Polyiodide-Doped Graphene - The Journal of Physical Chemistry C

Dec 9, 2016 - Robert A. Hoyt† , E. Marielle Remillard‡, Ekin D. Cubuk‡, Chad D. ... Emmanuel Flahaut , Marc Monthioux , Iann C. Gerber , Pascal ...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Otago Library

Article

Polyiodide-Doped Graphene Robert Alexander Hoyt, Erin Marielle Remillard, Ekin D. Cubuk, Chad David Vecitis, and Efthimios Kaxiras J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.6b11653 • Publication Date (Web): 09 Dec 2016 Downloaded from http://pubs.acs.org on December 13, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Polyiodide-Doped Graphene Robert A. Hoyt,† E. Marielle Remillard,‡ Ekin D. Cubuk,‡ Chad D. Vecitis,‡ and Efthimios Kaxiras∗,†,‡ Department of Physics, Harvard University, Cambridge, MA 02138, USA, and John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts 02138, USA E-mail: [email protected]



To whom correspondence should be addressed Harvard University Department of Physics ‡ Harvard University John A. Paulson School of Engineering and Applied Sciences †

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract Iodine-doped graphene has recently attracted significant interest as a result of its enhanced conductivity and improved catalytic activity. Using density functional theory calculations, we obtain the formation energy, desorption rate, and electronic properties for graphene systems doped with polyiodide chains consisting of one to six iodine atoms in the low-concentration limit. We find that I3 and I5 act as p-type surface dopants that shift the Fermi level 0.46 and 0.57 eV below the Dirac point, respectively. For these two molecules, molecular orbital theory and analysis of the charge density show that doping transfers electronic charge to iodine π ∗ molecular orbitals oriented perpendicular to the graphene sheet. For even-length polyiodides, we find that I6 decomposes to I2 and I4 , both of which readily desorb at 300 K. Adsorption energy calculations further show that I3 acts as an effective catalyst for the oxygen reduction reaction on graphene by stabilizing the rate-limiting OOH intermediate.

Keywords graphene, iodine, oxygen reduction, DFT, electronic structure, surface doping

Introduction Despite a wide range of extraordinary features, practical applications of graphene in modern electronics remain limited due to the difficulties of controlling its electronic properties. For example, the lack of an electronic band gap restricts application in semiconductor electronics; the presence of dangling bonds on the exposed carbon atoms at the edges of graphene make samples highly reactive and prone to reactions with impurities; 1 and the low density of states near the Fermi level contributes to high sheet resistance such that even high quality graphene samples have a sheet resistance in excess of 1 kΩ/sq. 2 In order to use graphene in advanced electronics, it is necessary to develop better control of its structural 3–5 and

2

ACS Paragon Plus Environment

Page 2 of 25

Page 3 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

electronic 6 properties. Doping provides one possible path in this direction: depending on the type of dopant, the concentration, and the synthesis method, it is possible to create band gaps ranging from 0.1 to 6.4 eV, 7–9 to increase conductivity, 10 to form ultra-thin insulators, 11 and to improve catalytic performance. 12 Conventional ion implantation techniques introduce lattice defects that disrupt graphene’s electronic band structure and ultimately degrade desirable properties, such as ballistic electron transport, that arise from the Dirac-like fermions 6 near the Fermi level. One way to introduce dopants without creating structural defects is to use adsorbed atoms or molecules. Iodine-doped graphene is of particular interest as a reversible p-type surface dopant: recent experiments have shown that iodine enhances conductivity 13,14 and catalytic response 12,15 making it ideal for use in optoelectronic devices as a transparent conducting film 7,11,16,17 and as a low-cost alternative to platinum catalysts in fuel cells. 12 Moreover, reversible iodine doping allows graphene to transition from a semi-metal to a metal and back without disrupting the sp2 hybridization in pristine graphene. 14 To date, theoretical studies suggest that iodine is adsorbed to the surface of graphene by van der Waals (vdW) interactions. 18–20 However, structures that contain a 1:1 ratio of C, H, and halogen X (X=F, Cl, Br, I) atoms, like CX and CHX, or a 2:1 ratio like C2 HX, are not likely to be stable at room temperature. 11 The main difference between experimental and theoretical studies has been the treatment of iodine. Thus far, all but one theoretical study considered isolated atoms of iodine or I2 molecules interacting with the graphene sheets, whereas Raman spectroscopy and X-ray photoelectron spectroscopy data show that iodine dopants form I3 and I5 chains. 13,14,21,22 More recent work by Tristant and coworkers 20 focused on the formation of I3 and I5 on monolayer and bilayer graphene from high concentrations of I2 , as well as their geometric and vibrational properties. These authors, using molecular dynamics simulations, found vibrational spectra in good agreement with Raman spectroscopy, and from these studies concluded that I3 and I5 are stable on graphene near 300 K. What has not yet been been systematically explored for I3 and larger

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

polyiodides is the formation energy, thermodynamic stability, and charge transfer in the lowconcentration limit of order 1 at. % or less found in experimental samples. 23–27 In particular, the doping behavior of I4 and I6 is unknown. These species could be formed from precursor I2 subunits during the thermal annealing stage but are absent in experimental samples. Determining their structure and thermodynamic stability would complement the current understanding of the polyiodide-doped graphene system. In a related, important application it is also unknown how polyiodides can serve as active sites for the oxygen reduction reaction (ORR). Recent work 28–31 demonstrated that oxygen reduction proceeds primarily through a four-electron associative pathway involving OOH adsorption on undoped and substitutionally-doped graphene. Jiao at al 28 showed that the rate-limiting step is the weak adsorption of this OOH intermediate, which prefers regions of high spin and positive charge density on graphene clusters. 30,31 This preference is related to the fact that dopants at edge sites yield higher ORR activity than those in the bulk of a cluster. ORR activity in N-doped graphene can be improved by co-doping with adsorbed molecules such as polyiodides, 27 benzoate, 32 and 1-pyrenebutyrate 32 by a similar increase of positive charge induced by N atoms. In contrast to these covalently-bonded substitutional dopants, polyiodides are weakly-bound surface adsorbates. The mechanism for stabilizing the OOH intermediate by polyiodides is ambiguous since OOH could bind directly to adsorbed polyiodides rather than to positively-charged carbon atoms in graphene.

Computational Details To explore the issues raised above we studied the electronic properties of graphene doped with adsorbed polyiodide chains and OOH using first-principles electronic structure calculations based on density functional theory (DFT). We obtain band structures and formation energies within DFT with the generalized gradient approximation exchange-correlation functional of Perdew, Burke, and Ernzerhof 33,34 and the projector-augmented wave (PAW) formalism with

4

ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

frozen core states, using the standard pseudopotentials in the VASP computational package. 35–39 We employ a 7×7 graphene supercell (containing 98 C atoms) in which polyiodides and OOH adsorbates are placed, with 24.7 ˚ A vacuum spacing between images of the graphene sheet to minimize interactions. We determine the equilibrium geometries with the conjugate A. Convergence of gradient relaxation method and a maximum force tolerance of 0.02 eV/˚ total energies to within 1 meV per atom requires a 3×3×1 Monkhorst-Pack k-point sampling and a 500 eV plane-wave cutoff. We include vdW interactions using the methods of Grimme 40 (D2) and Tkatchenko-Scheffler 41 (TS). Gas phase calculations for polyiodide and OOH molecules were performed using cubic cells with side length 36 ˚ A. For visualization of the results we use the VESTA 42 and VMD 43 graphics packages. We calculate the formation energies, Ef (N ), and desorption energies, ∆EN , for each iodine chain of size N from the expressions

Ef (N ) = Etot (N ) − EG −

N Evac (2) 2

∆EN = Evac (N ) + EG − Etot (N )

where Etot (N ) is the total energy for the polyiodide-graphene system, EG is the total energy of a pristine graphene sheet, and Evac (N ) is the energy of an isolated IN chain. Similarly, for OOH calculations we obtain the adsorption energy Eads of OOH bound to system X relative to the independent systems using the expression

Eads = Etot (OOH@X) − Etot (OOH) − Etot (X)

(1)

To determine the minimum energy configurations for polyiodides, we place single iodine atoms at several high-symmetry points above the graphene lattice. To test the effects of rotation, we rotate the linear I3 chains originating above the hollow site to form angles of 0,

π , 12

and

π 6

radians relative to a C-C bond. To test the effects of translation, we consider

a linear I3 chain originating at the top site and running parallel to a C-C bond. We allow 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

all these initial geometries to relax to their ground-state. The ground-state energies for all tested rotations and displacements differ by less than 2 meV, showing that there is no strong preference for the orientation of the iodine chains, and consequently we focus on a single orientation for the rest of the structures considered. For OOH adsorption, the minimumenergy geometries were found with a similar approach by testing multiple rotations and translations relative to undoped and I3 -doped graphene. In addition to possible energy differences arising from molecular orientation, polyiodides can exhibit a wide range of structural conformations, especially when combined with salts and metallic structures. 44 We therefore consider a range of possible geometries to determine the preferred doping structure, some of which are depicted in Fig. 1. These include: 1) linear geometries (denoted as ILN ) for all polyiodide dopants; 2) a triangular geometry for I3 ; 3) various shapes, like V, T, and dice-like for I5 , shown in Fig. 1; 4) for I6 , two IL3 units were placed 3.3 ˚ A apart both collinearly, denoted as ILL 6 , and perpendicular to each other, denoted as IT6 . In most cases, the geometry obtained after relaxation was nearly identical to the initial geometry. The only exceptions are the geometries for I3 , which always relaxed to IL3 , and the geometry for IL5 , which relaxed to IV5 (see Fig. 1(c)).

Results and Discussion Formation Energies The most stable geometries (lowest formation energy per iodine atom) are ILN for N 6= 5, and IV5 ; we denote these as simply IN in the following. All relaxed structures are coplanar and parallel to the graphene sheet, and the formation energy generally decreases with increasing chain length: I5 has the lowest formation energy (−0.376 eV/I) suggesting that it is the most stable polyiodide. The energy for adsorbed atomic I is positive since the formation energy is defined with reference to molecular iodine (I2 ), and therefore includes the full bond dissociation energy for converting I2 to atomic I. In all cases, the vdW corrections of either 6

ACS Paragon Plus Environment

Page 6 of 25

Page 7 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1: (a-c) Minimum-energy configurations for the odd-length polyiodides. (d) The next-most stable pentaiodide IT5 . (e-g) Minimum-energy configurations for each even-length polyiodide. (h) Most stable adsorption geometry for OOH on I3 -doped graphene. kind (D2 or TS) substantially decrease the formation energy for all chain lengths as seen in Fig. 2(a). To investigate the iodine concentration’s influence on the binding energy, we normalize by including multiple copies of the shorter chains (N = 1 through 4) in the same unit cell to achieve approximately 6% and 8% doping. Higher iodine concentration increases the formation energies per iodine for I1 , I3 and I4 , but does not affect the formation energy for I2 . The formation energy change is greater for the odd-length chains, with Ef (1) increasing by 0.39 eV and Ef (3) increasing by 0.06 eV, while Ef (4) increases by only 0.02 eV. The values of Ef (N ) and ∆EN , along with the structural information for each case, are summarized in Table 1. The high total formation energies, all above −2 eV, and long iodinegraphene distances indicate only weak interactions between polyiodides and graphene. From comparing formation energies and final geometries, a key result is the tendency to form weaker bonds (about 3 ˚ A) between I2 subunits, a common feature among polyiodides. 44 These appear in the I4 and I6 geometries as well as in the I2 subunit of IT5 , implying weaker bonds than those found within I3 and I5 . For the even-length chains, comparing formation

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2: (a) Formation energy, Ef (N ) per iodine atom for chain length N = 1 − 6 with and without vdW corrections. (b) Induced hole density σN versus polyiodide chain length N .

8

ACS Paragon Plus Environment

Page 8 of 25

Page 9 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

energies reveals that each of these weak I2 -I2 bonds yields approximately 0.4 eV, and is responsible for stabilizing the longer, even-length polyiodides. In contrast, the local minima for I3 and I5 are the result of favorable orbital hybridization, yielding shorter (∼ 2.9 ˚ A) bonds, and larger charge transfer from the graphene sheet as discussed below. Similar results for the geometries of the odd-length polyiodides were reported by Tristant et al.; 19 the lower binding energies and higher mean iodine-graphene distances reported here are consistent with the tendency of the GGA exchange-correlation functional to underbind structures relative to optB86b-vdW. 45 Table 1: Formation energies per iodine atom (Ef ), desorption energy (∆EN ), desorption lifetime (τN ), mean polyiodide-graphene distance (dI-G ), and nearestneighbor distances (dI-I ) for the most stable length-N polyiodide. Numbers in parentheses are from Tristant et al. 19 N 1 2 3 4 5 6

Ef (eV/I) ∆EN (eV) 0.370, (0.43) 0.884 −0.177, (−0.21) 0.354 −0.312 1.150 −0.289 0.869 −0.376 1.595 −0.323 1.352

τN (s) 1.1×102 1.4×10−7 3.3×103 6.3×101 1.0×1013 8.3×109

dI-G (˚ A) 3.60, (3.35) 3.74, (3.62) 3.70 3.73 3.68 3.76

dI-I (˚ A) 17.3, (12.30) 15.0, (12.30) 12.6 10.5 10.2 8.3

The low formation energy per elongated I-I bond in I4 and I6 suggests low reaction barriers toward decomposing into smaller polyiodide chains. To explore this in more detail, we calculated the bond dissociation curves for I4 and I6 as a function of I-I bond length by abstracting an I2 subunit from the rest of the chain. This was done by constraining the second and third iodine atoms along the chain transverse to the graphene plane while allowing all other degrees of freedom to relax, see Fig. 3 for I6 . This is an application of the distinguished (or driven) coordinate method, 46 which provides a reasonable reaction path and transition state structure when the reaction path is dominated by the selected coordinate, in this case the length of the already-elongated 3.01 ˚ A I-I bond. The resulting dissociation energies, 0.446 eV for I6 and 0.418 eV for I4 , agree well with the total energy difference between the IN precursor and the resulting IN −2 and I2 chains in separate supercells, 0.427 eV for I6 and 9

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

0.445 eV for I4 .

Figure 3: Bond dissociation curve for I6 as a function of I4 -I2 distance, with energies relative to the geometric ground state. The inset shows the dissociated geometry a distance of 7.50 ˚ A.

Electronic Structure To investigate the extent of orbital hybridization and the effect of iodine doping on graphene’s electronic structure, we show results for I2 and I3 in Fig. 4. The band structure and density of states (DOS) vary minimally with spin polarization and choice of vdW correction (we present only spin unpolarized results here). The presence of the Dirac cone, flat iodine bands, and the remarkable similarity between pristine graphene and the IN -graphene band structure for all N values considered confirms that graphene’s pz -hybridization is not disrupted by the presence of iodine atoms (the differences are minimal on the scale of Fig. 4). The Fermi level is always lower than the value of the Dirac point indicating that all polyiodide configurations considered are p-type dopants. As illustrated in Fig. 4 for I2 and I3 , the energy difference between the Dirac point and the Fermi level is generally higher for oddlength chains than even-length chains. The only exception is I6 , where the energy difference (0.48 eV) is substantially higher than that of I1 (0.11 eV). This predicts greater charge transfer for odd-length polyiodides than even-length ones, as suggested in prior work for I1 and I2 by Tristant et al. 19 To reveal the source of this difference, we project the DOS onto iodine-centered spherical harmonics, and find that the molecular orbitals of iodine are almost exclusively 5s and 5p 10

ACS Paragon Plus Environment

Page 10 of 25

Page 11 of 25

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 25

hybrids. With a large 10 eV energy difference between the 5s and 5p orbitals, there is minimal s-p hybridization; for a chain of length N this yields N 5s molecular orbitals with energies centered around −13 eV, and 3N 5p hybrid orbitals centered around −2 eV relative to the Dirac point. The shapes and characters of the 5p-hybridized orbitals for I2 and I3 within the energy range of Fig. 4(a)-(b) are shown alongside the DOS. Since all IN in their minimum-energy configurations are coplanar and approximately linear, molecular symmetry further divides the 5p hybrids into N orbitals of σ character and 2N orbitals of π character. The strong interaction between 5p orbitals oriented along the chain yields larger energy gaps between σ molecular orbitals than between π molecular orbitals, as seen in Fig. 4(a)-(b), thus explaining the difference between even and odd-length polyiodides. The odd-length neutral I3 and I5 chains have a lowest unoccupied molecular orbital (LUMO) with π ∗ character, and the small π-orbital energy gaps place these states at least 0.72 eV below the Dirac point, in good agreement with the maximum 0.8 eV Fermi level shift observed in experimental transport measurements by Wu and coworkers 14 at high iodine exposure. In contrast, the LUMO for even-length chains has σ ∗ character, and the larger energy gaps tend to place them much closer to the Dirac point. In the case of I2 and I4 this lowest unoccupied σ ∗ is very close to the Dirac point, while for I6 the increasingly-dense spacing of σ ∗ orbitals places the LUMO somewhat lower, 0.48 eV below the Dirac point. The lone anomaly is I1 , where the lack of I-I bonding destabilizes the 5p orbitals by removing exchange and correlation effects present for N ≥ 2. As a result, the nearly-degenerate 5p orbitals for I1 lie only −0.11 eV below the Dirac point. The spatial character of the transferred charge gives additional support to this molecular orbital interpretation. Fig. 4(c) and (d) show isosurfaces of the transferred charge density for I2 and I3 , defined as the difference between the charge density of the doped system and the sum of densities of the isolated, neutral polyiodide and undoped graphene sheet with the same atomic positions. The I3 and I5 chains primarily gain charge in π ∗ orbitals, while I1 and the even-length chains gain charge in σ ∗ orbitals parallel to the graphene sheet. The

12

ACS Paragon Plus Environment

Page 13 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

spatial distribution of excess charge on the iodine chains agrees well with the density of the corresponding LUMO: note, for example, the similarity for I3 between Fig. 4(d) and the corresponding π ∗ LUMO that accepts the charge in Fig. 4(b).

Charge Transfer The charge transfer between iodine and graphene offers additional insight into the role of iodine dopants in catalysis and electronic devices. We employ Bader charge analysis 47–49 to determine how much charge the polyiodide dopants acquire in their optimal configurations. We obtain the net charge transfer QN by finding the total charge QB tot contained in Bader volumes attributed to the iodine atoms, then subtracting the seven valence electrons per iodine atom to obtain the net charge. For a more direct comparison to experiment, where the iodine concentration is generally of order 1 at. %, the hole density is extrapolated to 1.0 at. % iodine doping by dividing the net charge QN by the corresponding area of graphene AG , giving a hole density σN : σN = (QB tot − 7N )/AG = QN /AG

We show the results of this calculation in Fig. 2(b). All three odd-length polyiodides draw more charge than the even-length polyiodides, generally consistent with the lower energy of the unoccupied π ∗ orbitals relative to the lowest unoccupied σ ∗ orbitals, and yield the highest hole densities induced on graphene. I1 still draws more charge than I6 , even though the the dopant band is farther below the Dirac point for I6 as mentioned above. Estimating charge transfer and hole density from the band structure alone, by integrating the DOS resulting from graphene’s linear dispersion near the Dirac point, 6,19 would therefore predict the opposite order between I6 and I1 . The Hirshfeld partial charges, 50 calculated as part of the TS vdW correction, are qualitatively consistent with the Bader charges so the latter are expected to be more reliable. We also assess the magnitude of finite size effects by calculating

13

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 25

the hole density for a pair of I3 dopants. The error bar in Fig. 2 indicates the difference: the 2×I3 structure yields a lower hole density than a single I3 . Charge transfer for larger N may be underestimated compared to calculations performed with actual 1 at. % doping. The good correspondence between charge transfer and the character of molecular orbitals explains the conductivity-enhancing effects of I3 and I5 . With the weak π hybridization in these species, their π ∗ LUMO is lower than the Dirac point, and can therefore yield significant hole densities in graphene with both I3 and I5 yielding hole densities of approximately 5 × 1012 cm−2 . This value is in reasonable agreement with a range experimental results, which demonstrate 5 × 1012 cm−2 at 4 at. % I, 25 2.6 × 1013 cm−2 at an estimated 1 to 1.7 at. % I, 17 and 4.75 × 1013 cm−2 . 14 In contrast, the even-length polyiodides have σ ∗ LUMOs that are much closer to the Dirac point due to the stronger 5p hybridization along the chain. The net charge transfer to these less-stable LUMOs is therefore reduced. The DC conductivity of graphene is proportional to its net charge density, 6 whether this density is created by an applied gate potential, impurities and adsorbates, 14,51 or even by photogenerated carriers. 52 For I3 and I5 , the significant induced hole density in graphene agrees well with the observation that polyiodide doping significantly increases the conductivity. 14 These results are also consistent with the observation that polyiodide doping improves the conductivity of carbon nanotubes. 10

Polyiodide Stability An open question is the reason for the abundance of I3 and I5 reported in the experimental literature. Although we find I5 to be the most stable per I atom, the formation energy of I3 is only 0.9 kB T lower per I atom than I4 , and is 0.4 kB T higher per I atom than I6 at 300 K. Formation energies alone cannot explain why I3 is the majority polyiodide dopant reported by experiments. An observation that may help explain this is related to the hole densities: I3 and I5 show the largest charge transfer per adsorbed molecule, suggesting that electrostatics play a role in keeping these polyiodides bound to graphene. To explore this 14

ACS Paragon Plus Environment

Page 15 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

idea further, we obtain the desorption energy ∆EN for each chain length to determine the stability of each IN species. To estimate the corresponding desorption rates kN and lifetimes −1 we use the Eyring equation: 53 kN = (kB T /h) e−∆EN /kB T e∆S/kB . The ∆SN term in the kN

full Eyring equation is assumed to be negligible since the polyiodides are only weakly bound to graphene, so that both adsorbed and free polyiodides have nearly identical vibrational modes; the only major difference is a hindered rotation into the graphene sheet for adsorbed chains. As a result, ∆SN should be negligible and is unlikely to vary with N . Table 1 shows the resulting desorption energies and lifetimes. The desorption barriers are highest for I3 , I5 and I6 , resulting in orders of magnitude longer desorption lifetimes that are in line with their larger charge transfer and dispersion interactions with graphene. Although these calculations neglect solvation and dielectric screening effects that would be present in an experimental setting, the relative rates presented here provide strong evidence that the high desorption barriers of I3 and I5 play a crucial role in their observed abundance. Combined with the small bond dissociation energies for I4 and I6 , the weak adsorption of I2 explains the absence of the even-length polyiodides studied here, and likely all longerlength polyiodides as well. Any I2 molecules remaining after the preparation of polyiodidedoped graphene should quickly desorb from the surface at 300 K given its low desorption barrier of 0.354 eV. For I4 and I6 , although the desorption barriers are significantly higher, the corresponding dissociation barriers for I2 abstraction are not. For example, the 0.445 eV dissociation barrier for I4 to form a pair of I2 molecules corresponds to a lifetime of only 4.8 × 10−6 s from the Eyring equation. Therefore I4 and I6 are expected to rapidly decompose toward I2 at 300 K. For even larger polyiodides IN with N ≥ 7, these generally feature similar elongated I-I bonds between some combination of I2 , I3 , and I5 subunits. 44 Assuming the corresponding bond dissociation energies to be similar to those for I4 and I6 , the observed lack of large polyiodides can be explained by repeated thermal decomposition toward the shorter polyiodides considered here.

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Mechanism of Catalysis In addition to reversibly doping and increasing the conductivity of graphene, polyiodides have attracted recent interest for their role in making graphene an effective catalyst for oxygen reduction. The work of Jiao et al. 28 shows that the weak binding of the OOH intermediate presents a large free energy barrier of 1.1 eV, yielding slow ORR kinetics in undoped graphene. The presence of dopants can significantly increase the binding energy of OOH and therefore reduce the corresponding barrier. Of the two observed polyiodide dopants, we focus on I3 since it has been observed to be the dominant active species in polyiodide-doped graphene. 12 To determine how strongly I3 stabilizes the OOH intermediate, we find the minimum-energy adsorption geometry for OOH for undoped graphene (G), I3 -doped graphene (I3 -G), and for both neutral I3 and charged I− 3 molecules in the gas phase. The adsorption energies, distances, and Bader charge values are shown in Table 2. On undoped graphene, OOH is only weakly adsorbed with a minimum O-graphene distance of 2.95 ˚ A, and the molecular plane is nearly parallel to graphene. Similar to polyiodides, OOH adsorption energies vary negligibly with translations and rotations. The minimum-energy adsorption geometry for I3 -doped graphene is shown in Fig. 1(h), and has a significantly stronger adsorption energy as a result of forming an O-I bond. The minimum O-graphene distance actually increases slightly in the presence of I3 , which shows that the lower adsorption energy comes from bonding to I3 rather than a carbon atom in graphene. Compared to substitutional B-doped graphene 28 and N-doped graphene, 28,30 I3 binds OOH by about −0.13 eV more and should therefore correspond to a smaller free energy barrier of 0.57 eV. This significant lowering of the free energy barrier for OOH adsorption, due to direct bonding to I3 dopants, explains how I3 can serve as the active site for OOH adsorption and hence promote the early stages of oxygen reduction. There is also an interesting relationship between OOH adsorption and the corresponding charge transfer, where the stronger binding of OOH on I3 -doped graphene is accompanied by a substantial increase in its partial charge. For reference, we compare the OOH adsorption 16

ACS Paragon Plus Environment

Page 16 of 25

Page 17 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 2: Adsorption energy (Eads ), minimum O-graphene distance (dO−G ), O-I bond length (dO−I ), and net OOH Bader charge (QOOH ) for each system interacting with OOH. system G I3 -G I3 gas I− 3 gas

Eads (eV) −0.441 −0.968 −1.440 −0.569

dO−G ˚ A dO−I ˚ A 2.95 3.06 2.23 2.12 2.33

QOOH (e) 0.26 0.35 0.33 0.44

strength to neutral I3 and the corresponding I− 3 anion in the gas phase. I3 yields significantly stronger OOH adsorption than I− 3 , see Table 2. I3 on graphene is roughly halfway between neutral and anionic with a net Bader charge of 0.55 e, and its corresponding OOH adsorption energy and O-I bond length are between those of neutral I3 and anionic I− 3 in the gas phase. This close agreement reflects the weak I3 -graphene interaction, and suggests that the reactivity of polyiodide-doped graphene could be improved by decreasing the net charge on the polyiodide dopants, for example by co-doping graphene with an electronegative element. This may explain the recent observation 27 that co-doping graphene with both polyiodides and N yields faster ORR kinetics than either dopant alone, despite lower I and N loading than the individual samples.

Conclusion In summary, we studied the electronic properties of iodine-doped graphene to understand the possibility of using polyiodides as dopants (I1 through I6 ). To foster comparisons between these different polyiodides, we calculated the binding energies, electronic band structures, densities of states, charge transfer, and desorption rates for each polyiodide on graphene. We find that, in general, longer chain lengths and lower concentrations have more favorable formation energies, with little dependence on the relative orientation of the polyiodide molecule and the graphene lattice. Band structure calculations and molecular orbital theory reveal that polyiodide molecules act as p-type dopants and do not noticeably disrupt graphene’s

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

band structure. There are also clear distinctions in doping behavior among the different chain lengths as a result of orbital hybridization: the I3 and I5 chains acquire charge in π ∗ orbitals that are oriented with lobes perpendicular to graphene, while I1 and even-length chains acquire charge in orbitals with lobes parallel to graphene. As a result of weak π hybridization and thus lower-energy LUMOs, the odd-length chains exhibit 25% greater charge transfer per iodine atom. This increased charge helps keep the experimentally-observed I3 and I5 species stably bound to graphene, while I2 rapidly desorbs at 300 K. Furthermore, although I4 and I6 are reasonably stable with respect to desorption, they are unstable with respect to dissociation into separate I2 molecules as a result of their elongated I-I bonds and low bond dissociation energies. This explains the observed lack of longer polyiodides IN (N ≥ 7), which generally feature longer bonds between shorter polyiodide subunits. We also find that the rate-limiting OOH intermediate is effectively stabilized by adsorbed I3 chains, resulting in a lower free energy barrier that agrees with the fast ORR kinetics observed for polyiodide-doped graphene. Note The authors declare no competing financial interests.

Acknowledgement The authors thank Dr. Geoffrey Beach, Matthew Montemore, and Thomas Markovich for technical assistance, and Dmitry Vinichenko and Sarah Schlotter for useful discussions. Calculations were performed on the Odyssey cluster, supported by the FAS Division of Science, Research Computing Group at Harvard University. This work was supported by ARO MURI Award W911NF-14-0247 and Harvard-UTEC Faculty Grants.

References (1) Jiang, D.; Sumpter, B. G.; Dai, S. Unique Chemical Reactivity of a Graphene Nanoribbon’s Zigzag Edge. J. Chem. Phys. 2007, 126, 134701. 18

ACS Paragon Plus Environment

Page 18 of 25

Page 19 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(2) Kholmanov, I. N.; Magnuson, C. W.; Aliev, A. E.; Li, H.; Zhang, B.; Suk, J. W.; Zhang, L. L.; Peng, E.; Mousavi, S. H.; Khanikaev, A. B. et al. Improved Electrical Conductivity of Graphene Films Integrated with Metal Nanowires. Nano Lett. 2012, 12, 5679–5683. (3) Wang, W. L.; Santos, E. J. G.; Jiang, B.; Cubuk, E. D.; Ophus, C.; Centeno, A.; Pesquera, A.; Zurutuza, A.; Ciston, J.; Westervelt, R. et al. Direct Observation of a Long-Lived Single-Atom Catalyst Chiseling Atomic Structures in Graphene. Nano Lett. 2014, 14, 450–455. (4) Li, X.; Cai, W.; An, J.; Kim, S.; Nah, J.; Yang, D.; Piner, R.; Velamakanni, A.; Jung, I.; Tutuc, E. et al. Large-Area Synthesis of High-Quality and Uniform Graphene Films on Copper Foils. Science 2009, 324, 1312–1314. (5) Chen, W.; Cui, P.; Zhu, W.; Kaxiras, E.; Gao, Y.; Zhang, Z. Atomistic Mechanisms for Bilayer Growth of Graphene on Metal Substrates. Phys. Rev. B 2015, 91, 045408. (6) Castro Neto, a. H.; Peres, N. M. R.; Novoselov, K. S.; Geim, a. K. The Electronic Properties of Graphene. Rev. Mod. Phys. 2009, 81, 109–162. (7) Guo, B.; Fang, L.; Zhang, B.; Gong, J. R. Graphene Doping: A Review. Insci. J. 2011, 1, 80–89. (8) Rani, P.; Jindal, V. K. Designing Band Gap of Graphene by B and N Dopant Atoms. RSC Adv. 2013, 3, 802. (9) Klintenberg, M.; Leb`egue, S.; Katsnelson, M. I.; Eriksson, O. Theoretical Analysis of the Chemical Bonding and Electronic Structure of Graphene Interacting with Group IA and Group VIIA Elements. Phys. Rev. B 2010, 81, 085433. (10) Zhao, Y.; Wei, J.; Vajtai, R.; Ajayan, P. M.; Barrera, E. V. Iodine Doped Carbon

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

Nanotube Cables Exceeding Specific Electrical Conductivity of Metals. Sci. Rep. 2011, 1, 83. (11) Karlick´ y, F.; Kumara Ramanatha Datta, K.; Otyepka, M.; Zboˇril, R. Halogenated Graphenes: Rapidly Growing Family of Graphene Derivatives. ACS Nano 2013, 7, 6434–6464. (12) Yao, Z.; Nie, H.; Yang, Z.; Zhou, X.; Liu, Z.; Huang, S. Catalyst-Free Synthesis of Iodine-Doped Graphene via a Facile Thermal Annealing Process and its Use for Electrocatalytic Oxygen Reduction in an Alkaline Medium. Chem. Commun. 2012, 48, 1027–9. ˇ (13) Poh, H. L.; Simek, P.; Sofer, Z.; Pumera, M. Halogenation of Graphene With Chlorine, Bromine, or Iodine by Exfoliation in a Halogen Atmosphere. Chem. - Eur. J. 2013, 19, 2655–62. (14) Wu, Z.; Han, Y.; Huang, R.; Chen, X.; Guo, Y.; He, Y.; Li, W.; Cai, Y.; Wang, N. Semimetallic-to-Metallic Transition and Mobility Enhancement Enabled by Reversible Iodine Doping of Graphene. Nanoscale 2014, 6, 13196–13202. (15) Baek, J. Y.; Jeon, I.-Y.; Baek, J.-B. Edge-Iodine/Sulfonic Acid-Functionalized Graphene Nanoplatelets as Efficient Electrocatalysts for Oxygen Reduction Reaction. J. Mater. Chem. A 2014, 2, 8690–8695. (16) Pei, S.; Zhao, J.; Du, J.; Ren, W.; Cheng, H.-M. Direct Reduction of Graphene Oxide Films into Highly Conductive and Flexible Graphene Films by Hydrohalic Acids. Carbon 2010, 48, 4466 – 4474. (17) Jung, N.; Crowther, A. C.; Kim, N.; Kim, P.; Brus, L. Raman Enhancement on Graphene: Adsorbed and Intercalated Molecular Species. ACS Nano 2010, 4, 7005– 7013.

20

ACS Paragon Plus Environment

Page 21 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(18) Medeiros, P. V. C.; Mascarenhas, A. J. S.; de Brito Mota, F.; de Castilho, C. M. C. A DFT study of Halogen Atoms Adsorbed on Graphene Layers. Nanotechnology 2010, 21, 485701. (19) Tristant, D.; Puech, P.; Gerber, I. C. Theoretical Study of Graphene Doping Mechanism by Iodine Molecules. J. Phys. Chem. C 2015, 119, 12071–12078. (20) Tristant, D.; Puech, P.; Gerber, I. C. Theoretical Study of Polyiodide Formation and Stability on Monolayer and Bilayer Graphene. Phys. Chem. Chem. Phys. 2015, 17, 30045–30051. (21) Jung, N.; Kim, N.; Jockusch, S.; Turro, N. J.; Kim, P.; Brus, L. Charge Transfer Chemical Doping of Few Layer Graphenes: Charge Distribution and Band Gap Formation. Nano Lett. 2009, 9, 4133–7. (22) Kalita, G.; Wakita, K.; Takahashi, M.; Umeno, M. Iodine Doping in Solid PrecursorBased CVD Growth Graphene Film. J. Mater. Chem. 2011, 21, 15209. (23) Lee, C.-H.; Kang, G.-S.; Lee, Y.-K.; Lee, S.; Jo, S. M.; Yoo, S. J.; Jang, J. H.; Lee, D. C.; Joh, H.-I. Synthesis and Properties of Nitrogen and Iodine Co-Functionalized Graphene Oxide and Its Electrochemical Applications. Sci. Adv. Mater. 2016, 8, 28–33. ˇ (24) Simek, P.; Kl´ımov´a, K.; Sedmidubsk´ y, D.; Jankovsk´ y, O.; Pumera, M.; Sofer, Z. Towards Graphene Iodide: Iodination of Graphite Oxide. Nanoscale 2015, 7, 261–270. (25) Kim, H.; Renault, O.; Tyurnina, A.; Simonator, J.-P.; Rouchon, D.; Mariolle, D.; Chevalier, N.; Dijon, J. Doping Efficiency of Single and Randomly Stacked Bilayer Grpahene by Iodine Adsorption. App. Phys. Lett. 2014, 105, 011605. (26) Kim, H.; Renault, O.; Tyurnina, A.; Guillet, J.-F.; Simonato, J.-P.; Rouchon, D.; Mariolle, D.; Chevalier, N.; Dijon, J. Doping Characteristics of Iodine on As-Grown Chemical Vapor Deposited Graphene on Pt. Ultramicroscopy 2015, 159, 470 – 475. 21

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 25

(27) Zhan, Y.; Huang, J.; Lin, Z.; Yu, X.; Zeng, D.; Zhang, X.; Xie, F.; Zhang, W.; Chen, J.; Meng, H. Iodine/Nitrogen Co-Doped Graphene as Metal Free Catalyst for Oxygen Reduction Reaction. Carbon 2015, 95, 930–939. (28) Jiao, Y.; Zheng, Y.; Jaroniec, M.; Qiao, S. Z. Origin of the Electrocatalytic Oxygen Reduction Activity of Graphene-Based Catalysts: A Roadmap to Achieve the Best Performance. J. Am. Chem. Soc. 2014, 136, 4394–4403. (29) Yang, L.; Jiang, S.; Zhao, Y.; Zhu, L.; Chen, S.; Wang, X.; Wu, Q.; Ma, J.; Ma, Y.; Hu, Z. Boron-Doped Carbon Nanotubes as Metal-Free Electrocatalysts for the Oxygen Reduction Reaction. Angew. Chem., Int. Ed. 2011, 50, 7132–7135. (30) Zhang, L.; Xia, Z. Mechanisms of Oxygen Reduction Reaction on Nitrogen-Doped Graphene for Fuel Cells. J. Phys. Chem. C. 2011, 115, 11170–11176. (31) Zhang, L.; Niu, J.; Li, M.; Xia, Z. Catalytic Mechanisms of Sulfur-Doped Graphene as Efficient Oxygen Reduction Reaction Catalysts for Fuel Cells. J. Phys. Chem. C. 2014, 118, 3545–3553. (32) Mohapatra, B. D.; Mantry, S. P.; Behera, N.; Behera, B.; Rath, S.; Varadwaj, K. S. K. Stimulation of Electrocatalytic Oxygen Reduction Activity on Nitrogen Doped Graphene Through Noncovalent Molecular Functionalisation. Chem. Commun. 2016, 52, 10385–10388. (33) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. (34) Perdew, J. P.; Burke, K.; Ernzerhof, M. Erratum: Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1997, 78, 1396. (35) G. Kresse and J. Hafner, Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558. 22

ACS Paragon Plus Environment

Page 23 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(36) G. Kresse and J. Hafner, Ab Initio Molecular-Dynamics Simulation of the Liquid-MetalAmorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994, 49, 14251. (37) G. Kresse and J. Furthm¨ uller, Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mat. Sci. 1996, 6, 15. (38) G. Kresse and J. Furthm¨ uller, Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169. (39) Hafner, J. Ab-Initio Simulations of Materials Using VASP: Density-Functional Theory and Beyond. J. Comp. Chem. 2008, 29, 2044–2078. (40) Grimme, S. Semiempirical GGA-Type Density Functional Constructed With a LongRange Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. (41) Tkatchenko, A.; Scheffler, M. Accurate Molecular van der Waals Interactions from Ground-State Electron Density and Free-Atom Reference Data. Phys. Rev. Lett. 2009, 102, 073005. (42) Momma, K.; Izumi, F. VESTA3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 1272–1276. (43) Humphrey, W.; Dalke, A.; Schulten, K. VMD – Visual Molecular Dynamics. J. Mol. Graphics 1996, 14, 33–38. (44) Svensson, P. H.; Kloo, L. Synthesis, Structure, and Bonding in Polyiodide and Metal Iodide-Iodine Systems. Chem. Rev. 2003, 103, 1649–1684. (45) Klimeˇs, J. c. v.; Bowler, D. R.; Michaelides, A. van der Waals Density Functionals Applied to Solids. Phys. Rev. B 2011, 83, 195131.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(46) Williams, I.; Maggiora, G. Use and Abuse of the Distinguished-Coordinate Method for Transition-State Structure Searching. J. Mol. Struct.: THEOCHEM 1982, 89, 365– 378. (47) Henkelman, G.; Arnaldsson, A.; J´onsson, H. A Fast and Robust Algorithm for Bader Decomposition of Charge Density. Comput. Mater. Sci. 2006, 36, 354 – 360. (48) Sanville, E.; Kenny, S. D.; Smith, R.; Henkelman, G. Improved Grid-Based Algorithm for Bader Charge Allocation. J. Comput. Chem. 2007, 28, 899–908. (49) Tang, W.; Sanville, E.; Henkelman, G. A Grid-Based Bader Analysis Algorithm Without Lattice Bias. J. Phys.: Condens. Matter 2009, 21, 084204. (50) Hirshfeld, F. L. Bonded-Atom Fragments for Describing Molecular Charge Densities. Theor. Chim. Acta 1977, 44, 129–138. (51) Jo, G.; Choe, M.; Lee, S.; Park, W.; Kahng, Y. H.; Lee, T. The Application of Graphene as Electrodes in Electrical and Optical Devices. Nanotechnology 2012, 23, 112001. (52) Kim, S. J.; Song, W.; Kim, S.; Kang, M.-A.; Myung, S.; Sook Lee, S.; Lim, J.; An, K.S. Tunable Functionalization of Graphene Nanosheets for Graphene-Organic Hybrid Photodetectors. Nanotechnology 2016, 27, 075709. (53) Eyring, H. The Activated Complex in Chemical Reactions. J. Chem. Phys. 1935, 3, 107–115.

24

ACS Paragon Plus Environment

Page 24 of 25

Page 25 of 25

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment