Accepted Manuscript Coexistence of Dirac cones and Kagome flat bands in a porous graphene Mina Maruyama, Nguyen Thanh Cuong, Susumu Okada PII:
S0008-6223(16)30745-X
DOI:
10.1016/j.carbon.2016.08.090
Reference:
CARBON 11276
To appear in:
Carbon
Received Date: 10 June 2016 Revised Date:
5 August 2016
Accepted Date: 29 August 2016
Please cite this article as: M. Maruyama, N.T. Cuong, S. Okada, Coexistence of Dirac cones and Kagome flat bands in a porous graphene, Carbon (2016), doi: 10.1016/j.carbon.2016.08.090. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
RI PT
Coexistence of Dirac cones and Kagome flat bands in a porous graphene Mina Maruyamaa,∗, Nguyen Thanh Cuongb , Susumu Okadaa a Graduate
SC
School of Pure and Applied Sciences, University of Tsukuba, 1-1-1 Tennodai, Tsukuba, Ibaraki 305-8571, Japan b International Center for Young Scientist (ICYS) and International Center for Materials Nanoarchitectonics (MANA), National Institute for Materials Science (NIMS), Tsukuba, Ibaraki 305-0044, Japan
M AN U
Abstract
We investigate the geometric and electronic structures of porous graphene networks consisting of phenalenyl and phenyl groups, which are connected alternately with C3 symmetry via single bonds to form a honeycomb network with an internal degree of freedom. The networks possess both Dirac cones and Kagome flat bands near the Fermi level (EF ) because the phenalenyl and phenyl form hexagonal and Kagome lattices, respectively. Because of the large spacing be-
TE D
tween phenalenyl units, the networks possess very slow massless electrons/holes at EF , the Fermi velocity of which is a hundred times lower than that of graphene, leading to spin polarization on the networks with antiferromagnetic (AF) and ferromagnetic (F) ordering as their stable states. Our calculations
EP
show the AF state is the ground state whose energy is lower by 14 meV than that of the F state. We also demonstrate that the electronic structure of the 2D networks is sensitive to the rotation of the phenyl unit connecting phenalenyl
AC C
units by changing the π electron network.
1. Introduction Electronic structures of nanoscale and low-dimensional materials consist-
ing of three-fold-coordinated (sp2 ) carbon are sensitive to their dimensionality, ∗ Corresponding
author Email address:
[email protected] (Mina Maruyama)
Preprint submitted to Carbon
August 3, 2016
ACCEPTED MANUSCRIPT
topology, and boundary condition. In contrast to graphene, which is a zerogap semiconductor, carbon nanotubes are either metallic or semiconducting
RI PT
5
depending on the periodic boundary condition along the circumference of the two-dimensional (2D) electronic energy band of graphene [1, 2, 3, 4]. An open
boundary condition results in a one-dimensional (1D) nanoribbon of graphene, leading to further variation of electronic structures [5, 6, 7, 8, 9, 10, 11, 12, 13].
The nanoribbons with armchair edges are metallic or semiconducting, and their
SC
10
band gap is inversely proportional to their width. In contrast, ribbons with zigzag or near-zigzag edges possess peculiar edge-localized states, namely edge
M AN U
states, with non-bonding nature of π electron states, causing magnetic ordering around the edge atomic sites. In addition to 1D boundary conditions, pores and 15
topological defects in hexagonal graphitic networks cause unusual flat dispersion bands at or near the Fermi level (EF ), leading to magnetic ordering throughout the networks in some cases [14, 15, 16, 17, 18].
Polymerization and oligomerization of hydrocarbon molecules can provide structurally well-defined π electron networks with various dimensionalities. These low-dimensional or porous graphene networks have attracted much attention
TE D
20
because of their high surface areas and electronic structure modulation, which are promising for application in energy and electronic devices in the near future [19, 20, 21, 22]. The electronic structures of such polymers and oligomers
25
EP
strongly depend on the combination of hydrocarbon molecules and polymer chains, which allows us to tailor their physical and chemical properties by fabrication under optimum external conditions [23, 24, 25, 26]. This indicates that
AC C
porous magnetic carbon nanomaterials may be synthesized by polymerizing hydrocarbon molecules with radical spins. Phenalenyl is one possible candidate as a constituent unit of such magnetic materials [27, 28, 29, 30, 31]. Dimerization
30
of triphenylphenalenyl and cyclodehydrogenated phenalenyl leads to complexes containing two phenalenyls possessing singlet and triplet spin coupling as their stable states [32, 33, 34, 35]. Despite intensive experimental and theoretical work on two phenalenyls connected via a covalent π network, the possible electronic structures of 2D networks consisting of phenalenyl molecules have not 2
ACCEPTED MANUSCRIPT
35
yet been investigated in detail. The three-fold symmetry of the phenalenyl net-
RI PT
work causes honeycomb networks of radical spins or non-bonding state to be distributed on phenalenyls with appropriate interconnect units, leading to in-
teresting electron systems that are characterized by competition between slow massless electrons and radical spin at EF . 40
In this work, we aim to theoretically investigate the possible structures of
SC
2D porous hydrocarbon networks of sp2 C obtained by assembling phenalenyl molecules in a hexagonal manner with phenyl interconnects and their electronic
and magnetic properties using density functional theory. Our calculation shows
45
M AN U
that the sheet exhibits unusual electronic structure near EF : Radical spin states on phenalenyl form a Dirac cone at EF . The Fermi velocity of the Dirac cone is one hundred times lower than that of graphene because of the hexagonally arranged phenalenyl units with large interunit spacing. Simultaneously, flat dispersion bands emerge just below and above EF arising from the highest occupied (HO) and lowest unoccupied (LU) states of the phenyl units, respectively, which 50
form a Kagome lattice [36, 37]. Slow massless electrons at EF lead to magnetic
TE D
ordering on the porous hydrocarbon sheet with antiferromagnetic (AF) and ferromagnetic (F) states even though it possesses zero density of states at EF . The spin exchange interaction between adjacent radical spins on phenalenyl is 10 meV per pair, preferring the singlet coupling. The electronic structure and spin state of the porous hydrocarbon network can be tuned by rotational angle
EP
55
of the phenyl connecting adjacent phenalenyls via single bonds.
AC C
2. Calculation method
All calculations were performed within the framework of density functional
theory [38, 39] using the simulation tool of atom technology (STATE) code [40].
60
To calculate the exchange-correlation energy among interacting electrons, we used the generalized gradient approximation with the Perdew-Burke-Ernzerhof functional [41]. To investigate the spin-polarized states of the porous hydrocarbon sheets, the spin degree of freedom was taken into account for all calculations.
3
ACCEPTED MANUSCRIPT
Vanderbilt ultrasoft pseudopotentials were used to describe the electron-ion interaction [42]. The valence wave functions and charge density were expanded
RI PT
65
in terms of the plane waves with cutoff energies of 25 and 225 Ry, respectively, which sufficiently describe the electronic structure and energetics of hydrocarbon molecules and graphene-related materials [43]. To simulate an isolated porous
hydrocarbon sheet, the sheet was separated by a vacuum spacing of 10.58 ˚ A. 70
Atomic structures of the sheets were optimized until the force acting on each
SC
atom was less than 1.33×10−3 HR/au. Integration over the Brillouin zone was carried out using an equidistance mesh of 2×2×1 k points. To inject holes into
M AN U
the sheet, we used the effective screening medium (ESM) method to solve the Poisson equation under the field-effect-transistor structure with a counter gate 75
electrode situated above the sheet with vacuum regions of 5.29 ˚ A [44].
3. Results and discussion
Figure 1 shows the optimized structures of 2D porous hydrocarbon net-
TE D
works consisting of phenalenyl connected via phenyl with various arrangements under the hexagonal cell parameter of 19.4 ˚ A. For all phenyl arrangements, 80
the phenalenyl units retained a planar structure as their stable conformation. As illustrated in Fig. 1a, two phenalenyl units per unit cell form a hexagonal lattice in which phenyl units connect adjacent phenalenyl units through cova-
EP
lent bonds like those of graphene. Simultaneously, three phenyl units form a Kagome lattice in which phenalenyl acts as an interunit bond. Because of the 85
planar conformations of both phenyl and phenalenyl, the porous hydrocarbon
AC C
sheet possesses a 2D π electron system. Rotation of phenyl modulates the π electron system: By rotating one of three phenyl units, the π electrons form a 1D network in which the alternating phenalenyl and phenyl units form zigzag chains (1D structure in Fig. 1b). By rotating two phenyl units, π electrons are
90
segmented into a phenalenyl dimer containing a phenyl unit (dimer structure in Fig. 1c). Finally, π electrons are localized on each phenalenyl and phenyl in the monomer structure depicted in Fig. 1d.
4
ACCEPTED MANUSCRIPT
The total energies of the porous hydrocarbon sheets with various phenyl con-
95
RI PT
formations are listed in Table 1. Among the four structures, the planar structure is the least stable with a total energy higher by 0.8 eV per cell than that
of the ground state conformation. The total energy monotonically decreases by increasing the number of rotated phenyl units, although the π conjugation de-
creases. The decrease of the total energy with respect to the rotation of phenyl
100
SC
units is caused by the steric hindrance between hydrogen atoms attached to the
phenyl and phenalenyl. Since the hydrogen atoms attached to the phenyl and phenalenyl are positively charged, the hydrogen atoms tend to separate each
M AN U
other to reduce the Coulomb’s repulsive interaction. The dynamical stability of the porous hydrocarbon sheets was investigated by ab initio molecular dynamics simulations conducted at a constant temperature up to 3000 K for simulation 105
times of 1 ps. Under all temperatures, all structures retained their initial network topologies. On the other hand, at the elevated temperatures, the phenyl units are tiled by angles of 20-40 degrees from the planar arrangement owing to the steric hindrance between hydrogen atoms attached to the phenalenyl and
110
TE D
phenyl. The sheet hardly retains the planar conformation under the finite temperature. Thus, we confirm that the 2D networks of phenalenyl and phenyl proposed here are both statically and dynamically stable and are expected to be stable under ambient conditions once they are synthesized experimentally.
EP
Fig. 2 presents the electronic structures of 2D porous hydrocarbon sheets consisting of phenalenyl and phenyl. The planar structure possesses a char115
acteristic feature around EF (Fig. 2a). The porous hydrocarbon sheet is a
AC C
zero-gap semiconductor with a pair of linear dispersion bands at EF . In addition to the linear dispersion bands, we find three bunched states just above and below the linear dispersion bands. One of the three branches exhibits perfect flat band nature while the remaining two have finite band dispersion, consistent
120
with their Kagome band structure. These observations indicate that Dirac cone and Kagome flat bands coexist in the porous hydrocarbon sheet of phenalenyl and phenyl with flat conformation. The Dirac cone is robust while the Kagome bands are fragile against the rotation of phenyl: a Dirac cone is observed for all 5
ACCEPTED MANUSCRIPT
porous hydrocarbon sheets, regardless of the phenyl conformation. In contrast, the Kagome flat bands are absent for the sheets with 1D and dimer structures,
RI PT
125
because the electron system of the phenyl unit does not retain the symmetry of the Kagome lattice.
We evaluated the tight-binding parameters of the Dirac cone and Kagome band in the different sheet conformations from their bandwidth. The calculated
transfer integrals of the Dirac cone tD are 17, 16, 14, and 10 meV for the
SC
130
planar, 1D, dimer, and monomer conformations, respectively. Because of the narrow bandwidth, the Fermi velocities of the Dirac cone are 1.2×104 , 1.4×104 ,
M AN U
0.5×104 , 1.6×104 m/s for the planar, 1D, dimer, and monomer conformations, respectively. Therefore, the porous hydrocarbon sheets are possible fields for 135
strongly correlated electron systems that induce peculiar physical phenomena. A low Fermi velocity may cause large Fermi level instability even though the sheets possess zero density of states at EF . For the Kagome bands, we also evaluated the transfer integral tK . The calculated tK are 27 and 21 meV for the planar and monomer conformations, respectively.
To clarify the physical origin of the Dirac cone in the planar system, we
TE D
140
investigated the squared wave functions of the states at the K point (Fig. 3a). The wave functions are localized on phenalenyl units even though this porous hydrocarbon sheet possesses a 2D planar π electron network. By focusing on
145
EP
the wave function distribution within each phenalenyl, the state exhibits nonbonding nature in each phenalenyl unit. This indicates that the electron states associated with the Dirac cone can be ascribed to the singly occupied states of
AC C
isolated phenalenyl monomers. In addition to the non-bonding nature within each phenalenyl unit, the wave function also exhibits non-bonding nature between adjacent phenalenyl units. Thus, the pair of linear dispersion bands at
150
EF is ascribed to the hexagonally arranged singly occupied state or radical spin state of phenalenyl units. Similar non-bonding nature within and between phenalenyl units was also found for the porous hydrocarbon sheets with 1D, dimer, and monomer conformations (Figs. 3b-3d). Next, we examined the wave function associated with the Kagome band. 6
ACCEPTED MANUSCRIPT
155
Figure 4 shows the isosurfaces of the squared wave functions of Kagome bands
RI PT
just below EF at the Γ point. For the planar conformation, the wave functions are mainly distributed on the phenyl unit with the character of the HO states
of phenyl molecules and are extended throughout the networks. Therefore, the flat dispersion band is ascribed to the delicate balance of the electron transfer 160
among the HO state of the three phenyl units of the Kagome network, as in the
SC
case of the usual Kagome flat band state. In contrast, for the 1D and dimer
conformations that lack the Kagome flat band, the states below EF exhibit anisotropic nature, because of the orthogonality between the π states of the
165
M AN U
planar network and rotated phenyl units (Figs. 4b and 4c). For the monomer conformation, which possesses a Kagome flat band, the wave functions are more localized at the phenyl unit with the HO state nature and weakly extended throughout the network (Fig. 4d). This weak extended nature leads to the smaller tK of the monomer conformation than that of the planar conformation. The narrow band width of the Dirac cone of the porous hydrocarbon sheets 170
results in a large density of states near EF , which induces the Fermi level insta-
TE D
bility, causing magnetic ordering on the sheet, even though the sheets possess zero density of states at EF because of the pair of linear dispersion bands at the K point. Figure 5 illustrates the spin density of the porous hydrocarbon sheets
175
EP
ρs (r) evaluated by the equation
ρs (r) = ρα (r) − ρβ (r)
where ρα (r) and ρβ (r) are the charge densities of α and β spin components,
AC C
respectively. We found that all the porous hydrocarbon sheets show spin polarization. Polarized electrons are localized on the phenalenyl units; the α and β spins are distributed on one of two sublattices of the phenalenyl units, as in the case of the spin density of an isolated phenalenyl. Because of the spin
180
density of the phenalenyl unit, each phenalenyl unit possesses S = 1/2 radical spin corresponding to the difference between the numbers of two sublattices. Furthermore, the S = 1/2 spin of the phenalenyl unit exhibits AF and F spin ordering throughout the sheets. 7
ACCEPTED MANUSCRIPT
The total energies of the porous hydrocarbon sheets with AF and F states are listed in Table 1. The total energy of each sheet with spin polarization is
RI PT
185
about 300 meV lower than that of the corresponding sheet with non-magnetic
states for all phenyl conformations. Furthermore, the total energy of the AF state is lower than that of the F state by 14, 9, and 4 meV for the planar, 1D,
and dimer conformations, respectively. These values indicate that the S = 1/2 spin on the phenalenyl unit interacts with its neighboring sites with weak sin-
SC
190
glet spin interaction J. The calculated spin interaction J per phenalenyl pair is about 10 meV. Note that the AF and F states are degenerate for the sheet with
M AN U
monomer conformation, because the total energies of theses states are the same within numerical error. Thus, this sheet does not exhibit long-range magnetic 195
ordering. In addition, the rotation of the phenyl unit may turn off the spin interaction between adjacent phenalenyl units because of the monotonic decrease of the energy difference between the AF and F states. Therefore, the magnetic states of the porous hydrocarbon sheets can be tuned by controlling the orientation of phenyl units with respect to the phenalenyl units: the planar, 1D, and dimer conformations are a S = 1/2 AF sheet, magnetically independent array
TE D
200
of S = 1/2 AF chains, and magnetically isolated singlet-spin dimer network, respectively.
Figure 6 shows the electronic energy bands of the AF and F states of the
205
EP
porous hydrocarbon sheets with various phenyl conformations. In all cases, the spin polarization modulates the pair of linear dispersion bands at EF . For the AF state, flat dispersion bands emerge above and below EF instead of a Dirac
AC C
cone because of the antiparallel coupling between two S = 1/2 radical spins of phenalenyl units. The energy gap between the upper and lower branches of the states associated with the spin polarization is 0.69-0.67 eV. For the F state,
210
the Dirac cone is split into α and β spin states with a spin exchange energy of 0.65 eV, almost retaining its characteristic shape. In addition to the Dirac cone, electronic states near EF are also modulated by the spin polarization. For instance, the Kagome band is split into α and β spin states with an energy difference of 0.16 eV. 8
ACCEPTED MANUSCRIPT
215
The unusual electron systems near EF of the porous hydrocarbon sheets may
RI PT
exhibit further interesting physical properties upon hole injection into the Dirac cone and Kagome band. Because of the localized and extended natures of the
electronic states associated with the Dirac cone and Kagome band, respectively,
hole injection changes spin polarization in the porous hydrocarbon sheet. By 220
injecting holes into the porous hydrocarbon sheet with flat conformation by a
SC
counter planar electrode, we find that the number of polarized electron spins
strongly depends on the number of holes injected (Table 2). The spin density in the porous hydrocarbon sheet also strongly depends on the hole concentration
225
M AN U
(Fig. 7). For a low hole concentration corresponding to hole injection into the Dirac cone, the polarized electron spin exhibits qualitatively the same distribution as that of the undoped system. For a high hole concentration at which EF crosses the Kagome band, the spin density exhibits different characteristics to that of the undoped case: The polarized electron spin is distributed on not only the phenalenyl but also the phenyl units, reflecting the extended nature of the 230
wave functions associated with the Kagome band. Furthermore, in contrast to
TE D
the non-doped network, the porous hydrocarbon sheet with a high hole concentration exhibits F spin ordering in its ground state. It should be noted that the carrier density to induce the magnetism associated with the Kagome flat band is about 0.8 × 1014 cm−2 . Thus, the magnetic state of the porous hydrocarbon 235
sheet is tunable using ionic gating by the formation of electrical double layers
EP
(EDLs) which can inject carrier density up to 8 × 1014 cm−2 [45].
AC C
4. Discusion
Since the phenyl prefers the tiled/rotated conformation in their ground state
owing to the steric hindrance between hydrogen atoms attached to phenalenyl
240
and phenyl, we further investigate the detailed energetics and spin interaction of a constituent unit of the porous hydrocarbon network. Here, we consider a phenalenyl dimer connected via phenyl as for the structural model of the the porous hydrocarbon network (Fig. 8a). Figure 8b show the total energy of the
9
ACCEPTED MANUSCRIPT
model molecule as a function of the rotational angle of phenyl with respect to the phenalenyl moieties. By rotating the phenyl, the total energy of the molecule
RI PT
245
monotonically decreases and possesses a minimum at the angle of about 35 degree. The total energy of the molecule with the rotation angle of 35 degree
is lower by about 120 meV than that of the flat arrangement. Then, by further increase the angle, the total energy monotonically increase. Therefore, the
phenyl connecting the phenalenyl units of the free-standing porous hydrocarbon
SC
250
sheet may be rotated by about 35 degree under the ambient condition as their ground state conformation. On the other hand, the porous hydrocarbon sheet
M AN U
may retain its flat conformation by adsorbing on appropriate substrates, such as graphene and h-BN, since the energy gain upon the interaction between the 255
sheet and substrates compensate the energy loss upon the rotation of phenyl. In addition to the energetics, it is worth to investigate how the spin interaction depends on the mutual rotational angle of the phenyl, because the spin interaction is terminated by the phenyl with the angle of 90 degree. Figure 8c shows the antiferromagnetic spin exchange energy J of the localized electron spin on phenalenyl moiety as a function of the rotational angle. The antifer-
TE D
260
romagnetic spin exchange interaction is the largest at the planar conformation and monotonically decrease with increasing the rotational angle of phenyl. The spin exchange interaction at the stable angle of 35 degree is approximately half
265
EP
of that of the flat conformation, indicating that the spin interaction among the phenalneyl still extend throughout the sheet under the conformation with tilted phenyl. Thus, the sheet still exhibit the antiferromagnetic ordering under the
AC C
ground state conformation.
5. Conclusion
In this work, we theoretically investigated the possible structures of 2D
270
porous hydrocarbon networks of sp2 C consisting of phenalenyl and phenyl units. Phenalenyl formed a hexagonal lattice while phenyl formed a Kagome lattice because of their alternating arrangement with each phenalenyl connected to
10
ACCEPTED MANUSCRIPT
Table 1:
Relative total energies of porous hydrocarbon sheets consisting of phenalenyl and
RI PT
phenyl groups with antiferromagnetic (AF), ferromagnetic (F), and non-magnetic (NM) states. The energies are measured from that with the planar conformation without spin polarization.
AF state [meV]
F state [meV]
NM [meV]
2D
803
817
1116
Chain
301
310
612
Dimer
159
163
471
Monomer
0
0
313
SC
Structures
Table 2: Dependence of the number of polarized electron spins (∆ρ) on hole concentration.
1
2
3
4
5
6
7
∆ρ
1
0
1
2
3
2
1
M AN U
Number of hole doping
three adjacent phenalenyls via phenyls. Our density functional theory calculations showed that the sheet exhibited unusual electronic structure near EF : 275
Radical electron states localized on phenalenyl units led to a Dirac cone at EF .
TE D
The Fermi velocity of the Dirac cone was one hundred times lower than that of graphene. Simultaneously, flat dispersion bands emerge below and above the Dirac cone arising from the HO and LU states of phenyl, respectively, which form a Kagome lattice. Slow massless electrons at EF lead to magnetic ordering 280
on the porous hydrocarbon sheet with AF and F states. The spin exchange in-
EP
teraction between adjacent radical spins on phenalenyl units is J = 10 meV per pair, preferring the singlet coupling. We also demonstrated that the electronic structures and spin states of the porous hydrocarbon networks can be tuned
AC C
by tilting the phenyl unit connecting adjacent phenalenyl via a single bond,
285
lowering the dimensionality of the electron system near EF .
Acknowledgement M.M. acknowledges a Grant-in-Aid for Japan Society for the Promotion of Science (JSPS) Fellows. This work was supported in part by a Grant-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science and 11
RI PT
ACCEPTED MANUSCRIPT
(b)
(d)
TE D
(c)
M AN U
SC
(a)
EP
Figure 1: Optimized structures of porous hydrocarbon sheets consisting of phenalenyl and phenyl groups with (a) planar, (b) 1D, (c) dimer, and (d) monomer conformation. Black and pink balls denote carbon and hydrogen atoms, respectively. Black dotted lines show the unit
AC C
cell of each network. Blue hexagons and red triangles show hexagonal and Kagome networks of phenalenyl and phenyl units, respectively.
12
ACCEPTED MANUSCRIPT
2
1
1
0
-1
-2
-2
L
K
(c)
TE D
2 1 0
EP
Energy [eV]
K
-1
AC C
-2
K
L
L
M K
(d)
3
3
-3
-3
M K
2 1
Energy [eV]
-3
0
M AN U
-1
SC
2
RI PT
(b) 3
Energy [eV]
Energy [eV]
(a) 3
0
-1 -2
M K
-3
K
L
M K
Figure 2: Electronic structures of porous hydrocarbon sheets consisting of phenalenyl and phenyl groups with (a) planar, (b) 1D, (c) dimer, and (d) monomer conformation. Energies are measured from that of EF .
13
ACCEPTED MANUSCRIPT
(d)
M AN U
SC
(c)
RI PT
(b)
(a)
Figure 3: Isosurfaces of the wave function of the Dirac cone of porous hydrocarbon sheets consisting of phenalenyl and phenyl groups with (a) planar, (b) 1D, (c) dimer, and (d) monomer conformation at the K point.
Technology of Japan and the Joint Research Program on Zero-Emission Energy
TE D
290
Research, Institute of Advanced Energy, Kyoto University. Computations were performed on a NEC SX-8/4B at the University of Tsukuba, a SGI ICE XA/UV at the Institute for Solid State Physics, The University of Tokyo, and a NEC
295
EP
SX-Ace at the Cybermedia Center, Osaka University.
References
AC C
[1] N. Hamada, S.-I. Sawada, A. Oshiyama, New one-dimensional conductors: graphitic microtubules, Phys. Rev. Lett. 68 (1992) 1579–1581.
[2] R. Saito, M. Fujita, G. Dresselhaus, M.S. Dresselhaus, Electronic structure of chiral graphene tubules, Appl. Phys. Lett. 60 (1992) 2204–2206.
300
[3] J.W. Mintmire, B.I. Dunlap, C.T. White, Are fullerene tubules metallic? Phys. Rev. Lett. 68 (1992) 631–634.
14
HOMO-1
(b) HOMO-2
(c) HOMO-2
(d) HOMO-2
HOMO-3
HOMO-3
AC C
EP
HOMO-1
TE D
HOMO-1
HOMO-3
M AN U
HOMO-2 HOMO-3
SC
(a) HOMO-1
RI PT
ACCEPTED MANUSCRIPT
Figure 4: Isosurfaces of the wave function of the Kagome flat dispersion band of porous hydrocarbon sheets consisting of phenalenyl and phenyl groups with (a) planar, (b) 1D, (c) dimer, and (d) monomer conformation at the Γ point.
15
ACCEPTED MANUSCRIPT
(a) F
(b)
F
M AN U
AF
SC
RI PT
AF
(c)
F
EP
TE D
AF
(d) F
AC C
AF
Figure 5: Spin densities of porous hydrocarbon sheets consisting of phenalenyl and phenyl groups with (a) planar, (b) 1D, (c) dimer, and (d) monomer conformation with antiferromag-
16
netic and ferromagnetic states. Colors indicate the sign of the polarized electron spin.
(b)
AF 2
1
1
1
K
L
M K
-3
K
L
3
2
2
1
1
0
-1 -2
K
L
M K
-2
-2
L
-3
M K
0
3
3
2
2
1
1
0
-1
-1
-2
-2
-3
K
AF
F
K
L
M K
-3
EP
-3
0
-1
-3
M K
TE D
Energy [eV]
Energy [eV]
AF
0
-1
K
L
M K
(d)
(c) 3
1
M AN U
-2
-2 -3
0
-1
Energy [eV]
0
F
2
Energy [eV]
2
Energy [eV]
2
-1
3
3
Energy [eV]
3
Energy [eV]
Energy [eV]
3
F
SC
(a) AF
RI PT
ACCEPTED MANUSCRIPT
F
0
-1 -2
K
L
M K
-3
K
L
M K
Figure 6: Electronic structures of porous hydrocarbon sheets consisting of phenalenyl and phenyl groups with (a) planar, (b) 1D, (c) dimer, and (d) monomer conformation with antiferromagnetic and ferromagnetic states. EF is located at zero. Black lines and blue dotted
AC C
lines denote majority and minority spin bands, respectively.
17
1h
7h
AC C
EP
6h
5h
TE D
4h
M AN U
SC
3h
RI PT
ACCEPTED MANUSCRIPT
Figure 7: Spin densities of a porous hydrocarbon sheet consisting of phenalenyl and phenyl groups with planar conformation and hole concentrations of 1h to 7h.
18
RI PT
ACCEPTED MANUSCRIPT
M AN U
(b)
(c)
0.002
0.15
0
-0.002
-J [eV]
0.05 0 -0.05 -0.1
0
10
20
TE D
Energy [eV]
0.1
-0.15
SC
(a)
30
40
50
[degree]
60
70
80
90
-0.004 -0.006 -0.008
-0.01 -0.012
0
10
20
30
40
50
60
[degree]
70
80
90
EP
Figure 8: (a) A geometric structure of a phenalenyl dimer connected via phenyl (phenalenylphenyl-phenaleny). Black and pink circles denote carbon and hydrogen atoms, respectively. (b) Calculated total energy of phenalenyl-phenyl-phenaleny as a function of the rotational angle of phenyl unit. Energy is measured from that under the flat conformation. (c) Spin
AC C
exchange energy of phenalenyl-phenyl-phenaleny as a function of the rotational angle of phenyl unit.
19
ACCEPTED MANUSCRIPT
[4] K. Tanaka, K. Okahara, M. Okada, T. Yamabe, Electronic properties of
RI PT
bucky-tube model, Chem. Phys. Lett. 191 (1992) 469–472. [5] M. Fujita, K. Wakabayashi, K. Nakada, K. Kusakabe, Peculiar localized 305
state at zigzag graphite edge, J. Phys. Soc. Jpn. 65 (1996) 1920–1923.
[6] K. Nakada, M. Fujita, G. Dresselhaus, M.S. Dresselhaus, Edge state in
Rev. B 54 (1996) 17954–17961.
SC
graphene ribbons: nanometer size effect and edge shape dependence, Phys.
[7] Y.-W. Son, M.L. Cohen, S.G. Louie, Energy gaps in graphene nanoribbons, Phys. Rev. Lett. 97 (2006) 216803–1–4.
M AN U
310
[8] Y. Miyamoto, K. Nakada, M. Fujita, First-principles study of edge states of H-terminated graphitic ribbons, Phys. Rev. B 59 (1999) 9858–9861. [9] Y. Kobayashi, K. Fukui, T. Enoki, Edge state on hydrogen-terminated graphite edges investigated by scanning tunneling microscopy, Phys. Rev. B 73 (2006) 125415–1–8.
TE D
315
[10] S. Okada, A. Oshiyama, Magnetic ordering in hexagonally bonded sheets with first-row elements, Phys. Rev. Lett. 87 (14) (2001) 146803–1–4. [11] A. Yamanaka, S. Okada, Energetics and electronic structures of graphene
320
EP
nanoribbons under a lateral electric field, Carbon 96 (2016) 351–361. [12] S. Bhowmick, Weber-fechner type nonlinear behavior in zigzag edge
AC C
graphene nanoribbons, Phys. Rev. B 82 (2010) 155448–1–10. [13] S. Bhowmick, A. Medhi, V.B. Shenoy, Sensory-organ-like response determines the magnetism of zigzag-edged honeycomb nanoribbons, Phys. Rev. B 87 (2013) 085412–1–6.
325
[14] N. Shima, H. Aoki, Electronic structure of super-honeycomb systems: a peculiar realization of semimetal/semiconductor classes and ferromagnetism, Phys. Rev. Lett. 71 (1993) 4389–4392.
20
ACCEPTED MANUSCRIPT
[15] K. Kusakabe, K. Wakabayashi, M. Igami, K. Nakada, M. Fujita, Magnetism
330
RI PT
of nanometer-scale graphite with edge or topological defects, Mol. Cryst. Liq. Cryst. 305 (1997) 445–454.
[16] M. Maruyama, S. Okada, A two-dimensional sp2 carbon network of fused pentagons: all carbon ferromagnetic sheet, Appl. Phys. Express 6 (2013)
SC
095101–1–4.
[17] S. Okada, T. Kawai, K. Nakada, Electronic structure of graphene with 335
topological line defect, J. Phys. Soc. Jpn. 80 (2011) 013709–1–4.
M AN U
[18] J. Fern´ andez-Rossier, J.J. Palacios, Magnetism in graphene nanoislands, Phys. Rev. Lett. 99 (2007) 177204–1–4.
[19] J. Wu, W. Pisula, K. M¨ ullen, Graphenes as potential material for electronics, Chem. Rev. 107 (2007) 718–747. 340
[20] K. M¨ ullen, J.P. Rabe, Nanographenes as active components of singlemolecule electronics and how a scanning tunneling microscope puts them
TE D
to work, Acc. Chem. Res. 41 (2008) 511–520. [21] M. Fujihara, Y. Miyata, R. Kitaura, Y. Nishimura, C. Camacho, S. Irle, et al., Dimerization-initiated preferential formation of coronene-based 345
graphene nanoribbons in carbon nanotubes, J. Phys. Chem. C 116 (2012)
EP
15141–15145.
[22] H.E. Lim, Y. Miyata, M. Fujihara, S. Okada, Z. Liu, Arifin, et al., Fabri-
AC C
cation and optical probing of highly extended, ultrathin graphene nanoribbons in carbon nanotubes, ACS Nano 9 (2015 ) 5034–5040.
350
[23] A.I. Cooper, Conjugated microporous polymers, Adv. Mater. 21 (2009) 1291–1295.
[24] J. Gao, D. Jiang, Covalent organic frameworks with spatially confined guest molecules in nanochannels and their impacts on crystalline structures, Chem. Commun. 52 (2016) 1498–1500. 21
ACCEPTED MANUSCRIPT
355
[25] X. Liu, J. Tan, A. Wang, X. Zhang, M. Zhao, Electron spin-polarization and
Phys. Chem. Chem. Phys. 16 (2014) 23286–23291.
RI PT
spin lattices in the boron- and nitrogen-doped organic framework COF-5,
[26] C. S´anchez-S´ anchez, S. Br¨ uller, H. Sachdev, K. M¨ ullen, M. Krieg, H.F. Bet-
tinger, et al., On-surfaces synthesis of BN-substitued heteromatic networks, ACS Nano 5 (2015) 9228–9235.
SC
360
[27] Y. Morita, S. Suzuki, K. Sato, T. Takui, Synthetic organic spin chemistry for structurally well-defined open-shell graphene fragments, Nature Chem.
M AN U
3, (2011) 197–204.
[28] V. Boekelheide, C.E. Larrabee, An investigation of the preparation and 365
some properties of perinaphthene1 , J. Am. Chem. Soc. 72 (1950) 1245– 1249.
[29] D.H. Reid, Stable π-electron systems and new aromatic structures, Tetrahedron 3 (1958) 339–352.
370
TE D
[30] K. Nakasuji, M. Yamaguchi, I. Murata, K. Yamaguchi, T. Fueno, H. OhyaNishiguchi, et al., Synthesis and characterization of phenalenyl cations, radicals, and anions having donor and acceptor substituents: three redox states of modified odd alternant systems, J. Am. Chem. Soc. 111 (1989)
EP
9265–9267.
[31] Z. Sun, J. Wu, Open-shell polycyclic aromatic hydrocarbons, J. Mater. Chem. 22 (2012) 4151–4160.
AC C
375
[32] T. Kubo, Phenalenyl-based open-shell polycyclic aromatic hydrocarbons, Chem. Rec. 15 (2015) 218–232.
[33] T. Kubo, A. Shimizu, M. Uruichi, K. Yakushi, M. Nakano, D. Shiomi, et al., Singlet biraical character of phenalenyl-based Kekule’ hydrocarbon
380
with naphthoquinoid structure, Org. Lett. 9 (2007) 81–84.
22
ACCEPTED MANUSCRIPT
[34] Z. Mou, K. Uchida, T. Kubo, M. Kertesz, Evidence of σ- and π-dimerization
RI PT
in a series of phenalenyls, J. Am. Chem. Soc. 136 (2014) 18009–18022. [35] Z-H. Cui, A. Gupta, H. Lischka, M. Kertesz, Concave or convex π-dimers:
the role of the pancake bond in substituted phenalenyl radical dimers, Phys. 385
Chem. Chem. Phys. 17 (2015) 23963–23969.
[36] K. Shiraishi, H. Tamura, H. Takayanagi, Design of a semiconductor fer-
SC
romagnet in a quantum-dot artificial crystal, Appl. Phys. Lett. 78 (2001) 3702–3704.
390
M AN U
[37] H. Ishiia, T. Nakayama, J. Inoue, Flat-band excitonic states in Kagome lattice on semiconductor surfaces, Surf. Sci. 514 (2002) 206–210. [38] P. Hohenberg, W. Kohn, Inhomogeneous electron gas, Phys. Rev. 136 (1964 ) B864–B871.
[39] W. Kohn, L.J. Sham, Self-consistent equations including exchange and cor-
395
TE D
relation effects, Phys. Rev. 140 (1965) A1133–A1138.
[40] Y. Morikawa, K. Iwata, K. Terakura, Theoretical study of hydrogenation process of formate on clean and Zn deposited Cu(111) surfaces, Appl. Surf. Sci. 169–170 (2001) 11–15.
[41] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation
400
EP
made simple, Phys. Rev. Lett. 77 (1996) 3865–3868. [42] D. Vanderbilt, Soft self-consistent pseudopotentials in a generalized eigen-
AC C
value formalism. Phys. Rev. B 41 (1990) 7892–7895.
[43] T.C. Nguyen, M. Otani, S. Okada, Semiconducting electronic property of graphene adsorbed on (0001) surfaces of SiO2 , Phys. Rev. Lett. 106 (2011) 106801–1–4.
405
[44] M. Otani, O. Sugino, First-principles calculations of charged surfaces and interfaces: a plane-wave nonrepeated slab approach, Phys. Rev. B 73 (2006) 115407–1–11. 23
ACCEPTED MANUSCRIPT
[45] H. Yuan, H. Shimotani, J. Ye, S. Yoon, H. Allah, A. Tsukazaki, M.
410
RI PT
Kawasaki, and Y. Iwasa, Electrostatic and electrochemical nature of liquidgated electric-double-layer transistors based on oxide semiconductors, J.
AC C
EP
TE D
M AN U
SC
Am. Chem. Soc. 132 (2010) 18402–18407.
24