Philippe F. Weck*a,
Eunja Kimb and
Grant W. Biedermanna
aSandia National Laboratories, Albuquerque, NM 87185, USA. E-mail: pfweck@sandia.gov
bDepartment of Physics and Astronomy, University of Nevada Las Vegas, Las Vegas, NV 89154, USA
First published on 21st April 2015
The interaction of Cs adatoms with mono- or bi-layered graphene (MLG and BLG), either free-standing or on a SiO2 substrate, was investigated using density functional theory. The most stable adsorption sites for Cs are found to be hollow sites on both graphene sheets and graphene-veiled SiO2(0001). Larger dipole moments are created when a MLG-veiled SiO2(0001) substrate is used for adsorption of Cs atoms compared to the adsorption on free-standing MLG, due to charge transfer occurring between the MLG and the SiO2 substrate. For the adsorption of Cs on BLG-veiled SiO2(0001) substrate, these differences are smoothed out and the binding energies corresponding to different sites are nearly degenerate; smaller dipole moments created by the Cs adatoms on BLG compared to MLG are also predicted.
However, the practical realization of such AM-containing nanodevices with satisfactory signal-to-noise performance has been hindered by motional heating in micro-fabricated traps. The experimental spectra over the observed range of frequencies suggest a 1/f variation with frequency and are consistent with a d−4 scaling, where d is the trap-surface distance, for a large variety of trapping geometries. This anomalous noise-induced heating contrasts with the most common noise source in conductors, the Johnson–Nyquist noise from the trap components, that is characterized by a frequency independent spectrum and decays as d−2 with increasing trap-surface distance. Recent theoretical research efforts have focused on the development of a microscopic model for this excess electric field noise, which points a way towards a more systematic understanding of surface adsorbates as progenitors of electric field jitter.9,10
One of the current strategies to limit or suppress this 1/f noise in nanodevices consists in stacking graphene monolayers with the natural graphite order on an insulating substrate.11 Silicon dioxide (SiO2) is widely utilized as a dielectric substrate in nanodevices and exfoliated graphene deposition on SiO2 or direct production of graphene on SiO2 by chemical vapour deposition were recently demonstrated.12–14 A limited number of first-principles studies of monolayer graphene (MLG) and bilayer graphene (BLG) on SiO2 substrate were also reported and focused on the perturbation of the graphene band structure by the oxide substrate.15–18
While the adsorption of AM on graphitic systems has been the subject of active experimental and computational research,19 no study has been carried out on the interaction of AM adatoms with MLG or BLG on SiO2 substrate, to the best of our knowledge.
In this study, atomistic models within the framework of density functional theory (DFT) are used to investigate the interaction of cesium atoms – envisioned for use in micro-fabricated traps – with graphene and graphene-veiled SiO2. Since existing experimental results for Cs on graphite are somewhat controversial,19 computational models are used to initially study the interactions of Cs atoms with free-standing MLG and BLG. These models are then extended to investigate the adsorption of Cs atoms on graphene-veiled SiO2. Particular attention is dedicated to the change in work function as a result of Cs adsorption and associated with the creation of adatom-surface dipole moments; the fluctuation of such surface dipoles was demonstrated to be responsible for the anomalous noise-induced heating in microtraps.9,10
Details of the computational approach used are given in the next section, followed by a summary of our Results and discussion.
The interaction between valence electrons and ionic cores was described by the projector augmented wave (PAW) method.24,25 The Cs (5s2,5p6,6s1), Si (3s2,3p2), C (2s2,2p2), and O (2s2,2p4) electrons were treated explicitly as valence electrons in the Kohn–Sham (KS) equations and the remaining cores were represented by PAW pseudopotentials. The KS equations were solved using the blocked Davidson iterative matrix diagonalization scheme26 followed by the residual vector minimization method. The plane-wave cutoff energy for the electronic wavefunctions was set to 500 eV, ensuring the total energy of the system to be converged to within 1 meV per atom.
All structures were optimized with periodic boundary conditions applied using the conjugate gradient method, accelerated using the Methfessel–Paxton Fermi level smearing27 with a width of 0.1 eV. The total energy of the system and Hellmann–Feynman forces acting on atoms were calculated with convergence tolerances set to 10−3 eV and 0.01 eV Å−1, respectively. Structural optimizations and properties calculations were carried out using the Monkhorst–Pack special k-point scheme28 with 5 × 5 × 5 and 5 × 5 × 1 meshes for integrations in the Brillouin zone (BZ) of bulk and slab systems, respectively.
The supercells for MLG and BLG models had a (2 × 2) mesh unit with a ca. 20.00 Å vacuum separating periodic slabs; the BLG model used a Bernal stacking (i.e. ABAB) similar to graphite. (2 × 2) graphene supercells were chosen in order to lead to a periodically repeated geometry for graphene placed on SiO2(0001) surface. O- and Si-terminated SiO2(0001) surfaces were built by cleaving the relaxed bulk structure of α-quartz and consisted of 5 and 6 bilayers of SiO2, with the back side of the slabs passivated by addition of hydrogen atoms. In the structural optimization calculations, only the top two SiO2 bilayers were allowed to relax. Although a large vacuum region of ca. 20 Å was used between periodic slabs, the creation of dipoles upon adsorption of atoms on only one side of the slab can lead to spurious interactions between the dipoles of successive slabs. In order to circumvent this problem, a dipole correction was applied by means of a dipole layer placed in the vacuum region following the method outlined by Neugebauer and Scheffler.29 MLG and BLG were then placed on top of the O-terminated SiO2(0001) surface and further optimization of these structures was carried out before studying the adsorption of Cs on their surface.
Since we are ultimately interested in evaluating the dipole fluctuation spectrum and thereby the corresponding heating rate upon adsorption of Cs atoms onto the surface, a first step towards this goal is to assess the change in the surface electric dipole caused by adsorption of Cs adatoms. This can be tracked by calculating the change of the work function upon atomic adsorption, ΔW = Wclean − Wads, where Wads and Wclean correspond to the values of the work function for a surface with and without atoms adsorbed, respectively. The work function, W, is defined as the minimum energy required to remove an electron from a solid to the vacuum region in the vicinity of the solid surface and is given by:
W = ![]() | (1) |
The electrostatic potential V(x,y,z) on a grid in real space can be obtained from a self-consistent electronic structure calculation using a plane wave basis set. Assuming that the surface normal is oriented along the z-axis, one can define a plane averaged potential:
![]() | (2) |
![]() | (3) |
It should be noted that the major contribution to the surface dipole results from the charge reordering associated with the formation of the chemical bonds between the metal surface and the adatoms. This contribution is foremost determined by the nature of the chemical bonds, but can also be modified by the packing density of the adatoms.
The structures of the (2 × 2) MLG and BLG supercells relaxed at the LDA/PWC level of theory are depicted in Fig. 1. For MLG, the structural lattice parameters obtained at the LDA/PWC level were found to be: a = b = 4.89 Å (with a vacuum layer along the z-axis fixed to 20 Å), α = β = 90°, γ = 120°. The lattice parameters relaxed at the GGA/PBE and GGA/PW91 levels of theory resulted in a = b = 4.94 Å and a = b = 4.93 Å, respectively. These values are in good overall agreement with experimental lattice parameters for graphite: a = b = 4.912 Å (c = 6.696 Å), α = β = 90°, γ = 120°.31 As expected, experimental lattice parameters are slightly larger (smaller) than the values calculated with LDA (GGA), due to the well-known overbinding (underbinding) character of LDA (GGA). The plane-averaged electrostatic potential in the vacuum, (∞) of MLG was found to be 1.55 eV for LDA/PWC, 1.50 eV for GGA/PBE and 1.51 eV for GGA/PW91. The Fermi energy, EF, of MLG was calculated to be −3.82 eV for LDA/PWC, −3.60 eV for GGA/PBE and −3.64 eV for GGA/PW91. As a result, the computed work function, W, of MLG was: 5.37 eV for LDA/PWC, 5.10 eV for GGA/PBE, and 5.15 eV for GGA/PW91. These values appear slightly larger than the value of ca. 4.66 eV determined experimentally for MLG.32 Unless otherwise stated, all the subsequent calculations were carried out using LDA/PWC.
![]() | ||
Fig. 1 Crystal supercells of (2 × 2) (a) monolayer graphene and (b) bilayer graphene. Geometries were relaxed at the LDA/PWC level of theory. |
For BLG, similar structural parameters as in the case of MLG were obtained with LDA/PWC: a = b = 4.89 Å, with an optimized interlayer spacing of 3.18 Å, and α = β = 90°, γ = 120°. Let us note that the computed interlayer spacing of BLG is slightly shorter than the distance c/2 = 3.348 Å in graphite with graphene layers periodically stacked with a Bernal sequence. The plane-averaged electrostatic potential in the vacuum was found to be (∞) = 3.11 eV for LDA/PWC, with the Fermi energy EF = −1.95 eV, resulting in a work function of 5.06 eV.
The relaxed structures for one Cs atom adsorption on the (2 × 2) MLG calculated at the LDA/PWC level of theory are show in Fig. 2. Cs atom adsorption at “atop” sites and “hollow” sites were investigated for a 1:
8 coverage ratio of Cs/C atoms. For adsorption of Cs atoms at atop sites, the computed bond distances were d(Cs–C) = 3.01 and 3.30 Å and correspond to a Cs binding energy of EB = 0.56 eV. For adsorption of Cs atoms at hollow sites, the average bond distance was d(Cs–C) = 3.27 Å and corresponds to a Cs binding energy of EB = 0.63 eV. Therefore, Cs adsorption at the hollow sites is predicted to be more energetically favorable by 0.07 eV than at atop sites. The predicted
(∞) values were 1.45 and 1.69 eV for Cs adsorption at atop and hollow sites, respectively, with a corresponding Fermi energy value of −0.68 eV in both cases. The resulting work function W is 2.13 and 2.37 eV for adsorption at atop and hollow sites, respectively. Therefore, the change in work function, ΔW, upon adsorption of a Cs atom on the (2 × 2) MLG is 3.24 eV for adsorption at atop sites and 3.00 eV for adsorption at hollow sites.
![]() | ||
Fig. 2 Crystal supercells of Cs adsorption on a (2 × 2) graphene monolayer at (a) atop sites and (b) hollow sites. Geometries were relaxed at the LDA/PWC level of theory. |
Structural optimization of Cs-MLG was also investigated for a 4:
8 coverage ratio of Cs/C atoms, with all the Cs atoms occupying hollow sites. However, it was found that only a limited number of Cs atoms were adsorbed and a tendency to form multilayer-thick Cs cubic structures from the Cs atoms that desorbed was observed (see, e.g., Fig. 3). Such situations are likely to be encountered when the Cs pressure in the trap increases, leading possibly to epitaxial growth of a Cs thin film on the MLG substrate. If this is the case, the change in work function would be expected to correspond to the difference in work function between MLG and bulk Cs.
![]() | ||
Fig. 3 Crystal supercell of Cs adsorption on (2 × 2) graphene monolayer, with a 4![]() ![]() |
The relaxed structures for one Cs atom adsorption on the (2 × 2) BLG calculated at the LDA/PWC level of theory are shown in Fig. 4, with Cs atom adsorption at atop and hollow sites (1:
8 coverage ratio of Cs/C atoms). The calculated Cs binding energies on BLG are identical to the case of Cs adsorption on MLG, i.e. EB = 0.56 eV for Cs at atop sites and EB = 0.63 eV for Cs at hollow sites. The computed bond distances were d(Cs–C) = 2.97 and 3.26 Å for Cs adsorption at atop sites and the average bond distance was d(Cs–C) = 3.23 Å for Cs adsorption at hollow sites.
![]() | ||
Fig. 4 Crystal supercells of Cs adsorption on a (2 × 2) graphene bilayer at (a) atop sites and (b) hollow sites. Geometries were relaxed at the LDA/PWC level of theory. |
The (∞) values were 2.88 and 2.96 eV for Cs adsorption at atop and hollow sites, respectively, with a corresponding Fermi energy value of 0.70 eV for Cs atop-site adsorption and 0.57 eV for Cs hollow-site adsorption. The resulting work function W is 2.18 and 2.39 eV for adsorption on BLG at atop and hollow sites, respectively; only small variations are predicted with regard to the values of 2.13 and 2.37 eV for adsorption on MLG at atop and hollow sites. Therefore, the change in work function, ΔW, upon adsorption of a Cs atom on the (2 × 2) BLG is 2.88 eV for adsorption at atop sites and 2.67 eV for adsorption at hollow sites. These changes in work function are slightly smaller than in the case of adsorption on a MLG (3.00 and 3.24 eV for hollow and atop adsorption, respectively), thus suggesting the formation of a smaller dipole moment created by the adatoms on BLG compared to MLG.
The trigonal crystal unit cell of bulk α-quartz (SiO2; space group P3221, IT no. 154) was relaxed at the LDA/PWC level of theory (cf. Fig. 5). The optimized lattice parameters were a = b = 4.91 Å, c = 5.40 Å, α = β = 90°, γ = 120°. These values are in excellent agreement with the diffraction measurements of Levien and co-workers,33 i.e. a = b = 4.916 Å, c = 5.4054 Å, α = β = 90°, γ = 120°.
![]() | ||
Fig. 5 The trigonal crystal unit cell of bulk α-quartz (SiO2; space group P3221, IT no. 154) relaxed at the LDA/PWC level of theory. |
As shown in Fig. 6, O- and Si-terminated SiO2(0001) surfaces were built by cleaving the relaxed bulk structure of α-quartz (Fig. 5) and consisted of 5 and 6 bilayers of SiO2, with the back side of the slabs passivated by addition of hydrogen atoms; a large vacuum region of ca. 20 Å was used between periodic slabs.
For the clean O-terminated SiO2(0001) surface, the vacuum potential (∞) was found to be 4.29 eV and the Fermi-level energy was −5.33 eV, resulting in a work function of 9.62 eV. For Si-terminated SiO2(0001) surface,
(∞) was calculated to be 3.76 eV and the Fermi energy −2.83 eV, thus the predicted work function was 6.59 eV, significantly lower than for the O-terminated quartz surface.
The optimized structures of O- and Si-terminated SiO2(0001) slabs covered with one MLG calculated at the LDA/PWC level are displayed in Fig. 7. The computed bond distances were d(C–O) = 1.45–1.46 Å and d(C–Si) = 2.06–2.10 Å for O- and Si-terminated slabs, respectively. The binding energy of the (2 × 2) MLG to O-terminated SiO2(0001) is about 3.20 eV and about 1.28 eV to Si-terminated SiO2(0001). For the (2 × 2) MLG-veiled SiO2(0001) slab with O-atom termination, (∞) = 3.63 eV and EF = −2.57 eV, thus the predicted work function W is 6.20 eV. Therefore, the addition of a MLG changes the work function of the clean O-terminated surface by ΔW = 3.42 eV. For the (2 × 2) MLG-veiled SiO2(0001) slab with Si-atom termination,
(∞) = 4.84 eV and EF = −0.28 eV, resulting in a smaller work function of W = 5.12 eV. The addition of a MLG changes the work function of the Si-terminated surface by only ΔW = 1.47 eV. This smaller change in work function upon covering the Si-terminated SiO2(0001), relative to the O-terminated quartz surface, is expected due to the weaker C–Si bonding compared to the strong C–O bonding [see above, d(C–O) < d(C–Si)]. Let us note that, under normal atmospheric conditions, O-terminated SiO2(0001) will be predominant relative to Si-terminated SiO2(0001). Therefore, we will focus on modelling O-terminated SiO2(0001) throughout the rest of this study.
The optimized structures for the adsorption of Cs at hollow sites of (2 × 2) MLG-veiled and BLG-veiled SiO2(0001) slabs with O-atom termination are depicted in Fig. 8 and 9, respectively.
![]() | ||
Fig. 8 Optimized structures for the adsorption of Cs at hollow sites of a (2 × 2) MLG-veiled SiO2(0001) slab with O-atom termination. Geometries were relaxed at the LDA/PWC level of theory. |
Three locations were studied for the adsorption of Cs at hollow sites of a (2 × 2) MLG-veiled SiO2(0001) slab [Fig. 8(a)–(c)]. The computed bond distances d(Cs–C) were all in the range 3.25–3.33 Å with the presence of four to six bonds. The computed binding energies were: (a) 0.54 eV; (b) 0.67 eV; (c) 0.52 eV. These values are close to the value of EB = 0.63 eV for Cs adsorption at hollow sites on a free-standing MLG. The predicted (∞) and EF values were, respectively: (a) 3.02 and 0.81 eV; (b) 3.23 and 0.70 eV; (c) 3.07 and 0.83 eV. Therefore, the corresponding work functions were: (a) 2.21 eV; (b) 2.53 eV; (c) 2.24 eV. In turn, the change in work function ΔW as a result of Cs adsorption at these three locations was: (a) 3.99 eV; (b) 3.67 eV; (c) 3.96 eV. These ΔW values are slightly larger than in the case of the adsorption of a Cs atom at hollow sites on the free-standing (2 × 2) MLG, i.e. ΔW = 3.00 eV. This means that larger adatom-surface dipole moments are created when a MLG-veiled quartz substrate is used compared to the adsorption on MLG alone.
The optimized structures of a clean (2 × 2) BLG-veiled SiO2(0001) slab and for the three possible adsorption sites of Cs at hollow sites of a (2 × 2) BLG-veiled SiO2(0001) slab with O-atom termination are shown in Fig. 9. For the clean BLG-veiled SiO2(0001) slab [Fig. 9(a)], the interlayer spacing for the BLG ranges between 3.11 and 3.48 Å, due to the buckling of the bottom graphene sheet as a result of the strain created by the formation of C–O covalent bonds and the slight lattice mismatch between the (2 × 2) graphene and the SiO2(0001) surface. This interlayer spacing is in line with the value of 3.18 Å calculated for a free-standing BLG and the value of 3.348 Å determined experimentally for graphite. For the clean BLG-veiled SiO2(0001) slab, the predicted (∞) and EF values were 4.53 and −0.91 eV, respectively, thus resulting in a work function of 5.44 eV. This work function is slightly larger than the value of 5.06 eV for the free-standing BLG, therefore suggesting the occurrence of charge transfer between the BLG and the SiO2(0001) surface.
The Cs adsorption at three different hollow sites of the (2 × 2) BLG-veiled SiO2(0001) slab [Fig. 9(b) and (c)] has been investigated. The computed Cs–C bond distances were all in the range 3.22–3.24 Å with the presence of six bonds per Cs atom; this contrasts with the case of Cs adsorption on GML where Cs exhibited both four- and six-fold coordination to the graphene sheet. The calculated binding energies for (b) and (c) are essentially degenerate with a value of EB = 0.65 eV, in contrast to Cs adsorption on the MLG-veiled SiO2(0001) which featured adsorption energies between 0.52 and 0.67 eV.
The computed (∞) and EF values were, respectively: (b) 4.02 and 1.52 eV; (c) 4.00 and 1.51 eV; (d) 4.02 and 1.52 eV. This leads to work function values of (b) 2.50 eV, (c) 2.49 eV, (d) 2.50 eV, and, therefore, a nearly identical change in work function of around ΔW = 2.94 eV as a result of Cs adsorption at these three locations shown in Fig. 9. This value is comparable to the ΔW = 3.00 eV obtained for adsorption of a Cs atom at hollow sites on the free-standing (2 × 2) MLG, while significantly smaller than the values of 3.67–3.99 eV for the adsorption of Cs at hollow sites of a (2 × 2) MLG-veiled SiO2(0001) slab. This, again, suggests the formation of smaller dipole moments created by the adatoms on BLG compared to MLG.
Results from LDA/PWC calculations for the Cs binding energies, asymptotic plane-averaged electrostatic potentials in the vacuum, Fermi energies, work function, and change of the work function upon Cs atom adsorption, are summarized in Table 1 for all the cases discussed in this study.
Structures | EB | ![]() |
EF | W | ΔW |
---|---|---|---|---|---|
a Cs adsorption at hollow and atop sites are indicated by h and a sub-indexes (followed by a number when multiple adsorption sites exist). | |||||
MLG | 1.55 | −3.82 | 5.37 | ||
BLG | 3.11 | −1.95 | 5.06 | ||
O–SiO2(0001) | 4.29 | −5.33 | 9.62 | ||
Si–SiO2(0001) | 3.76 | −2.83 | 6.59 | ||
MLG/O–SiO2(0001) | 3.63 | −2.57 | 6.20 | ||
BLG/O–SiO2(0001) | 4.53 | −0.91 | 5.44 | ||
MLG/Si–SiO2(0001) | 4.84 | −0.28 | 5.12 | ||
Csa/MLG | 0.56 | 1.45 | −0.68 | 2.13 | 3.24 |
Csh/MLG | 0.63 | 1.69 | −0.68 | 2.37 | 3.00 |
Csa/BLG | 0.56 | 2.88 | 0.70 | 2.18 | 2.88 |
Csh/BLG | 0.63 | 2.96 | 0.57 | 2.39 | 2.67 |
Csh1/MLG/O–SiO2(0001) | 0.54 | 3.02 | 0.81 | 2.21 | 3.99 |
Csh2/MLG/O–SiO2(0001) | 0.67 | 3.23 | 0.70 | 2.53 | 3.67 |
Csh3/MLG/O–SiO2(0001) | 0.52 | 3.07 | 0.83 | 2.24 | 3.96 |
Csh1/BLG/O–SiO2(0001) | 0.65 | 4.02 | 1.52 | 2.50 | 2.94 |
Csh2/BLG/O–SiO2(0001) | 0.65 | 4.00 | 1.51 | 2.49 | 2.95 |
Csh3/BLG/O–SiO2(0001) | 0.65 | 4.02 | 1.52 | 2.50 | 2.94 |
Further DFT calculations will investigate the interaction potentials at various distances between the Cs adatoms and the different MLG and BLG free-standing and veiled SiO2(0001) surfaces and predict the jitter noise in these systems.
This journal is © The Royal Society of Chemistry 2015 |