Interfaces formed between monolayer transition metal dichalcogenides and (metallo)phthalocyanine molecules are promising in energy applications and provide a platform for studying mixed-dimensional molecule-semiconductor heterostructures in general. An accurate characterization of the frontier energy level alignment at these interfaces is key in the fundamental understanding of the charge transfer dynamics between the two photon absorbers. Here, we employ the first-principles substrate screening GW approach to quantitatively characterize the quasiparticle electronic structure of a series of interfaces: metal-free phthalocyanine (H2Pc) adsorbed on monolayer MX2 (M = Mo, W; X = S, Se) and zinc phthalocyanine (ZnPc) adsorbed on MoX2 (X = S, Se). Furthermore, we reveal the dielectric screening effect of the commonly used α-quartz (SiO2) substrate on the H2Pc:MoS2 interface using the dielectric embedding GW approach. Our calculations furnish a systematic set of GW results for these interfaces, providing the structure–property relationship across a series of similar systems and benchmarks for future experimental and theoretical studies.
I. INTRODUCTION
Interfaces formed between a pair of semiconductors feature intriguing electronic and optoelectronic properties due to the alignment of the energy levels of one component with respect to the other at the interface. The direction of the charge transfer in the exciton splitting process depends on type of heterojunction,1 which, in turn, depends on the electronic structure of the interface. Mixed-dimensional heterostructures formed between 0D molecules and 2D substrates,2,3 in particular, organic molecules deposited on monolayer transition metal dichalcogenides (TMDs),4–6 have attracted much attention in recent years due to their promising applications in optoelectronic devices,7 photovoltaics,8 and photocatalysis.9 Among all molecular adsorbates, (metallo)phthalocyanines are notably interesting10,11 by virtue of their structural planarity, large π-conjugation, photon-absorbing capability,12 fruitful surface chemistry,13 as well as great tunability in electronic and optical properties via a change in the metal center.14–16 Once the two photon absorbers—a (metallo)phthalocyanine molecule and a monolayer TMD—are placed together to form an interface, the underlying electronic structure, i.e., the relative alignment of energy levels of the two components, dictates the mechanism and direction of the charge transfer across the interface, giving rise to distinct optoelectronic properties. To quantitatively characterize the electronic structure at these interfaces, various experimental efforts have been put forward,3,17–20 while benchmark results and a systematic account of the trends are still missing, which constitute the main goals of this paper.
Complementary to experimental techniques, first-principles calculations play a unique role in elucidating the electronic structure and structure–property relationship via modeling of atomistically well-defined systems. Notably, the energy levels at a heterogeneous interface pertinent to charge transfer are quasiparticle levels, whose accurate characterization, in principle, requires methods beyond the conventional density functional theory (DFT). Although many-body perturbation theory (MBPT), such as the GW formalism,21–23 has been very successful24–26 in resolving the gap problem of DFT,27,28 its relatively high computational cost hinders its routine applications in large systems such as the interfaces formed between (metallo)phthalocyanines and monolayer TMDs. As a result, most prior computational studies of the phthalocyanine:TMD interfaces still employ DFT with different complexities such as semi-local functionals,17,29 hybrid functionals,3,30 or the DFT + U approach.31
The key ingredient in GW that is responsible for an accurate determination of interfacial energy level alignment is the so-called surface polarization32,33 or, equivalently, dielectric screening due to the substrate. To effectively capture the dielectric screening while reducing the computational cost, the substrate screening GW was proposed in Ref. 34, which was shown to be accurate for weakly coupled interfaces with negligible orbital hybridization.34–38 In this work, we apply this approach to a series of interfaces: metal-free phthalocyanine (H2Pc) adsorbed on monolayer MX2 (M = Mo, W; X = S, Se) and zinc phthalocyanine (ZnPc) adsorbed on MoX2 (X = S, Se). We focus on the quasiparticle electronic structure, especially the frontier energy level alignment between the valence band maximum (VBM) and the conduction band minimum (CBM) of the monolayer TMD with the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of the H2Pc or ZnPc. Our results reveal the structure–property relationship and provide GW-quality benchmark results for future studies.
Furthermore, in most experimental studies, a substrate is used to support the composite molecule:TMD interface system from the bottom, with a commonly used one being α-quartz (SiO2).7,20,39–41 Due to the dielectric screening effect, substrates could greatly affect the electronic properties of the adsorbate on top of it.42–45 In this work, we use the dielectric embedding GW approach46 to address the dielectric effect of the SiO2 substrate on the H2Pc:MoS2 interface. We show that with additional screening from the SiO2 underneath, the energy level alignment at the H2Pc:TMD interface is modulated considerably. Our results on the embedded H2Pc:MoS2 system agree quantitatively with experimental measurements.41
This paper is organized as follows: In Sec. II, we detail the computational methodology and parameters. In Sec. III, we present our results in two aspects: the structure–property relationship across a series of systems and the dielectric effect of the SiO2 substrate on the H2Pc:MoS2 interface. We then conclude in Sec. IV with brief remarks. Appendixes A and B are devoted to draw a quantitative connection between two GW-based methods—interface GW and projection GW— to supplement our discussion in Sec. III B.
II. METHODOLOGY
As a first step, we relax the in-plane lattice parameter and atomic coordinates of each monolayer TMD unit cell using the vdw-DF-cx functional.47 This is the functional that we will use to relax the structure of the interfaces, so we also employ it here to ensure consistency. The calculation uses a k-mesh of 18 × 18 × 1 and a kinetic energy cutoff of 100 Ry. All DFT relaxations employ the optimized norm-conserving Vanderbilt (ONCV) pseudopotentials48,49 and the package.50 The resulting lattice constants are 3.15, 3.29, 3.15, and 3.28 Å for monolayers MoS2, MoSe2, WS2, and WSe2, respectively. These results agree very well with experimental measurements, which yield 3.15,51 3.30,52 3.15,53 and 3.28 Å53 for the four systems, respectively.
After the monolayer unit cell relaxation, we build 6 × 6 supercells and place one H2Pc molecule flat on each substrate and one ZnPc molecule flat on the MoS2 and MoSe2 substrates to form six interface systems. Each interface simulation cell is 30.0 Å along the c direction and includes about 23 Å of vacuum. During the relaxation of the interface, the atoms belonging to the substrate are kept fixed in their relaxed monolayer positions to ensure the exactness of the subsequent reciprocal-space folding of the noninteracting polarizability. The coordinates of the atoms belonging to the adsorbate molecule are fully relaxed until all residual forces are below 0.05 eV/Å. The relaxations are carried out using the vdw-DF-cx functional,47 a k-mesh of 3 × 3 × 1, and a kinetic energy cutoff of 70 Ry. We found an adsorption height of about 3.0 Å for each system we study, similar to the result of a prior calculation3 (3.3 Å). In our relaxed structures, the center of the H2Pc or ZnPc molecule is approximately at the bridge position of two S or Se atoms belonging to the top layer of TMD, which is a stable binding site for similar systems.31 We note that Ref. 54 reported that the binding energies and band alignments are roughly the same for different binding sites. Figure 1 shows the relaxed H2Pc:MoS2 structure in two different views.
(a) Side view and (b) top view of the optimized H2Pc:MoS2 structure. The black boxes represent periodic boundary conditions. This figure is rendered using VESTA.55
(a) Side view and (b) top view of the optimized H2Pc:MoS2 structure. The black boxes represent periodic boundary conditions. This figure is rendered using VESTA.55
Considering the large size of the interfaces and the associated computational cost of conventional GW, we apply the substrate screening GW approach34 for all systems and have explicitly benchmarked this approach against direct GW calculations of the H2Pc:MoS2 and H2Pc:MoSe2 systems. All GW calculations are performed at the G0W0 level and employ a mean-field starting point using the Perdew–Burke–Ernzerhof (PBE) functional,56 the Hybertsen–Louie generalized plasmon-pole model23 for the frequency dependence of the dielectric function, the semiconductor screening for the treatment of the q → 0 limit, the slab Coulomb truncation57 for the removal of spurious long-range interactions along the c direction, and the static remainder58 in the self-energy calculation to improve convergence, as implemented in the package.59 We note that our calculations do not include the spin–orbit coupling, which is known to cause a 0.4–0.5 eV splitting for WS2 and WSe2 and a 0.1–0.2 eV splitting for MoS2 and MoSe2 in the valence band.54,60
Here, we list the computational parameters involved in the substrate screening GW calculations. For the calculation of the noninteracting polarizability of the substrate unit cell, , we use a q-mesh of 18 × 18 × 1, a 5 Ry dielectric cutoff, and 200 bands in the summation. For the treatment of the q → 0 limit, 30 bands on a shifted q-grid are used. For the calculation of the noninteracting polarizability of the adsorbate molecule, , we use a simulation cell that is of the same size along a and b as the interface but is much smaller in size (10 Å) along c. We use a q-mesh of 3 × 3 × 1, a 5 Ry dielectric cutoff, and 2400 bands in the summation. For the treatment of the q → 0 limit, 360 bands on a shifted q-grid are used. After that, the is folded in the reciprocal space to a 6 × 6 supercell, and the is mapped in the real space to the interface simulation cell, following Ref. 34. These quantities are then combined at each q-point to approximate , the noninteracting polarizability of the interface. It is then inverted to generate the dielectric function in the interface simulation cell, after which the self-energies are computed (for the details of different approaches employed in this work, see below), where we use a k-mesh of 3 × 3 × 1, a 5 Ry dielectric cutoff, and 7200 bands of the interface in the summation for the Green’s function.
We consider two types of self-energy calculations for each interface, in line with our prior works. (i) “Interface GW” calculations, where we compute the expectation value of the self-energy operator using ϕtot, an orbital of the interface, i.e., ⟨ϕtot|Σ|ϕtot⟩. (ii) “Projection GW” calculations as proposed in Refs. 61 and 62, where we compute the expectation value of the self-energy operator using an orbital of the freestanding substrate (ϕsub) or that of the freestanding monolayer of adsorbate (ϕmol), i.e., ⟨ϕsub|Σ|ϕsub⟩ for the former and ⟨ϕmol|Σ|ϕmol⟩ for the latter. In both the approaches mentioned above, the self-energy operator Σ is calculated using iGtotWtot, i.e., both G and W are from the interface system with W calculated using the substrate screening approximation. We compare the interface GW and projection GW results for every system and show that they agree very well except for the LUMO of H2Pc when H2Pc is adsorbed on a MoS2 substrate. In Appendixes A and B, we show that this discrepancy is due to the strong orbital hybridization and explicitly establish a quantitative connection between the two.
At the end of this section, we clarify the terminologies we use in this paper. All results reported in Sec. III B are obtained using the “substrate screening GW”34 approach, where we compute the self-energy operator using iGtotWtot, together with the approximation . Within this framework, two types of self-energies are computed, “interface GW,” ⟨ϕtot|Σ|ϕtot⟩, and “projection GW,” ⟨ϕmol|Σ|ϕmol⟩ or ⟨ϕsub|Σ|ϕsub⟩, as discussed above. Note that the “substrate screening GW”34 is physically identical to the “XAF-GW” approach,36 while the difference between the two is only technical: the former first computes the in a small cell and then performs a real-space mapping of this quantity to the interface simulation cell and hence is more computationally efficient than the latter.
All results reported in Sec. III C are obtained using the “dielectric embedding GW” approach46 (but without the real-space truncation of as done in Ref. 46). Here, we compute ⟨ϕad|Σ|ϕad⟩, where Σ = iGadWtot. One can see that the key difference here is the use of Green’s function of the adsorbate, Gad, in the self-energy operator compared to the “projection GW” within the substrate screening framework. Because of this difference, the “dielectric embedding GW” further reduces the computational cost without sacrificing the accuracy for weakly coupled interfaces. Note that in Sec. III C, the “adsorbate” itself is already a H2Pc:MoS2 interface (i.e., both Gad and ϕad are for the H2Pc:MoS2), and only the SiO2 substrate is included implicitly as a dielectric environment through Wtot such that the “weak coupling” condition is satisfied.
III. RESULTS AND DISCUSSION
A. Convergence study
GW calculations are known to converge slowly. We describe our convergence study in this section to show that our calculations are reasonably well converged and the level alignment results are reliable.
For the monolayer MoS2 unit cell, we have checked that compared to a 10 Ry dielectric cutoff and 1000 bands in the summation, our choice of parameters (5 Ry and 200 bands) leads to a convergence in the bandgap within 0.02 eV, although the individual quasiparticle energies are off by about 0.3 eV. Our prediction of the monolayer MoS2 gap is 2.81 eV, in good agreement with prior GW calculations63 (2.80 eV). We also note that for such low-dimensional materials, the nonuniform neck subsampling (NNS) method64 leads to faster convergence. Although we do not use it here for the interface, our calculations of the monolayer MoS2 unit cell achieve reasonably good agreement (within 0.1 eV in the gap) with NNS results with a 10 Ry dielectric cutoff.
For the interface systems, using H2Pc:MoSe2 as an example, we have checked that our choice of parameters (a 5 Ry dielectric cutoff and 7200 bands) leads to a convergence of the quasiparticle energies and energy level alignments at the interface within 0.05 eV compared to using a 7.5 Ry cutoff and 9000 bands in the summation. This choice of parameter is thus adopted for all other interface systems.
B. Molecule:TMD interfaces
When a molecule is adsorbed on a semiconductor substrate, the interface could, in principle, exhibit the so-called type-I (straddling bandgap), type-II (staggered bandgap), or type-III (broken bandgap) energy level alignment.1 Specific to the H2Pc:TMD or ZnPc:TMD interfaces studied in this work, type-I and type-II heterostructures are possible based on our results below. Depending on the relative ordering between TMD and molecular levels, we further categorize the interfaces to type-Ia, type-Ib, type-IIa, and type-IIb, as schematically shown in Figs. 2(a)–2(d), respectively. All interfaces studied in this work have a direct bandgap at Γ, so all interface gaps and energy level alignments are reported for the Γ point.
Energy level alignment diagrams showing relevant gaps across the interface. is the bandgap of the TMD monolayer (H2Pc or ZnPc molecular layer) in its freestanding form, while ΔTMD (Δmol) is the gap of the TMD (H2Pc or ZnPc) within the interface system. ΔLL (ΔHH) is the gap between the TMD CBM (VBM) and the LUMO (HOMO) of the molecular layer. (a) and (b) display type-I heterostructures between TMD and the molecule. In (a) [(b)], both the VBM and CBM of the interface are localized on the TMD (molecule), which we denote “type-Ia” (“type-Ib”). (c) and (d) display type-II heterostructures between TMD and the molecule. In (c) [(d)], the VBM of the interface is localized on the molecule (TMD) and the CBM of the interface is localized on the TMD (molecule), which we denote “type-IIa” (“type-IIb”). In both (c) and (d), ΔHL denotes the bandgap of the interface.
Energy level alignment diagrams showing relevant gaps across the interface. is the bandgap of the TMD monolayer (H2Pc or ZnPc molecular layer) in its freestanding form, while ΔTMD (Δmol) is the gap of the TMD (H2Pc or ZnPc) within the interface system. ΔLL (ΔHH) is the gap between the TMD CBM (VBM) and the LUMO (HOMO) of the molecular layer. (a) and (b) display type-I heterostructures between TMD and the molecule. In (a) [(b)], both the VBM and CBM of the interface are localized on the TMD (molecule), which we denote “type-Ia” (“type-Ib”). (c) and (d) display type-II heterostructures between TMD and the molecule. In (c) [(d)], the VBM of the interface is localized on the molecule (TMD) and the CBM of the interface is localized on the TMD (molecule), which we denote “type-IIa” (“type-IIb”). In both (c) and (d), ΔHL denotes the bandgap of the interface.
In Fig. 2, we show frontier energy levels (bands) of the freestanding monolayer TMD and those of the freestanding molecular layer, together with their counterparts within the interface. Blue lines represent TMD levels, and red lines represent molecular levels. We discuss the following quantities that characterize the electronic structure of these interfaces: is the bandgap of the TMD monolayer (H2Pc or ZnPc molecular layer) in its freestanding form, while ΔTMD (Δmol) is the gap of the TMD (H2Pc or ZnPc) within the interface system. ΔLL (ΔHH) is the gap between the TMD CBM (VBM) and the LUMO (HOMO) of the molecular layer, which is of interest for both type-I and type-II heterostructures because the sign and magnitude of ΔLL (ΔHH) dictate the direction and barrier for electron (hole) transfer across the interface, respectively. For the type-II heterostructures in Figs. 2(c) and 2(d), we further consider ΔHL, the fundamental (transport) gap of the entire interface, which is between the HOMO of the molecule (the VBM of the TMD) and the CBM of the TMD (the LUMO of the molecule) for type-IIa (type-IIb). is calculated at the K point of the Brillouin zone for the TMD unit cell, and all other quantities are calculated at the Γ point.
Table I shows the computed results for all quantities labeled in Fig. 2 from both DFT and substrate screening GW. To verify that the substrate screening approximation holds for the systems, we compare substrate screening GW results with direct GW calculations for two interfaces, H2Pc:MoS2 and H2Pc:MoSe2. This comparison shows that the substrate screening GW is very accurate: for all quantities reported in Table I, substrate screening GW leads to an agreement with direct GW results within 0.05 eV. With the same accuracy, the substrate screening GW approach costs only about 30% in computing time and 10% in memory (for the χ step) compared to the direct GW for the systems studied. We also note that our GW results on H2Pc:MoS2 and ZnPc:MoS2 are largely in agreement with other calculations of the same systems (but with slightly different simulation cells) using range-separated hybrid functionals65 and GW.54
Key descriptors of the electronic structure for different interface systems as calculated from DFT (using the PBE functional) and GW. All values are in eV. and are the bandgaps of the freestanding TMD monolayer and molecular layer, respectively. is calculated at the K point for the unit cell, and is calculated at the Γ point. ΔTMD, ΔLL, ΔHL, ΔHH, and Δmol are the energy level differences within the interface systems defined in Fig. 2, all calculated at the Γ point.
Interface . | Method . | Type . | . | ΔTMD . | ΔLL . | ΔHL . | ΔHH . | Δmol . | . |
---|---|---|---|---|---|---|---|---|---|
H2Pc:MoS2 | DFT | IIa, Fig. 2(c) | 1.79 | 1.79 | 0.09 | 1.21 | 0.58 | 1.30 | 1.39 |
GW | IIa, Fig. 2(c) | 2.81 | 2.78 | 0.14 | 2.33 | 0.45 | 2.47 | 3.86 | |
H2Pc:MoS2:SiO2 | GW | IIa, Fig. 2(c) | 2.81 | 2.07 | 0.68 | 1.62 | 0.45 | 2.30 | 3.86 |
H2Pc:MoSe2 | DFT | IIb, Fig. 2(d) | 1.55 | 1.55 | 0.24 | 1.31 | 0.06 | 1.37 | 1.37 |
GW | Ia, Fig. 2(a) | 2.40 | 2.39 | 0.20 | ⋯ | 0.18 | 2.77 | 3.86 | |
H2Pc:WS2 | DFT | Ib, Fig. 2(b) | 1.93 | 1.96 | 0.24 | ⋯ | 0.36 | 1.36 | 1.36 |
GW | IIa, Fig. 2(c) | 3.05 | 2.98 | 0.14 | 2.76 | 0.22 | 2.90 | 3.86 | |
H2Pc:WSe2 | DFT | IIb, Fig. 2(d) | 1.68 | 1.65 | 0.54 | 1.11 | 0.26 | 1.37 | 1.39 |
GW | Ia, Fig. 2(a) | 2.65 | 2.44 | 0.05 | ⋯ | 0.35 | 2.84 | 3.86 | |
ZnPc:MoS2 | DFT | IIa, Fig. 2(c) | 1.79 | 1.80 | 0.27 | 1.15 | 0.65 | 1.42 | 1.42 |
GW | IIa, Fig. 2(c) | 2.81 | 2.78 | 0.59 | 2.29 | 0.49 | 2.88 | 3.91 | |
ZnPc:MoSe2 | DFT | Ib, Fig. 2(b) | 1.55 | 1.55 | 0.16 | ⋯ | 0.04 | 1.35 | 1.35 |
GW | Ia, Fig. 2(a) | 2.40 | 2.38 | 0.28 | ⋯ | 0.30 | 2.96 | 3.91 |
Interface . | Method . | Type . | . | ΔTMD . | ΔLL . | ΔHL . | ΔHH . | Δmol . | . |
---|---|---|---|---|---|---|---|---|---|
H2Pc:MoS2 | DFT | IIa, Fig. 2(c) | 1.79 | 1.79 | 0.09 | 1.21 | 0.58 | 1.30 | 1.39 |
GW | IIa, Fig. 2(c) | 2.81 | 2.78 | 0.14 | 2.33 | 0.45 | 2.47 | 3.86 | |
H2Pc:MoS2:SiO2 | GW | IIa, Fig. 2(c) | 2.81 | 2.07 | 0.68 | 1.62 | 0.45 | 2.30 | 3.86 |
H2Pc:MoSe2 | DFT | IIb, Fig. 2(d) | 1.55 | 1.55 | 0.24 | 1.31 | 0.06 | 1.37 | 1.37 |
GW | Ia, Fig. 2(a) | 2.40 | 2.39 | 0.20 | ⋯ | 0.18 | 2.77 | 3.86 | |
H2Pc:WS2 | DFT | Ib, Fig. 2(b) | 1.93 | 1.96 | 0.24 | ⋯ | 0.36 | 1.36 | 1.36 |
GW | IIa, Fig. 2(c) | 3.05 | 2.98 | 0.14 | 2.76 | 0.22 | 2.90 | 3.86 | |
H2Pc:WSe2 | DFT | IIb, Fig. 2(d) | 1.68 | 1.65 | 0.54 | 1.11 | 0.26 | 1.37 | 1.39 |
GW | Ia, Fig. 2(a) | 2.65 | 2.44 | 0.05 | ⋯ | 0.35 | 2.84 | 3.86 | |
ZnPc:MoS2 | DFT | IIa, Fig. 2(c) | 1.79 | 1.80 | 0.27 | 1.15 | 0.65 | 1.42 | 1.42 |
GW | IIa, Fig. 2(c) | 2.81 | 2.78 | 0.59 | 2.29 | 0.49 | 2.88 | 3.91 | |
ZnPc:MoSe2 | DFT | Ib, Fig. 2(b) | 1.55 | 1.55 | 0.16 | ⋯ | 0.04 | 1.35 | 1.35 |
GW | Ia, Fig. 2(a) | 2.40 | 2.38 | 0.28 | ⋯ | 0.30 | 2.96 | 3.91 |
Figure 3 shows the GW interfacial energy level alignment for the six heterostructures, where we use different colors to represent different substrates or molecules, and all energy levels are measured with respect to a common vacuum. In Fig. 3, solid bars or lines are interface GW results (same as those reported in Table I), i.e., ⟨ϕtot|Σ[GtotWtot]|ϕtot⟩, where the ϕtot is chosen as the interface orbital that mostly resembles the orbital of interest (HOMO, LUMO, VBM, or CBM) of the freestanding monolayer TMD or molecular layer, as quantitatively determined from orbital projections. Dashed lines are projection GW results, i.e., ⟨ϕsub|Σ[GtotWtot]|ϕsub⟩ for the TMD, where ϕsub is the VBM or CBM of the freestanding TMD, and ⟨ϕmol|Σ[GtotWtot]|ϕmol⟩ for the molecule, where ϕmol is the HOMO or LUMO of the freestanding molecular layer. For all cases, except for the H2Pc LUMO on MoS2 substrate, the interface GW results agree very well with projection GW results, which indicates negligible orbital hybridization upon formation of the interface such that ⟨ϕtot|ϕsub⟩ and ⟨ϕtot|ϕmol⟩ are close to unity. The special case of H2Pc LUMO on the MoS2 substrate is discussed in Appendixes A and B.
Energy level alignment for the molecule:TMD interfaces as calculated from substrate screening GW. The bars denote the bands of MoS2 (pink), MoSe2 (blue), WS2 (green), and WSe2 (purple). The lines denote the energy levels of H2Pc (orange) and ZnPc (red). Solid bars or lines indicate results from interface GW calculations (same as those in Table I), and dashed lines indicate results from projection GW calculations. All energy levels are measured with respect to vacuum.
Energy level alignment for the molecule:TMD interfaces as calculated from substrate screening GW. The bars denote the bands of MoS2 (pink), MoSe2 (blue), WS2 (green), and WSe2 (purple). The lines denote the energy levels of H2Pc (orange) and ZnPc (red). Solid bars or lines indicate results from interface GW calculations (same as those in Table I), and dashed lines indicate results from projection GW calculations. All energy levels are measured with respect to vacuum.
Below in this section, we discuss the results from three aspects: (i) the renormalization of gaps upon the formation of the interface, (ii) the qualitative difference between DFT and GW in the prediction of the type of some heterostructures, and (iii) the structure–property relationship across the six different systems.
For the freestanding monolayer TMD, our GW calculations yield bandgaps of 2.81, 2.40, 3.05, and 2.65 eV for MoS2, MoSe2, WS2, and WSe2, respectively. Our results are in good agreement with Ref. 63, where scGW0 calculations of monolayer TMDs with similar lattice parameters (within 0.01 Å of our relaxed values) were performed, resulting in bandgaps of 2.80, 2.40, 3.11, and 2.68 eV, respectively, for the four materials. For the freestanding H2Pc (ZnPc) molecular layer, our GW calculations yield a HOMO–LUMO gap of 3.86 (3.91 eV) and an ionization potential of 6.24 (6.16 eV), on par with Ref. 66, where GW calculation of an isolated H2Pc molecule yielded a bandgap of 3.67 eV and an ionization potential of 6.08 eV. The difference between our results and Ref. 66 lies in the physical difference between a periodic molecular layer and an isolated molecule.
At the interface, the gap renormalization for the H2Pc or ZnPc molecule is significant with about 1 eV decrease in the HOMO–LUMO gap when the molecule is brought in contact with the substrate. This is consistent with well-established understanding of the surface renormalization at molecule–substrate interfaces.32,33 On the contrary, Table I shows that TMD bandgaps are still very similar to those of the freestanding phase, indicating negligible gap renormalization for the substrate, which we attribute to the relatively low coverage of the molecule on the substrate (see Fig. 1). This situation is different from a previous semiconductor–semiconductor interface that we studied before,37 where we observed gap renormalization on both sides of an organic bulk heterojunction. For the H2Pc:MoS2:SiO2 interface, the MoS2 gap is significantly renormalized due to the dielectric screening effect of the extensive SiO2 substrate underneath, as we will elaborate in Sec. III C. Needless to say, DFT clearly underestimates the relevant gaps and does not capture the gap renormalization at the interface.
A remarkable observation here is that for some systems, DFT and GW yield qualitatively different heterostructure types, as listed in Table I. To be specific, for H2Pc:MoSe2 and H2Pc:WSe2, DFT predicts a type-II, while GW predicts a type-I interface; for H2Pc:WS2, DFT predicts a type-I, while GW predicts a type-II interface; for ZnPc:MoSe2, although both DFT and GW predict type-I, the relative ordering of TMD and molecular levels are opposite (type-Ia and type-Ib as shown in Fig. 2). In all cases, it is the self-energy correction to the LUMO of the molecular adsorbate that pushes this orbital upward in energy, causing a change in the heterostructure type (there is an additional effect in ZnPc:MoSe2, where the self-energy correction to the HOMO of ZnPc shifts this orbital downward). We note that the experimental characterizations of these quasiparticle energy level alignments require techniques such as (inverse) photoemission spectroscopy, while most existing experiments3,7,18,20,67,68 focused on the description of excitons at such interfaces using, e.g., photoluminescence. Although we have not found direct experimental verification of most of the quasiparticle energy level alignments that we have computed (except for H2Pc:MoS2:SiO2), we believe our work provides a reference point for future experiments.
Table I and Fig. 3 reveal the structure–property relationship that is helpful in materials design of similar systems. All sulfur-based interfaces form type-IIa heterostructures with the HOMO (LUMO) of the adsorbed molecule lying higher than VBM (CBM) of the TMD substrate. All selenium-based ones form type-Ia heterostructures with the HOMO (LUMO) of the adsorbed molecule lying lower (higher) than the VBM (CBM) of the TMD substrate. The gap renormalization of H2Pc is larger on Mo-based substrates compared to the W-based counterparts. Moreover, comparing ZnPc with H2Pc on the same substrates, we find that the gap renormalization of ZnPc is smaller, resulting in larger ΔLL and ΔHH values than H2Pc-based interfaces. Since these values represent the charge transfer barrier across the interface in the exciton splitting process, this trend suggests that the electron or hole transfer rates across H2Pc-based interfaces might be generally higher than those across ZnPc-based interfaces.
C. The effect of SiO2 substrate on the H2Pc:MoS2 interface
In typical experimental studies, the molecule:TMD interfaces are further supported by a substrate underneath the monolayer TMD.41,42 It is well known that the bandgaps of TMDs are sensitive to the dielectric environment provided by other adjacent 2D materials or substrates.43,44,69 Therefore, it is imperative to include the dielectric screening effects from any additional substrates to achieve quantitative agreement with experimental characterization of the electronic structure of the molecule:TMD interfaces. In this work, we employ the dielectric embedding GW approach as developed in Ref. 46 to include the effect of the substrate using the H2Pc:MoS2 system as an example. We focus on this system because an experimental measurement is available41 for a direct comparison.
A commonly used substrate in experimental studies of molecule:TMD interfaces is the α-quartz (SiO2).3,7,20,39–41 To model the composite H2Pc:MoS2:SiO2 system, we have applied a 5% compressive strain to the experimental lattice constant of SiO2 to enforce a commensurate simulation cell with H2Pc:MoS2. Under this strain, the PBE bandgap of 5.93 eV for bulk SiO270 increases by about 4% while other qualitative features of the band structure remain intact. Using the dielectric embedding GW approach,46 we first compute the noninteracting polarizability of the SiO2 substrate in its unit cell, then fold this quantity in reciprocal space to the supercell, and finally, combine the folded quantity with the noninteracting polarizability of the H2Pc:MoS2 (which, in turn, is calculated using the substrate screening GW34 approach). As a result, the self-energy calculation is only explicitly performed for the H2Pc:MoS2 interface, where the W includes the dielectric effect of the SiO2 substrate.
The dielectric embedding GW approach46 applied here adds little extra computational cost compared to the GW calculation of H2Pc:MoS2 without the SiO2 substrate. The inner products between the H2Pc:MoS2 orbitals of interest (the orbitals that represent the resonances of H2Pc HOMO/LUMO and MoS2 VBM/CBM) and the H2Pc:MoS2:SiO2 orbitals are close to unity, suggesting negligible orbital hybridization between H2Pc:MoS2 and SiO2 and the validity of the dielectric embedding GW approach. We note in passing that within this approach, the orbital hybridization within H2Pc:MoS2, i.e., between H2Pc and MoS2, is captured exactly.
Figure 4(a) shows the optimized structure of the H2Pc:MoS2:SiO2 system, where we consider a two-layer SiO2 substrate with silicon termination. Other terminations exist,71,72 and we have checked explicitly that for a hydrogen-passivated SiO2 surface, the band alignment results stay unchanged compared to what we will report below. The separation between the SiO2 and MoS2 is optimized using vdw-DF-cx to be 3.5 Å. Figure 4(b) shows the quasiparticle electronic structure of the H2Pc:MoS2 interface embedded in the dielectric environment of SiO2, computed using the dielectric embedding GW approach,46 with key energy gaps listed in Table I. Compared to the H2Pc:MoS2 system without the SiO2 substrate, the MoS2 bandgap is further renormalized to 2.07 eV due to the additional dielectric screening from SiO2, which is in very good agreement with the experiment (2.10 eV as in Ref. 41). The H2Pc gap is further renormalized only moderately, possibly due to its large distance to the SiO2 substrate. Furthermore, the ΔLL is considerably changed to 0.68 eV (compared to 0.14 eV without the SiO2 substrate), and the ΔHL is changed to 1.62 eV (compared to 2.33 eV without the SiO2 substrate). Interestingly, the ΔHH does not change compared to the case without the SiO2 substrate.
(a) Optimized structure of the H2Pc:MoS2:SiO2 interface. (b) Quasiparticle energy level alignment from an embedding GW calculation, where the H2Pc:MoS2 interface is embedded in the dielectric environment of a SiO2 substrate. All energy levels are measured with respect to vacuum.
(a) Optimized structure of the H2Pc:MoS2:SiO2 interface. (b) Quasiparticle energy level alignment from an embedding GW calculation, where the H2Pc:MoS2 interface is embedded in the dielectric environment of a SiO2 substrate. All energy levels are measured with respect to vacuum.
As a direct comparison with the experiment, Ref. 41 characterized the energy level alignment of the H2Pc:MoS2 interface deposited on the SiO2 substrate. Ultraviolet photoemission spectroscopy and inverse photoemission spectroscopy measurements were performed for pristine monolayer MoS2 and the H2Pc:MoS2 interface, both on the SiO2 substrate. By aligning the Fermi level of the two systems, Ref. 41 deduced that ΔTMD = 2.1 eV, ΔLL = 1.0 eV, ΔHL = 1.2 eV, ΔHH = 0.9 eV, and Δmol = 2.2 eV. We note that our results agree quantitatively with Ref. 41 in ΔTMD and Δmol values, but our computed H2Pc HOMO and LUMO levels are both lower by 0.3–0.4 eV than those reported in Ref. 41, resulting in lower ΔLL, lower ΔHH, and higher ΔHL. We discuss two possible sources for the discrepancy. First, Ref. 41 aligned the Fermi level of the pristine MoS2 (without the H2Pc adsorbate) and that of the H2Pc:MoS2 interface. The precise position of the Fermi level within the bandgap might depend on the specific experimental condition, while it is not well-defined in first-principles calculations at zero temperature. This difference between an experiment and a computation affects how the MoS2 and H2Pc levels are relatively aligned without affecting the bandgap of each component. Second, the coverage of the H2Pc adsorbate on the MoS2 substrate might be different between Ref. 41 and our modeling, which might lead to different interface dipoles and consequently different ΔLL, ΔHL, and ΔHH values.
The result is a clear indication of the dielectric screening effect of the substrate on molecule:TMD interfaces. Our dielectric embedding GW approach provides an accurate account of this effect without further increasing the computational cost compared to calculations that only involve molecule:TMD interfaces. Finally, we note in passing that our calculated GW bandgap is 8.69 eV for bulk SiO2, in agreement with Ref. 73, which reports a GW gap of 8.77 eV. Our GW gap is 7.10 eV for the freestanding bilayer SiO2 with Si termination and 6.64 eV when the bilayer SiO2 is in contact with monolayer MoS2, as shown in Fig. 4(b). These gaps are large enough such that the precise positions of SiO2 band edges relative to TMD bands will unlikely have large effects on the energy level alignment at the H2Pc:MoS2 interface.
IV. CONCLUSION
In this work, we have systematically and quantitatively characterized the quasiparticle electronic structure of a series of interfaces formed between (metallo)phthalocyanine molecules and monolayer TMDs using first-principles GW calculations. Besides the well-known gap renormalization of the adsorbate, we have found that in certain cases, GW and DFT yield qualitatively different heterostructure types, which is of interest for future experimental validation. Furthermore, we have elucidated the dielectric screening effect of the SiO2 substrate on the electronic structure of H2Pc:MoS2, leading to quantitative agreement with existing experiments. Our findings provide a useful structure–property relationship for materials design, insight into charge transfer processes across such molecule:TMD interfaces, as well as benchmark data for future theoretical and experimental investigations.
ACKNOWLEDGMENTS
Z.-F.L. acknowledges support from NSF CAREER Award No. DMR-2044552. O.A. acknowledges A. Paul and Carole C. Schaap Endowed Distinguished Graduate Award and Rumble Fellowship from Wayne State University. This research used computational resources at the Wayne State Grid and additionally via a user project at the Center for Nanoscale Materials at Argonne National Laboratory, an Office of Science User Facility, which was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-06CH11357. Furthermore, large-scale calculations used computational resources from the National Energy Research Scientific Computing Center (NERSC), a U.S. Department of Energy Office of Science User Facility operated under Contract No. DE-AC02-05CH11231.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.
APPENDIX A: DERIVING PROJECTION GW RESULTS FROM INTERFACE GW CALCULATIONS
We investigate the quantitative difference between projection GW calculations (dashed lines in Fig. 3) and interface GW calculations (solid lines in Fig. 3). They agree very well except for the H2Pc LUMO on the MoS2 substrate, which involves strong orbital hybridization. Here, we quantitatively connect the projection GW results with interface GW results.
To make the discussion self-contained in the appendix, we repeat here that in “projection GW,”61,62 we compute ⟨ϕsub|Σ|ϕsub⟩ for the TMD substrate and ⟨ϕmol|Σ|ϕmol⟩ for the adsorbed molecule, where ϕsub (ϕmol) is an orbital of the freestanding substrate (molecule). In “interface GW,” we compute ⟨ϕtot|Σ|ϕtot⟩ with ϕtot, where ϕtot is an orbital of the interface that mostly resembles a substrate or molecular orbital (i.e., a resonance). In both cases, Σ is the same operator that involves the G and W of the entire interface.
We focus on the H2Pc LUMO on the MoS2 substrate and expand the H2Pc LUMO (calculated for the freestanding H2Pc molecular layer in the same simulation cell as the interface) in terms of the orbitals of the H2Pc:MoS2 interface,
Here, we have neglected expansion coefficients whose magnitude is below 0.05. The projection GW computes . Substituting Eq. (A1), this quantity is related to diagonal and off-diagonal matrix elements of Σ involving the three interface orbitals on the right-hand side of Eq. (A1). These matrix elements are from interface GW calculations, which are then connected to projection GW results that involve the left-hand side of Eq. (A1). For diagonal elements, we use the difference between the quasiparticle energies and the corresponding Kohn–Sham eigenvalues. For off-diagonal elements, we use the matrix elements of Σ − Vxc, with Vxc being the exchange-correlation potential operator.
The expansion coefficients and matrix elements (in eV) are
Using these values from an interface GW calculation, one can calculate from Eq. (A1) that = 1.011 eV. This is in very good agreement with the projection GW calculation of the H2Pc LUMO on MoS2, whose self-energy correction is 1.023 eV. The above analysis successfully explains the difference between the solid and dashed lines in Fig. 3: the solid line computes , which only involves the first term on the right-hand side of Eq. (A1), while the dashed line computes , which involves the left-hand side of Eq. (A1).
As a limiting case, if only one of the coefficients in Eq. (A1) is unity, then the isolated molecular orbital and its resonance at the interface are the same (the weak-coupling limit). The projection GW and the interface GW will then yield identical values in the quasiparticle energy. This is, in fact, a very good approximation for most interfaces. Specific to the systems studied in this work, the two approaches agree well for every resonance except for the H2Pc LUMO on the MoS2 substrate.
APPENDIX B: DERIVING INTERFACE GW RESULTS FROM PROJECTION GW CALCULATIONS
In a similar manner to the above analysis, we can derive interface GW results from projection GW calculations of the substrate and of the adsorbate, which can also explain the difference between the solid and dashed lines in Fig. 3. For the H2Pc LUMO on the MoS2 substrate, its resonance in the interface is the CBM+2 of the interface. We then consider the following expansion:
We again neglect expansion coefficients whose magnitude is below 0.05. The left-hand side of Eq. (B1) is the orbital used in an interface GW calculation, and the right-hand side consists of orbitals used in projection GW calculations of the substrate and of the adsorbate, respectively.
The expansion coefficients and the matrix elements (in eV) are
and the off-diagonal elements are all close to zero.
Using these values from projection GW calculations of the TMD and of the molecule, one can derive from Eq. (B1) that eV. This is in very good agreement with the interface GW calculation of the CBM + 2 of the H2Pc:MoS2 system (the resonance of H2Pc LUMO), whose self-energy is 0.667 eV.