Arul Varman
Kesavan
ab,
Atul C.
Khot
a,
Tukaram D.
Dongale
ac,
Kyung Rock
Son
a,
Praveen C.
Ramamurthy
d and
Tae Geun
Kim
*a
aSchool of Electrical Engineering, Korea University, Seoul, 02841, Republic of Korea. E-mail: tgkim1@korea.ac.kr; Fax: +82 2 924 5119; Tel: +82 2 3290 3255
bDepartment of Physics & Nanotechnology, SRM Institute of Science & Technology, Kattankulathur, 603203, India
cSchool of Nanoscience and Biotechnology, Shivaji University, Kolhapur, 416 004, India
dDepartment of Materials Engineering, Indian Institute of Science, Bangalore, 560012, India
First published on 21st September 2022
In electronic devices, the work function (WF) of the electrodes must be tailored to achieve a well-aligned Ohmic or Schottky contact. Low- and high-WF electrodes are typically used to ensure effective injection/extraction of electrons and holes. In this study, composite graphite–aluminum (G:Al) and graphite–nickel (G:Ni) electrodes were deposited on a glass substrate using electron beam evaporation, and ambient pressure photoemission spectroscopy was conducted to evaluate the WF of the fabricated electrodes. The WF of the G:Al electrode was successfully tuned from 4.24 ± 0.047 eV to 5.10 ± 0.031 eV (a range of ∼0.9 eV) by increasing the graphite content. Similarly, the WF of the G:Ni composite electrode was tuned from 4.67 ± 0.041 eV to 5.11 ± 0.031 eV (a range of ∼0.4 eV). The shift in the WF in the composite graphite–metal electrodes could be explained by the formation of metal–metal (or semiconductor) junctions. The optical reflectance, sheet resistance, and morphology were also able to be tuned. The sheet resistance of the G:Al and G:Ni electrodes varied from 2.28 ± 0.03 Ω sq−1 to 80.05 ± 9.1 Ω sq−1 and from 4.92 ± 0.04 Ω sq−1 to 166.30 ± 4.1 Ω sq−1, respectively, while the total tunable reflectance was 53.77% and 45.70%, respectively. This research demonstrates a novel exploratory technique for tailoring the WF of hybrid graphite materials.
The introduction of small molecules or a non-conjugated polyelectrolyte interface layer can also change the electrical field distribution between the semiconductor and the electrode. This electric field redistribution leads to Fermi level pinning, electric double-layer formation, charge transfer, spontaneous dipole orientation, spontaneous interfacial dipole orientation,5 electron pushback (repulsion), and interface dipole formation. These effects are associated with changes in the electrode WF. The degree of this shift in the electrode WF depends on the strength of the electric field and the choice of polyelectrolyte. Acidic and basic polyelectrolytes result in dipole formation pointing inward and outward from the electrode, leading to an increase and decrease in the WF, respectively. Nonionic polyelectrolytes do not lead to a change in the WF due to null dipole formation. Notably, chemical modification influences electrode stability in the long term.
In most electronic devices, metal electrodes such as gold, silver, copper, nickel, and aluminum are routinely used, the WF of which must be tuned. Alkanethiols and other polymers and small molecules are widely used to modify the WF of Au,6 Ag,7 and Cu. The self-assembled monolayer (SAM) of alkanethiol on Au can lead to a charge transfer of less than 0.05 eV. However, the thiol (R–S⋯H) anchor group of alkanethiol adsorbed onto the Au surface leads to the formation of Au–S bonds. This observation suggests that the WF shift is primarily associated with the dipole movement of individual molecules, while gauche defects unavoidably occur in SAMs. Compared to isolated molecules, the dipole moment decreases (known as the depolarization effect) in the SAM. However, the intensity of this effect is negligible, and it is thus not considered in many practical applications.8–10
The WF of an electrode can also be modified using doping, the preparation of metal alloys, and the engineering of the crystallographic orientation. The WF of metals is an anisotropic function of the crystallographic orientation of the grains. This anisotropy is associated with the variation in the atomic packing density with the crystallographic orientation. A surface with a low atomic packing density exhibits a low WF, while a high atomic packing density leads to a high WF. The atomic packing density of many face-centered cubic (fcc) metals (Ni, Cu, and Ag) varies according to the crystallographic orientation in the following manner: (110)fcc < (001)fcc < (111)fcc. Moreover, based on the crystallographic orientation, the atomic packing fraction increases in the following order: (111)bcc < (001)bcc < (110)bcc. Therefore, the surface free energy is inversely related to the WF, leading to the following trend in the surface free energy for fcc and body-centered cubic (bcc) metals: γ(110)fcc > γ(001)fcc > γ(111)fcc and γ(111)bcc > γ(001)bcc > γ(110)bcc. The anisotropic WF of many metals can be attributed to the presence of p- and d-orbitals, which are anisotropically distributed near the Fermi surface. For Al, the p-orbital is directionally dependent, leading to anisotropic characteristics, while s–d orbital hybridization causes anisotropy in metals such as Cu and Au. In principle, many materials exhibit an anisotropic WF due to the presence of asymmetric orbitals near the Fermi surface.11,12 On the other hand, the properties of graphite can be characterized by the type of hybridization of carbon in graphite (sp3, sp2, and sp1). Higher amounts of sp3 hybridized bonds in graphite results in diamond-like carbon, which is less conductive.
Based on these considerations, this study aimed to systematically investigate the structural, optical, morphological, and electrical properties of composite graphite–metal (graphite–aluminum [G:Al] and graphite–nickel [G:Ni]) electrodes prepared using electron beam (e-beam) evaporation. The WF of the G:Al and G:Ni electrodes was tuned by modifying the electrode composition. Using the composite electrodes, the unique physical and chemical properties of the parent materials (i.e., graphite, Al, and Ni) were able to be exploited.
We analyzed the photocurrent/photoelectron spectra in accordance with the photon energy using the Fowler theory to determine the WF. Fowler's theory can be expressed as R ∞ (Eph − hυ0)2, where R is the obtained photocurrent (electrons) per incident photon, Eph is the incident photon energy, h is Planck's constant, and υ0 is the threshold frequency. The density of states (DOS) near the Fermi energy level considerably influences the mechanical, chemical, and physical properties of a material. To determine the DOS for the nanocomposite electrodes, the photocurrent/photoelectron intensity (Y) was plotted against the photon energy (E) was done. The plot of dY/dE against the photon energy yielded the DOS distribution.13
The absorption, transmittance, and reflectance of the neat graphite and composite graphite–Al/Ni electrodes were measured using a UV spectrometer (PerkinElmer Lambda 35). The total reflectance spectra of the composite electrodes were obtained using the UV spectrometer with an integrating sphere setup. The equation 1-R-T = A was used. An optical microscope (Olympus BX51M) was used to examine the composite electrode surface. Atomic force microscopy (Park Systems XE series) was employed in non-contact mode to analyze the root mean square (RMS) roughness and the surface morphology.
(1) |
(2) |
(3) |
(4) |
Fig. 1 XRD spectra for the composite (a) G:Al and (b) G:Ni electrodes. Sheet resistance for the (c) G:Al and (d) G:Ni electrodes on a log scale. |
Fig. 1b displays the XRD spectra for the G:Ni electrodes and Ni (fcc). The calculated 2θ, FWHM, and crystallite size for these electrodes are also presented in Table S4 (ESI†), and the variation with respect to the graphite content is shown in Fig. S2a–c (ESI†). As the graphite content increased, the 2θ of the (111) peak shifted to lower values (from 45.02° for Ni to 43.68° for G8:Ni2). This decrease in 2θ of ∼1.34° was related to an increase in the lattice constant a for Ni (eqn (2)), which indirectly indicated a change in the unit cell volume (V). The FWHM of Ni and G8:Ni2 was 0.66° and 6.29°, respectively, demonstrating that the crystallite size for G8:Ni2 (1.4 nm) was smaller than that for the Ni (13.5 nm) electrode. The variation in 2θ, FWHM, and crystallite size according to the G:Ni ratio is presented in Fig. S2d–f (ESI†). The variation in the lattice constant and the interplanar spacing was calculated for the composite electrodes (Table S4, ESI†). For the G:Ni electrodes, as the graphite content increased, both interplanar spacing d and the lattice constant a increased.
Due to the mechanisms described above, the addition of graphite to the Ni electrode led to an increased Rs. The sheet resistance of various G:Ni composite electrode is presented in Table S2 (ESI†). The pure Ni electrode exhibited a sheet resistance of 1.244 ± 0.005 Ω sq−1 (d-119.1 ± 11 nm), while the composite electrodes had a sheet resistance range of 161 Ω sq−1 (from 4.903 ± 0.043 Ω sq−1 for G1:Ni9 to 166.3 ± 4.101 Ω sq−1 for G9:Ni1). For the G:Al and G:Ni electrodes, regardless of their surface defects or porosity, the sheet resistance increased with increasing graphite content due to the low conductivity of graphite. The electrical resistivity (eqn (5)) and conductivity (eqn (6)) were determined using Rs, while the factors influencing the DC conductivity of the G:Al and G:Ni composite electrodes are summarized in eqn (7).
ρ = Rsd | (5) |
(6) |
(7) |
In general, charge transfer is a spontaneous process that occurs at any interface between two materials that are in contact (metals, semiconductors, or insulators). The charge transfer mechanism across the G:Al and G:Ni interfaces and the change in the WF in these composite electrodes can be understood in terms of metal–metal (or semiconductor) junction formation and the associated band energy structure (Fig. S6a–e, ESI†). Metal–metal junctions are usually formed between two conducting materials. When a metal–metal (or semiconductor) Ohmic contact is formed between materials with dissimilar WFs, a contact potential is generated. Electrons in the low-WF metal tunnel into the high-WF metal until the Fermi level of both metals is similar at equilibrium (i.e., no external bias at a given temperature). The electrons from the high-WF metal also flow to the low-WF metal and occupy empty energy levels, resulting in a minimum Helmholtz free energy. At equilibrium, a new EF is generated that differs from that of the individual metals. This process results in a positive pole in the high-WF metal and a negative pole in the low-WF metal. Consequently, at equilibrium, a contact potential is generated between the metals, the magnitude of which depends on the difference in the WF of the contacting metals.31,32Fig. 2a and b present a schematic overview of the graphite–Al/Ni system before and after junction formation, respectively. Similarly, Fig. 2c–e display the charge density distribution, electric field distribution, and electric potential at the junction, respectively. The change in the WF of the composite electrodes according to the graphite content is shown in Fig. 2f and g, respectively. In both cases, the WF increased as the graphite content increased in the G:Al and G:Ni electrodes. The difference in WF between the electrodes can be explained by the formation of metal–metal (or semiconductor) Ohmic junctions. The WF of the neat Al and graphite electrodes was 4.24 ± 0.047 eV and 5.10 ± 0.031 eV, respectively (Fig. 2f). In the G:Al composites, however, the electrons flowed from Al to the graphite, thus generating a new Fermi level that depended on the ratio of Al and graphite.
The Fermi level alignment in these composites can also be explained based on the graphite and Al phase distributions within the composite film. The phase distributions can be divided into three categories: (i) low graphite content, (ii) low Al (or high graphite) content, and (iii) an equal ratio of graphite and Al. In the electrodes with a low graphite content, such as G1:Al9 (4.353 eV), G2:Al8 (4.421 eV), G3:Al7 (4.478 eV), and G4:Al6 (4.499 eV) (Table S1, ESI†), a larger region is covered by the Al phase and an extremely small region is occupied by graphite. In this configuration, the graphite is assumed to be embedded in the Al matrix like a series of islands (Fig. S7, ESI†), which generates more electron-donating states (i.e., the graphite boundary area) and fewer electron-accepting empty states (i.e., the Al boundary area). Therefore, only some of the electrons from Al can tunnel into the low-energy states of graphite at equilibrium. Nanoscale metal–metal (or semiconductor) Ohmic junctions form, followed by electron transfer. In other words, the volume fraction of the active interfacial area across the graphite and Al is low, thus only a marginal change in the WF is observed for the electrodes with a low Al content.
The change in the contact potential (ΔV) for the G1:Al9, G2:Al8, G3:Al7, G4:Al6, G5:Al5, G6:Al4, G7:Al3, G8:Al2, G9:Al1 and G electrodes is presented in Fig. 2h. The actual contact potential between graphite and Al is ∼0.9 eV, whereas the contact potential for the composites was less than this. A similar charge transfer mechanism was expected to occur in the G:Ni composite electrode. Fig. 2g and i show the difference in the WF and contact potential, respectively, between the G:Ni composite electrodes. The measured WF for the pristine Ni and graphite electrodes was 4.673 ± 0.041 eV and 5.111 ± 0.031 eV, respectively. For the C:Ni composite electrode system, the contact potential (ΔV) and tunable WF range was approximately 0.5 eV. In principle, the volume fraction of the active interfacial boundary formed between graphite and Ni dictated the change in the WF.
The LDOS, which is the DOS primarily associated with the valence band electrons, is the number of electronic states present in the unit energy range. The energy distribution and carrier concentration can be obtained from the DOS. The LDOS near the Fermi energy level of materials considerably influences the physical and chemical properties of a material.13 In this study, the LDOS was derived from the wavelength-dependent photoelectron emission spectra of the composite electrodes. Fig. 3a and b present the photoemission spectra for the G:Al and G:Ni electrodes, respectively, while the local DOS (LDOS; the valence band DOS) of the G:Al and G:Ni electrodes is presented in Fig. 3c and d, respectively, with the specific values summarized in Table S5 (ESI†). As the graphite content in the G:Al and G:Ni electrodes increased, the LDOS state decreased. The LDOS corresponding to the Fermi energy is the LDOS that is associated with the energy in the LDOS state plot. Thus, the DOS corresponding to the Fermi energy can be estimated based on the x-axis on the LDOS plot. The spectra for the variation in the integrated LDOS with the energy for the G:Al and G:Ni electrodes are displayed in Fig. S8a and b (ESI†), respectively, while the composition-dependent LDOS for the G:Al and G:Ni electrodes are presented in Fig. S8c and d (ESI†), respectively.
The surface morphology and RMS roughness of the G:Al and G:Ni electrodes were also examined using AFM in noncontact mode (Fig. 5). As the Al content in the G:Al electrodes increased, larger grains appeared on the surface (Fig. 5: G1:Al9, G2:Al8, G3:Al7, G4:Al6, G5:Al5, G6:Al4, G7:Al3, G8:Al2, G9:Al1). Though the RMS roughness of the G:Al electrodes varied with the Al content, the maximum roughness was 10 nm (Fig. S10a, ESI†). The average thickness of the G:Al and G:Ni electrodes is presented in Fig. S10c and d (ESI†). The difference in the crystallite size for the G:Al electrodes is summarized in Table S3 (ESI†). Similar trends were also noted for the G:Ni composite electrode (Fig. 5: G1:Ni9, G2:Ni8, G3:Ni7, G4:Ni6, G5:Ni5, G6:Ni4, G7:Ni3, G8:Ni2, G9:Ni1). In the composite G:Al/G:Ni electrodes, the graphite was amorphous in nature, whereas Al and Ni were in crystalline form. The difference in the crystallite size for the G:Ni electrodes is given in Table S4 (ESI†), while the RMS roughness obtained from AFM is listed in Table S1 and S2 (ESI†). Fig. S10b (ESI†) displays the average RMS roughness for the G:Ni electrodes and AFM images of the Ni electrode, Al electrode and graphite film are presented in Fig. S11 (ESI†).
Fig. 5 Surface topography (area: 1 × 1 μm) of the G:Al and G:Ni electrodes measured using AFM in noncontact mode. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2tc02848d |
This journal is © The Royal Society of Chemistry 2022 |