
Oxygen F-CNTs [26]ExperimentalC2, N2Not reported
PANI/MWCNTs [32]ExperimentalNH3Not reported
F-MWCNTs and A-MWCNTs [14]ExperimentalH2Not reported
1,3-diaminopropane MWCNTs [22]ExperimentalCO2Not reported
Table 2. Summary of adsorption energies Eads (in eV) for different non-metallic doped graphene systems and adsorbates. The type of study is also specified.
Table 2. Summary of adsorption energies Eads (in eV) for different non-metallic doped graphene systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
ZGNR [51]DFTH2S−0.364
Pristine G (PG) [52]DFTSO2−0.157
PG [49]Exp. and MDNH3Not reported
PG [44]DFT and MDH2Not reported
Graphene [48]DFTH2S
CH4
−0.038
−0.022
DG (vacancy) [48]DFTH2S
CH4
−2.934
−0.154
G-OH [48]DFTH2S
CH4
−1.263
−0.047
Single layer G (SLG) [43]ExperimentalH2Not reported
Bilayer G (BG) [53]DFTH2S−0.360
PG [54]DFTCH4
C2H2
CO
−0.086
−0.102
−0.093
PG [46]DFTNO2
H2S
[−0.214, −0.185]
[−0.201, −0.122]
PG [55]DFTCO
NO
−0.084
−1.166
Vacancy G (VG) [55]DFTCO
NO
−0.069
−1.203
PG [56]DFTNO
NO2
[−0.1280, −0.1176]
[−0.1679, −0.1427]
PG [57]DFTAsH3
CO
Not reported
PG [58]DFTH2S−0.360
PG supercell [47]DFTNO[−2.3713, −1.7453]
PG [59]DFTCOCl2−0.554
G nanoflakes (GNFs) [60]DFTCO
CO2
−1.18
−0.58
PG [45]DFTH2O
NH3
CO
NO2
NO
−0.047
−0.031
−0.014
−0.067
−0.029
PG [42]DFTH[−0.84, −0.75]
Table 3. Summary of adsorption energies Eads (in eV) for different non-metallic doped fullerene systems and adsorbates. The type of study is also specified.
Table 3. Summary of adsorption energies Eads (in eV) for different non-metallic doped fullerene systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
C60 [70]Exp. and DFTH2[0.0495, 0.0641]
C60 [74]PES (potential energy surface)Ne, Ar, Kr, XeNot reported
C60 [69]ExperimentalH−0.0247
C60 [72]DFTN2[−0.28, −0.03]
C60 [64]GCMCC2H4[−0.0207, 0.0207]
C60 [79]DFTCO
NO
−0.006
0.00008
C60 [76]DFTC2N2−4.78
OH-C60 [76]DFTC2N2−5.13
C60 [77]DFTC9H13N + N
C9H13N + H
−0.017
−0.0516
C32 [71]DFTH2[−0.118, −0.0086]
C60 [65]ExperimentalAromatic vaporNot reported
C60-Nx [73]DFTNO
NO2
[−0.35, −0.28]
[−1.01, −0.94]
C24 + P + N24 [80]DFTCO2[−0.94, −0.30]
C60+P [63]DFTCO2[−1.97, 0.06]
Si@C54N4 [66]DFTN2O
CO
O2
[−3.50, −0–77]
C60 [67]DFTN2O
CO
−0.169
−0.092
C20 [75]
B-C20 [75]
N-C20 [75]
DFTC5H11N2O2P[−0.054, −0.043]
[−1.092, −0.072]
[−0.027, −1.65]
C460 [62]DFTH2
CO2
−4.37
[−0.49, −0.42]
C60 [68]Modified LJ-potentialH2, HeNot reported
Table 4. Summary of adsorption energies Eads (in eV) for different non-metallic doped graphdiyne systems and adsorbates. The type of study is also specified.
Table 4. Summary of adsorption energies Eads (in eV) for different non-metallic doped graphdiyne systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
GDY [86]DFTDMA
TMA
[−0.503, −0.757]
[−0.607, −0.796]
GDY supercell [85]DFTC2H5NO2
C5H9NO4
C6H9N3O2
C9H11NO2
[−1.10, −0.59]
[−1.14, −0.54]
[−1.46. −0.73]
[−1.53, −0.77]
GDY, BGDY [92]DFTTMZ[−1.97, −0.95]
GDY nanoflakes [105]DFTCWA A-230
CWA A-232
CWA A-234
−0.594
−0.713
−0.745
GDY [81]DFTCO
O2
−1.43
−3.27
GDY-NS [87]DFTC21H16ClF3N4O3
C21H15ClF4N4O3
[−0.660, −0.085]
[−0.641, −0.081]
GDY [89]DFTCH2O
CH2O2
[−1.502, −0.342]
[−0.945, −0.390]
GDY [107]DFTAg
Cu
Ni
Zn
[−0.792926, −1.236]
[−0.622651, −2.783]
[−2.913467, −3.446]
[−0.0196, 0.0356]
GDY [84]DFTH
O
−3.73
−7.53
GDY nanoflakes [106]DFTL1
L2
L3
−0.441
−0.534
−0.567
GDY [97]Hybrid DFTLi−1.82
Ca-GDY [83]DFTCH4O
CO
[−0.349, −0.122]
[−0.128, −0.060]
GDY [104]DFT, QTAIMCWA GA
CWA GB
CWA GD
CWA GF
−0.707
−0.520
−0.543
−0.382
GDY [88]DFTNH3[−0.465, −0.435]
GDY [108]DFTPd clusters[−4.0, −3.0]
Table 5. Summary of adsorption energies Eads (in eV) for different non-metallic doped hybrid systems and adsorbates. The type of study is also specified.
Table 5. Summary of adsorption energies Eads (in eV) for different non-metallic doped hybrid systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
CNO [117]ExperimentalH2ONot reported
CNO [115]DFTmN, m = 1–10[−0.18, 0.33]
Buckybowls [116]Hybrid-DFTCO2, CH4, C2H2Not reported
C20, C20 (bowl) [118]DFTN, HNot reported
Nanobuds [109]DFTH2[0.069, 0.115]
Sandwiched G-fullerene + Li [113]GCMCH2Not reported
Pillared-graphene [110]MDH2Not reported
Fullerene pillared-graphene [112]GCMCCH4Not reported
Pillared-graphene [111]GCMC-MDCH4Not reported
Table 6. List of transition metals used for doping carbon nanotubes and the corresponding adsorbed gases.
Table 6. List of transition metals used for doping carbon nanotubes and the corresponding adsorbed gases.
MetalAdsorbed Molecules
AgNO2 [133]
AuNH3, NH2 [132,133,135]; SO2, H2S [141]; C2H6O [31]; C3H6O [142,148]
PtH2, NO2, H2O, NH3 [132]; C7H8 [136]; NO [145]
PdH2 [120,121,122,123,124,125,128]; NO2 [135], Cl2 [135]; CH4 [126,127]; NO [145]; N2O [146]
FeH2 [137]
RhCO [134]; CO, C2H4 SO2 [143]; O2, O3 [147]
Table 7. Summary of adsorption energies Eads (in eV) for different TM-doped nanotubes systems and adsorbates. The type of study is also specified.
Table 7. Summary of adsorption energies Eads (in eV) for different TM-doped nanotubes systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
CNT(APTS) [18]ExperimentalCO2Not reported
Fe-SWCNTs [137]DFTH2[−0.238, −0.141]
Au-CNTs [31]ExperimentalC2H6ONot reported
Pt-SWCNTs [140]DFTSO2
H2S
CO
−1.225
−0.977
−1.386
Au-SWCNTs [141]DFTSO2
H2S
−1.258
−1.317
Au-CNTs [142]ExperimentalC3H6ONot reported
Rh–CNT [143]DFTCO
CO2
CH4
C2H4
SO2
[−3.527, −1.308]
−0.348
−0.253
−1.189
−1.158
Pd-, Pt-SWNTs [145]DFTCO
NO
[−1.8, −1.6]
[−1.814, −1.46]
Pd-CNT [146]DFTN2O−0.91
Rh–CNT [147]DFTO2
O3
−1.384
[−2.711, −1.824]
Pt-MWCNTs [136]ExperimentalC7H8Not reported
Au-CNTs [148]ExperimentalC3H6ONot reported
Pd-, Pt-, Rh-, Au-SWNTs [134]ExperimentalH2, CH4, CO, H2SNot reported
Table 8. Summary of adsorption energies Eads (in eV) for different TM-doped graphene systems and adsorbates. The type of study is also specified.
Table 8. Summary of adsorption energies Eads (in eV) for different TM-doped graphene systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
Zn-ZGNR [51]DFTH2S−2.237
Cu-ZGNR [51]DFTH2S−1.129
Cu/Zn-ZGNR [51]DFTH2S−7.043
Ni-G [52]DFTSO2−2.297
TM-PG [154]
TM = Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn, Pd, Ag, Pt, Au
DFTHCHO[0.83, 2.01]
Cu+1N-GNR [150]DFTH2−0.020
Cu+2N-GNR [150]DFTH2−0.200
Cu+3N-GNR [150]DFTH2−0.780
TiO2-GO [149]ExperimentalH2Not reported
Ni-G [48]DFTH2S
CH4
−0.699
−0.099
TM-BG [53]
TM = V, Cr, Mn, Fe, Co, Ni
DFTH2S[−0.58, −0.18]
Pt13-G [153]DFTCO2
NO2
SO2
[−4.217, −2.422]
[−3.767, −2.586]
[−3.260, −2.238]
Pt13-DG [153]DFTCO2
NO2
SO2
[−3.201, −0.916]
[−3.345, −2.309]
[−2.978, −2.065]
Mn-G [54]DFTCH4
C2H2
CO
−0.073
−2.424
−1.954
Ni-G [46]DFTNO2
H2S
[−2.631, −2.395]
[−1.846, −1.811]
Pd-G [46]DFTNO2
H2S
[−1.586, −1.294]
[−1.228, −1.224]
Pt-G [46]DFTNO2
H2S
[−2.003, −1.804]
[−2.034, −1.858]
Pd-G [55]DFTCO
NO
[−1.227, −0.909]
[−3.916, −1.308]
Ag-G [56]DFTNO
NO2
[−6.9262, −6.9101]
[−7.8293, −7.7806]
Pt-G [56]DFTNO
NO2
[−6.2225, −6.1646]
[−7.3758, −7.3723]
Au-G [56]DFTNO
NO2
[−8.4730, −8.3567]
[−9.3391, −9.3209]
TM-G [57]
TM = Ti, Mn, Fe, Co, Ni, Ag
DFTAsH3
CO
[−0.95, −1.45]
[−1.00, 2.02]
Ni-G [58]DFTH2S−0.97
Cu-G [58]DFTH2S−1.15
Zn-G [58]DFTH2S−1.16
TM-G [151]
TM = Fe, Ni, Co, Cu
DFTCO2
NO
NO2
SO2
[−0.89, −1.19]
[−0.68, −1.23]
[−2.06, −2.57]
[−0.89, −1.47]
Fe- AGNR [152]DFTCO
CO2
NO
NO2
−2.4
−1.3
−3.1
−3.0
Zr-G [59]
Mo-G [59]
Ti-G [59]
Mn-G [59]
Fe-G [59]
Co-G [59]
DFTCOCl2−0.894
−0.960
−1.065
−1.677
−1.378
−0.828
TM-GNF [60]
TM = Sc, Ti, V, Cr, Mn, Fe, Co, N, Cu, Zn
DFTCO
CO2
[−8.13, −37.56]
[ −5.05, −16.11]
MN4-G [155]
M = Sc, Ti, V, Cr, Mn, Co, Ni, Cu, Zn
DFTCO2[−0.0032, −0.0125]
Table 9. Summary of adsorption energies Eads (in eV) for different TM-doped fullerene systems and adsorbates. The type of study is also specified.
Table 9. Summary of adsorption energies Eads (in eV) for different TM-doped fullerene systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
Ti-C60 [156]DFTH2−0.14
TM-C60 [79]
TM = Cr, Mn, Fe, Co, Ni, Cu; Zn
DFTCO
NO
[−2.94, −1.03]
[−6.52, −1.95]
Cr-C70 [158]TD-DFTNO2
SO2
[−1.55, −0.605]
[−1.178, −0.025]
Co-C70 [158]TD-DFTNO2
SO2
[−1.919, −0.806]
[−0.627, −0.0047]
Mx-C60 [159]
M = Mn, Co, Ni, Cu
(x = 2–5)
DFTCH3OH
HCOOH
CH3CH2OH
O2
CO
SO2
[−0.26, −0.17]
[−0.19, −0.08]
[−0.27, −0.16]
[−0.19, −0.10]
[−0.35, −0.12]
[−0.22, −0.09]
Con-C60 [157]
(n = 1–8)
DFTH2
2H
[1.31, 0.60]
[1.78, 1.08]
TM-C60 [67]
TM = Cu, Zn, Ni
DFTN2O
CO
[−2.30, −1.41]
[−3.5, −1.56]
Table 10. Summary of adsorption energies Eads (in eV) for different TM-doped graphdiyne systems and adsorbates. The type of study is also specified.
Table 10. Summary of adsorption energies Eads (in eV) for different TM-doped graphdiyne systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
TM-GDY [98]
TM = Li, Na, K, Rb, Cs
DFTCO2[−0.54, −0.21]
Cu-GDY [100]
Cu-B-GDY [100]
Cu-N-GDY [100]
DFTCO2[−0.5, −0.4]
[−0.45, −0.3]
[−0.5, −0.31]
Pt-GDY [103]DFTO2
CO
−1.14
−1.56
Ni-GDY [168]DFTO2
CO
O2 + CO
O
CO2
−0.69
−1.68
−1.09
−3.21
−0.08
Cu-GDY [168]DFTO2
CO
O2 + CO
O
CO2
−0.79
−1.25
−1.43
−3.21
−0.37
Sc-GDY [165]
Ti-GDY [165]
DFTHCHO−2.59
−2.24
TM-GDY [161]
TM = Ti, Sc, Li, Na, K, Ca
DFT (GGA)
DFT (vdW-DF)
DFT (DFT-D3)
8H2[−0.197, −0.10]
[−0.77, −0.194]
[−0. 345, −0.173]
Mo-GDY [163]DFTN2[−1.35, −0.93]
TM-GDY [82]
TM = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Y, Zr, Rh, Pd, Ag, La, Hf, Pt
DFTN2[−1.8, +0.12]
TM-GDY [160]
TM = Cr, Mn, Fe, Co
DFT-D3O2
CO
[−2.52, −1.21]
[−1.57, −1.25]
TM-1N-GDY [160]
TM = Cr, Mn, Fe, Co
DFT-D3O2
CO
[−2.56, −1.07]
[−1.78, −1.21]
TM-2N-GDY [160]
TM = Cr, Mn, Fe, Co
DFT-D3O2
CO
[−2.31, −1.17]
[−1.94, −1.39]
Ru-GDY [162]ExperimentalH2ONot reported
Fe-GDY [167]DFT-D3/AIMDCO2Not reported
Table 11. Summary of adsorption energies Eads (in eV) for different TM-metallic doped hybrid systems and adsorbates. The type of study is also specified.
Table 11. Summary of adsorption energies Eads (in eV) for different TM-metallic doped hybrid systems and adsorbates. The type of study is also specified.
SystemType of StudyAdsorbateEads
Ti-C30 [171]DFTCO
CO2
[−0.897, −1.673]
[−1.605, −1.247]
Ti-G-Semifullerene [172]DFTH2−1.41
Ni-, Fe-CNOs [170]ExperimentalH2Not reported
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ramirez-de-Arellano, J.M.; Canales, M.; Magaña, L.F. Carbon Nanostructures Doped with Transition Metals for Pollutant Gas Adsorption Systems. Molecules 2021, 26, 5346. https://doi.org/10.3390/molecules26175346

AMA Style

Ramirez-de-Arellano JM, Canales M, Magaña LF. Carbon Nanostructures Doped with Transition Metals for Pollutant Gas Adsorption Systems. Molecules. 2021; 26(17):5346. https://doi.org/10.3390/molecules26175346

Chicago/Turabian Style

Ramirez-de-Arellano, J. M., M. Canales, and L. F. Magaña. 2021. "Carbon Nanostructures Doped with Transition Metals for Pollutant Gas Adsorption Systems" Molecules 26, no. 17: 5346. https://doi.org/10.3390/molecules26175346

Article Metrics

Back to TopTop