0
Views
0
CrossRef citations to date
0
Altmetric
Reviews

Probing excitons with time-resolved momentum microscopy

, &
Article: 2378722 | Received 15 Jan 2024, Accepted 06 Jul 2024, Published online: 26 Jul 2024

ABSTRACT

Excitons – two-particle correlated electron-hole pairs – are the dominant low-energy optical excitation in the broad class of semiconductor materials, which range from classical silicon to perovskites, and from two-dimensional to organic materials. The study of excitons has been brought on a new level of detail by the application of photoemission momentum microscopy – a technique that has dramatically extended the capabilities of time- and angle resolved photoemission spectroscopy. Here, we review how the photoelectron detection scheme enables direct access to the energy landscape of bright and dark excitons, and, more generally, to the momentum-coordinate of the exciton wavefunction. Focusing on two-dimensional materials and organic semiconductors, we first discuss the typical photoemission fingerprint of excitons in momentum microscopy and highlight that it is possible to obtain information not only on the electron- but also hole-component. Second, we focus on the recent application of photoemission orbital tomography to such excitons, and discuss how this provides a unique access to the real-space properties of the exciton wavefunction. We detail how studies performed on two-dimensional transition metal dichalcogenides and organic semiconductors lead to very similar conclusions, and, in this manner, highlight the strength of momentum microscopy for the study of optical excitations in semiconductors.

1. Introduction

Angle-resolved photoemission spectroscopy (ARPES) provides access to the binding energy E and the momentum k of photoelectrons in a condensed matter material. The photoelectron distribution therefore contains direct information on the equilibrium single-particle band structure, and also on many-body correlation effects via the electronic self energy  [Citation1,Citation2]. In order to access the non-equilibrium properties of a material after an external perturbation, ARPES experiments can be performed in a pump-probe scheme. In time-resolved ARPES (trARPES)  [Citation3–5], a pump laser pulse photoexcites the sample before a time-delayed probe laser pulse is used to detect the energy- and in-plane momentum-resolved photoemission spectrum. Very generally, the optical excitation creates a non-equilibrium distribution of electrons and holes that respond to the screening cloud of surrounding quasiparticles and dissipate their excess energy in secondary scattering processes. For materials with a high charge carrier density, the Coulomb attraction between the single-particle electrons and holes is efficiently screened, and the non-equilibrium distribution of single-particle electrons and holes thermalizes by carrier-carrier and carrier-phonon scattering events. In this manner, trARPES (aka time-resolved two-photon photoemission spectroscopy (tr2PPE)) has been successfully applied to a huge class of materials, ranging from pristine and adsorbate covered surfaces  [Citation6–9] to quantum materials such as graphene  [Citation10–13], topological insulators  [Citation14–17] and the melting of charge-ordered states  [Citation18,Citation19].

For semiconductors that are characterized by a small charge carrier density and a weak dielectric screening, the Coulomb interaction between the optically excited single-particle electrons and holes is not efficiently screened and correlated electron-hole two-particle states with significant binding energy Ebin - so-called excitons – are formed  [Citation20–24]. Excitons cannot be described in the single-particle band structure, and they manifest as additional peaks in absorption spectra at energies of one exciton binding energy below the single-particle conduction band. Being in most cases the lowest energy excitation, it is crucial to understand the excitonic response of semiconductors to an optical excitation on a fundamental level.

Given that excitons represent electronic excitations that are usually in the visible frequency range, their excitation energies and relaxation dynamics are often studied with methods such as (time-resolved) absorption or fluorescence spectroscopy. As a photon-in-photon-out technique, besides notable exceptions  [Citation25–27], all-optical techniques are only sensitive to transitions that couple to the light field. Fluorescence, for example, can only be used to observe excitons that can decay via the emission of a photon in a radiative process and thus have a measurable oscillator strength  [Citation28–30]. Hence, so-called dark excitons, where the electron- and the hole-component have a finite center-of-mass momentum or are of different spin, can only be accessed if an additional scattering event with a phonon or a spin-flip occurs  [Citation31–34]. Consequently, these optical spectroscopies provide no information on the momentum of the exciton, which, however, is key to the full description of the exciton wavefunction  [Citation35]. In contrast, as a photon-in-electron-out technique, trARPES provides direct access to the energy- and the momentum-coordinate of the photoemitted single-particle electrons that originate from the breakup of two-particle excitons. Therefore, momentum-indirect and spin-forbidden (optically dark) excitons can be accessed in a trARPES experiment.

While time-resolved photoemission spectroscopy has already been applied to various types of excitons in different material classes  [Citation36–45], recent developments on the side of the femtosecond laser light sources, the photoelectron analyzer, and also the material aspect have established trARPES as one of the most powerful tools to study exciton dynamics  [Citation46–60]. First, as mechanical exfoliation of two-dimensional materials nowadays allows for atomic-scale control of artificial heterostructures  [Citation61], several materials have become available that show strong excitonic response and a large exciton binding energy in the range of a few 100 meV  [Citation62,Citation63]. This has made excitonic systems more accessible to trARPES, as the large binding energy enables to discriminate photoelectrons originating from excitons or from the conduction band. Second, the rapid development of photoelectron spectroscopy detectors has culminated in the development of the photoelectron momentum microscope, which enables efficient detection of full 2D momentum-, kinetic energy-, and even spin-resolved photoelectron spectra  [Citation64–67]. Third, the development of momentum microscopes was complemented by the development of high-repetition rate femtosecond extreme ultraviolet light sources  [Citation68–73]. Here, the high repetition rate enables an efficient combination with momentum microscopes, where only about one electron/pulse can be detected. At the same time, time-resolved momentum microscopy becomes possible with simultaneous access to electrons with large in-plane momenta > 1 Å−1  [Citation74], as is necessary to probe the full Brillouin zone of typical two-dimensional materials and the orbital fingerprints of organic semiconductors. This is highly beneficial for exciton spectroscopy, as the multidimensional resolution directly allows to distinguish between various optically bright and dark excitonic states, and even allows access to the spatial exciton wavefunction in two-dimensional and organic semiconductors  [Citation48–50,Citation56,Citation57,Citation75–77] within the framework of orbital tomography  [Citation78]. Given this rapid progress in recent years, here, we present a review of the recent efforts in this research field to characterize excitons in two-dimensional and organic semiconductors. Importantly, we do this from the perspective of the momentum-resolved measurement scheme and discuss how this information can be exploited to reconstruct information on the real-space wavefunction of the optical excitation, i.e. the exciton.

The review article is structured as follows: In section 2, we review the basic theory of excitons, and answer the question how it is, in theory, possible to probe an exciton by measuring single-particle photoelectrons. We also present a concise overview of the recently developed capabilities of photoemission momentum microscopy. In section 3, we will first focus on the impact that momentum microscopy has had on the study of excitons in two-dimensional TMDs and organic semiconductors. Thereby, we will discuss how the energy- and momentum-resolved photoelectron detection can be used to quantify the energy landscape of excitons and how distinct signatures of the exciton’s hole-component are found in the photoemission signal. In section 4, we outline how the framework of orbital tomography has been used to reconstruct the real-space wavefunction of excitons in TMDs and organic semiconductors.

2. Methods: probing excitons with momentum microscopy

The first question that comes to mind when discussing photoemission spectroscopy of excitons is: how does an exciton show-up in an energy- and momentum-resolved photoemission spectrum? This question follows naturally from the basic description of an exciton: an exciton is an excited state of the many-body electron wavefunction of a semiconductor that can be conveniently described as a quasiparticle consisting of a Coulomb-correlated electron-hole pair. Thus, the theoretical description of an exciton requires a description of the many-body electron wavefunction, while in trARPES, single-particle photoelectrons are emitted and detected. In a simplistic picture, the photoemission process can be described as follows (): After a light pulse with photon energy ωpump (red) excites the semiconductor and creates an exciton, the photoemission laser pulse ωprobe (blue) excites the already-excited electron to a free-electron-like state, and thereby breaks the Coulomb-correlated electron-hole pair. The resulting single-particle photoelectron then travels to the surface, leaves the sample, and is detected with the photoelectron analyzer, while the single-particle hole remains in the sample. This simplified scenario provides an intuitive representation of the photoemission process, however, also a series of major questions arise: Very generally, in how far does the trARPES experiment provide access to information on the exciton’s electron and hole components and also the correlation between both? More specifically, at which single-particle photoelectron energy is the former exciton’s electron detected? Which information is contained in the momentum-resolved photoemission signal? Can the framework of orbital tomography be applied to reconstruct the real-space distribution of the exciton wavefunction?

Figure 1. Description of excitons and their detection in a time- and angle-resolved photoemission spectroscopy experiment. (a) Schematic illustration of the exciton formation and detection process. A pump laser pulse (red) is used to excite optically bright excitons that reside in a single TMD layer. In the photoemission process, the Coulomb correlation between the electron and the hole is broken, the single-particle electron is detected with the photoelectron analyzer and the single-particle hole remains in the sample. (b) Schematic energy level diagram indicating the single-particle valence band maximum (or HOMO) and conduction band minimum (or LUMO) at binding energies EVBM and ECBM, respectively, that are separated by the single-particle band gap energy Eg. In this picture, the binding energy Ebin of a two-particle exciton can be defined by comparing the exciton energy Eexc with the single-particle band gap Eg. (c) In the exciton picture (shown for w=0), Coulomb correlated electron-hole pairs are described based on EquationEquation (2) and have a parabolic dispersion with regard to their center-of-mass momentum Q. Within the light cone (orange area, vanishing Q), excitons are labeled to be optically bright because they can be excited by light and can decay in a radiative process. In contrast, excitons with a finite Q are termed optically dark. (d) A related pictorial description of excitons can be drawn in the electron-hole picture. Here, in addition, momentum-indirect excitons are sketched where the electron- and the hole-component are separated by momentum w and reside in different valleys of the Brillouin zone. Panel a is adopted from ref.  [Citation53] under Creative Commons Attribution License 4.0 (CC BY). Panels c,d are adopted from ref.  [Citation79].

Figure 1. Description of excitons and their detection in a time- and angle-resolved photoemission spectroscopy experiment. (a) Schematic illustration of the exciton formation and detection process. A pump laser pulse (red) is used to excite optically bright excitons that reside in a single TMD layer. In the photoemission process, the Coulomb correlation between the electron and the hole is broken, the single-particle electron is detected with the photoelectron analyzer and the single-particle hole remains in the sample. (b) Schematic energy level diagram indicating the single-particle valence band maximum (or HOMO) and conduction band minimum (or LUMO) at binding energies EVBM and ECBM, respectively, that are separated by the single-particle band gap energy Eg. In this picture, the binding energy Ebin of a two-particle exciton can be defined by comparing the exciton energy Eexc with the single-particle band gap Eg. (c) In the exciton picture (shown for w=0), Coulomb correlated electron-hole pairs are described based on EquationEquation (2)(2) Eexc(Q)=Eg−Ebin+Q22M.(2) and have a parabolic dispersion with regard to their center-of-mass momentum Q. Within the light cone (orange area, vanishing Q), excitons are labeled to be optically bright because they can be excited by light and can decay in a radiative process. In contrast, excitons with a finite Q are termed optically dark. (d) A related pictorial description of excitons can be drawn in the electron-hole picture. Here, in addition, momentum-indirect excitons are sketched where the electron- and the hole-component are separated by momentum w and reside in different valleys of the Brillouin zone. Panel a is adopted from ref.  [Citation53] under Creative Commons Attribution License 4.0 (CC BY). Panels c,d are adopted from ref.  [Citation79].

In the following, we will first briefly describe the current state-of-the art of time-resolved momentum microscopy experiments. Next, we will review theoretical proposals on how to detect photoelectron signals from excitons.

2.1. Femtosecond momentum microscopy – a new variant of trARPES

Time- and angle-resolved photoemission spectroscopy (trARPES) is a powerful technique to map the electronic band structure, identify many-body interactions and quantify ultrafast energy dissipation processes under equilibrium and non-equilibrium conditions  [Citation2–5,Citation17–19,Citation80–82]. Hence, with the fabrication of two-dimensional monolayers and their heterostructures, the idea to apply trARPES to study their non-equilibrium response has been straightforward. At the same time, it has been highly desirable to extend the framework of orbital tomography  [Citation78] into the non-equilibrium regime in order to reconstruct orbitals of initially unoccupied states  [Citation76] or even excitons  [Citation48–50,Citation55–57,Citation83]. For the case of van-der-Waals semiconductors, early experiments could address bulk and wafer-scale materials  [Citation49,Citation84–93]; however, major experimental challenges had to be overcome before high-quality exfoliated van-der-Waals heterostructures could be studied  [Citation46–48,Citation50–54,Citation94,Citation95]. The first experimental challenge lies in the small diameter of the exfoliated sample systems and the limited spatial resolution of common trARPES experiments. Second, extreme ultraviolet photon pulses at several hundreds of kHz repetition rate needed to be generated before high-quality momentum microscopy experiments could be reported.

As an integrating technique, the spatial resolution of established trARPES experiments equipped with hemispherical energy analyzers is determined by the spot sizes of the probe laser pulse on the surface. While typical laser beam diameters are in the range of 100 µm and more, the highest quality exfoliated two-dimensional materials have diameters of roughly 10 µm. Hence, it is not possible to study such samples with established laser-based table-top experiments. In static ARPES experiments performed at synchrotron light sources, this challenge has been overcome by using Fresnel zone plates or capillary devices to reduce the beam diameter on the sample to the sub-1-µm regime  [Citation96,Citation97]. By scanning the laser-spot across the sample, the occupied electronic band structure and many-body interaction effects of mono-, homo- and heterolayers of two-dimensional materials have been characterized  [Citation98–100]. It has even been possible to perform ARPES experiments on electrostatically gated samples  [Citation101,Citation102]. While such experiments are fully established and well known under the acronyms µARPES and nanoARPES, until the recent addition of momentum microscopes in ARPES experiments, it has not been possible to extend µARPES and nanoARPES towards femtosecond time-resolved experimental schemes using extreme ultraviolet light pulses.

We want to emphasize that the usage of extreme ultraviolet pulses is indispensable for the study of exciton dynamics in two-dimensional and organic materials. This is because of restrictions in the photoemission process  [Citation1,Citation74,Citation103] that require photon energies larger than 20 eV in order to probe large in-plane momenta. In materials such as graphene and TMDs, for example, the K valleys are found at ≈ 1.7Å−1 and ≈ 1.2Å−1, respectively, and are not accessible with usual femtosecond laser sources. Likewise, without extreme ultraviolet light pulses, it would not be possible to map the dynamics of the momentum fingerprint of a molecular orbital, as necessary for the evaluation of the trARPES data in the framework of photoemission orbital tomography  [Citation56,Citation76,Citation78,Citation83,Citation104]. With the development of high-repetition rate, high-harmonic generation beamlines  [Citation68–72] and the newest generation free-electron lasers  [Citation105,Citation106], it is possible to generate bright EUV light pulses with pulse durations in the attosecond to few-femtosecond range. However, the added complexity of (sub-)micron-scale focusing at EUV wavelengths has so far hindered focusing these pulses to achieve time-resolved EUV µARPES and nanoARPES.

Nevertheless, the last few years have shown that a complementary photoelectron detection scheme is extraordinarily well suited for trARPES of exfoliated samples: Momentum microscopes  [Citation64,Citation66] represent a family of photoelectron spectroscopy devices that are based on the electrostatic lens system of a (photoemission) electron microscope (PEEM). Crucially, this lens system can be controlled such that not only the real-space image plane can be projected onto the detector but also the Fourier plane (i.e. momentum space) is accessible (). Combined with spectral selection, i.e. via time-of-flight and position-sensitive photoelectron detectors, momentum microscopes then provide three-dimensional (3D) access to the photoemission intensity as a function of the photoelectron energy E and two in-plane momenta kx and ky  [Citation66,Citation67,Citation71,Citation107–109] () or the photoelectron energy E and two real-space coordinates rx and ry  [Citation54,Citation110,Citation111]. In order to access exfoliated samples with a diameter of 10 µm, it is then straightforward to insert an aperture into the real-space plane of the microscope and to measure only the photoelectrons coming from the selected region of interest. Note that, depending on the available apertures and lens settings, this approach also allows to set regions of interest with a diameter less than 10 µm. Naturally, this then increases the measurement time.

Figure 2. Schematic illustration of the momentum microscopy setup and the accessible multi-dimensional photoemission data. (a) Momentum microscopes are a novel type of photoelectron analyzers that are assembled by a microscope type electronic lens system, a time-of-flight drift tube or a hemisphere-based energy filter, and a position sensitive photoelectron detector. By projecting either the Fourier or the real-space plane onto the detector, spectrally resolved momentum maps (b) or real-space maps of the sample (c) can be collected. By inserting an aperture into the real-space plane, a region-of-interest with a diameter of approximately 10 µm can be selected on the sample. (d) If the momentum microscope is equipped with a time-of-flight detector, it is possible to simultaneously collect three-dimensional data sets that contain information on the kinetic energy E and two in-plane momenta (kx, ky) of the photoelectrons. The 3-dimensional data representation shows the occupied bands of homobilayer 2H-MoS2 and excitonic photoemission signal at a pump-probe delay of 140 fs. Panels a, b, and c are reproduced from ref.  [Citation50] (Copyright by Springer Nature). The data in panel d was taken in the Göttingen ultrafast momentum microscopy laboratory.

Figure 2. Schematic illustration of the momentum microscopy setup and the accessible multi-dimensional photoemission data. (a) Momentum microscopes are a novel type of photoelectron analyzers that are assembled by a microscope type electronic lens system, a time-of-flight drift tube or a hemisphere-based energy filter, and a position sensitive photoelectron detector. By projecting either the Fourier or the real-space plane onto the detector, spectrally resolved momentum maps (b) or real-space maps of the sample (c) can be collected. By inserting an aperture into the real-space plane, a region-of-interest with a diameter of approximately 10 µm can be selected on the sample. (d) If the momentum microscope is equipped with a time-of-flight detector, it is possible to simultaneously collect three-dimensional data sets that contain information on the kinetic energy E and two in-plane momenta (kx, ky) of the photoelectrons. The 3-dimensional data representation shows the occupied bands of homobilayer 2H-MoS2 and excitonic photoemission signal at a pump-probe delay of 140 fs. Panels a, b, and c are reproduced from ref.  [Citation50] (Copyright by Springer Nature). The data in panel d was taken in the Göttingen ultrafast momentum microscopy laboratory.

The strategy to perform femtosecond momentum microscopy experiments on exfoliated van-der-Waals materials is then as follows: First, in the real-space mode of the microscope, the sample structure can be mapped based on the photoemission intensity contrast. As an example, the orange and the dark red polygons in indicate monolayers of WSe2 and MoS2, respectively. Subsequently, three-dimensional ARPES spectra can be collected by inserting the real-space aperture in the respective region of interest (orange and red circles in ) and then projecting the Fourier plane onto the photoelectron detector (). In combination with a pump-probe scheme, multi-dimensional data sets of the temporal evolution of the non-equilibrium dynamics of quasiparticles can be measured.

2.2. The exciton wavefunction

Before it is possible to set up a theory of photoemission from excitons, it is first necessary to place the excitonic wavefunction on a theoretical basis. An exciton is essentially caused by a renormalization of the many-body electron wavefunction after an optical excitation, which requires by definition a step beyond ground-state theory. Whereas ground-state electronic structure calculations such as density functional theory (DFT) or the many-body interaction-corrected GW framework provide a good description of the valence and conduction orbitals ϕv(r) and χc(r) and their corresponding energies of the unexcited system, these theories ultimately describe single-quasiparticle orbitals and do not directly provide a good indication of either the exciton energy spectrum or the associated exciton wavefunctions. Rather, the wavefunction of an exciton must generally be expressed as a linear combination of electron-hole pairs ϕv(rh)χc(re). Instead of a single (electron) coordinate, this expression depends on the real-space coordinates re and rh of the electron and hole wavefunctions, respectively. In the Tamm-Dancoff approximation [Citation112], the wavefunction of the mth exciton can be written as

(1) ψm(rh,re)=v,cAvc(m)ϕv(rh)χc(re).(1)

Here, the expansion coefficients Avc(m) describe the relative contribution of each electron-hole-pair to the exciton wavefunction. Equivalently, |Avc(m)|2 indicates the probability to simultaneously find an electron in the conduction orbital χc and a hole in the valence orbital ϕv (note that χ and ϕ indicate the conduction and valence orbitals of the same material). The exciton energies Eexc(m) and the corresponding coefficients Avc(m) can be calculated by several methods  [Citation113,Citation114]. Notably, solving the Bethe-Salpeter equation enables a fully-interacting calculation of the exciton spectrum  [Citation112,Citation115,Citation116]. Similar information can be retrieved from time-dependent density functional theory (TD-DFT) using the Casida formalism  [Citation117–119]; however, this approach is still limited by the description of the many-body correlations that lead to the formation of bound excitonic states  [Citation120,Citation121].

In principle, the full exciton wavefunction (EquationEquation (1)) includes excitation and de-excitation contributions of all valence and conduction orbitals (bands) in the material, which can challenge the intuitive understanding of the exciton wavefunction. However, often the situation is not so complex: many exciton states in semiconducting TMDs, in particular those of low energy, are described by a hole in a single valence and an electron in a single conduction band (). Similarly, in organic semiconductors, the energetically lowest excitonic states are often, but not always  [Citation24], composed of an electron in the lowest unoccupied molecular orbital (LUMO) and a hole in the highest occupied molecular orbital (HOMO). Also for higher-lying excitons, the number of orbitals or bands contributing is often limited  [Citation24,Citation122].

2.3. Excitons in the band structure of a two-dimensional semiconductor

To understand the benefits of trARPES for the study of excitons, it is first useful to consider the prominent case of excitons in the family of semiconducting TMDs, which include WS2, WSe2, MoS2 and MoSe2  [Citation23]. These materials already show a significant excitonic response in the bulk phase, but they are especially interesting in their monolayer form: here, a direct band gap forms at the K valleys  [Citation123,Citation124], which, among others, hosts the bright A1s-exciton  [Citation62,Citation63].

For the A1s-exciton, the electron orbitals are given by the quasi-parabolic conduction band at the K valley, while the hole orbitals are given by the negative quasi-parabolic valence band at the K valley. The interaction between electron and hole is given by the Coulomb interaction. Consequentially, the exciton can be described by a two-dimensional analogue of the hydrogen atom. For such an exciton, the exciton energy Eexc can be expressed as

(2) Eexc(Q)=EgEbin+Q22M.(2)

where Eg is the single-particle band gap () and the exciton mass M can be calculated from the effective masses of the electron and the hole in the conduction and valence bands, respectively. As in the case of the hydrogen atom, the exciton binding energy Ebin depends on the quantum number n, and, for the lowest energy exciton n = 1, can be on the order of 0.5 eV  [Citation62,Citation125]. The last term of EquationEquation (2) carries information on the kinetic energy of the exciton, which is parameterized by the kinetic center-of-mass (COM) momentum Q.

If a direct semiconductor with a single valley is optically and resonantly excited with the exciton energy Eexc, the momenta of the exciton’s electron and hole are ke and −kh, respectively, with |ke| = |kh| such that Q=ke+kh=0 (). In the case that Q0, the electron-hole pairs are shifted out of the light-cone, and, because momentum needs to be conserved in a radiative decay process, they are not anymore straightforwardly accessible in all-optical experiments. These excitons are therefore termed to be optically dark. In addition to such intravalley excitons, for two-dimensional semiconductors, the thermalization of the optical excitation is strongly determined by intervalley scattering and the formation of momentum-indirect excitons where the electron- and hole-components are in different valleys of the Brillouin zone  [Citation31,Citation126–128]. As illustrated in , this includes the formation of excitons where the electron- and the hole-component reside in different valleys of the Brillouin zone, separated by momentum w. For those intervalley excitons, the COM is then described by Q=ke+khw, and these excitons are also not straightforwardly accessible in an all-optical experiment.

2.4. Breaking the exciton in the photoemission process: energy- and momentum-conservation

In the photoemission process, the probe laser pulse breaks the exciton into its single-particle components. The former exciton’s electron is detected by the photoelectron analyzer and the former exciton’s hole remains in the sample (). In this process, the energy and the momentum of the total system need to be conserved. For excitons with a center-of-mass momentum Q=0, Weinelt et al.  [Citation36] formulated it as follows: Before and after the breakage of the exciton, the total energy of the system is given by Etot=E0+Eexc+ωprobe and Etot=E0Ehole+Eelec, respectively. Here, E0 is the ground state energy before the optical excitation and Ehole (Eelec) are energies of the single-particle holes (electrons) after the breakup of the exciton. The energy of the single-particle photoelectrons can then be expressed as

(3) Eelec=Ehole+Eexc+ωprobe.(3)

EquationEquation (3) directly shows that the energy- and the momentum-coordinate of the photoemitted electrons Eelec depends on the energy and the momenta of the single-particle hole that remains in the sample Ehole: The photoelectrons are detected one exciton energy Eexc above the valence bands whereas the single-particle hole remains in the sample. For the case that the exciton is composed of a single pair of valence and conduction bands, the result of EquationEquation (3) can be intuitively expressed in terms of the exciton binding energy Ebin (): At the cost of the exciton binding energy Ebin, single-particle photoelectrons originating from excitons are found below the conduction band minimum (i.e. EVBM+Eexc=ECBMEbin).

This connection between the electron- and the hole-component of the exciton in the photoemission process is now formulated by multiple groups on a different level of theory  [Citation35,Citation83,Citation129–133]. exemplarily shows the analysis of Rustagi and Kemper  [Citation35] for the case of an indirect semiconductor where the exciton’s electron and hole component reside in different valleys (w0). In the limit of a low exciton density where the excitons have no center-of-mass momentum (), the photoemission from the exciton is found one exciton energy above the valence band (respectively, one exciton binding energy below the conduction band). Importantly, the energy-momentum dispersion shows the hole-like dispersion of the valence band, as expected from energy- and momentum-conservation following EquationEquation (3) for Q = 0 excitons. Notably, the situation is different if excitons with a finite center-of-mass momentum contribute (). In this scenario, the pure valence band-like dispersion of the excitonic photoemission signal is not anymore observed. However, the photoemission signal is still found one exciton binding energy Ebin below the conduction band, facilitating the discrimination of photoelectrons being emitted from excitons or the conduction band.

Figure 3. The photoemission signature from excitons is affected by the joint electron-hole nature of the compound quasiparticle. (a) For Q = 0 excitons in TMDs, it was found that the dispersion of the hole state is directly imprinted on the trARPES data. (b) For excitons including a wider range of momenta Q (see insets), the trARPES signature is broadened and the hole-dispersion is no longer apparent. Figure reproduced from ref.  [Citation35]. Copyright 2018 by the American Physical Society.

Figure 3. The photoemission signature from excitons is affected by the joint electron-hole nature of the compound quasiparticle. (a) For Q = 0 excitons in TMDs, it was found that the dispersion of the hole state is directly imprinted on the trARPES data. (b) For excitons including a wider range of momenta Q (see insets), the trARPES signature is broadened and the hole-dispersion is no longer apparent. Figure reproduced from ref.  [Citation35]. Copyright 2018 by the American Physical Society.

2.5. Generalization to multi-orbital excitons

So far, excitons have been considered that are derived from a single pair of valence and conduction bands, such as sketched in . In this case, the connection between the exciton binding energy Ebin, the exciton energy Eexc and the single-particle valence and conduction band energies is straightforward (). However, as described by EquationEquation (1), an exciton can be composed of multiple orbital pairs, where the single-particle electrons and holes reside in different single-particle conduction and valence bands, respectively.

Hence, EquationEquation (1) can be understood as the coherent sum of multiple electron-hole pairs. It is known that such multi-orbital excitons strongly contribute to the optical properties of organic semiconductors  [Citation24,Citation56,Citation83,Citation134–136], but they can also contribute to the optical response of multilayer TMD systems, where interlayer hybridization effects lead to the formation of valence and conduction bands with bonding and anti-bonding character  [Citation124,Citation132,Citation137]. In such a complex scenario, it is not directly clear at which energy and momentum the photoelectrons originating from the excitons are expected. While time-dependent density functional theory methods provide a powerful and generally applicable approach to achieve this information  [Citation138], a more direct theoretical description would be beneficial, and was recently developed in two publications by Kern et al.  [Citation83] and Meneghini et al.  [Citation132].

Here, we first review the work of Kern et al.  [Citation83] and discuss the expected photoemission intensity for the case of multi-orbital excitons in an organic semiconductor. Kern et al. realized that the complex exciton wavefunction (1) can be split into four prototypical cases, where each case has a unique photoemission signature (). Case 1 is the simplest scenario, where the exciton is composed of a single valence band hole and a single conduction band electron. The excitons that have been discussed so far fit into this category. Beyond case 1, we find exciton wavefunctions that are composed of multiple elementary transitions between orbitals and bands. Case 2 specifies an exciton that is composed of two elementary transitions, namely from the second-highest occupied molecular orbital (HOMO-1) to the LUMO and from the HOMO to the second-lowest unoccupied molecular orbital (LUMO +1). This exciton can be considered as a coherent superposition of a HOMO-1 - LUMO electron-hole pair and a HOMO – LUMO + 1 electron-hole pair. Cases 3 and 4 are also composed of two elementary transitions; however, in these cases, the electron (case 3) or the hole (case 4) are uniquely occupying a single orbital. Of course, even more complicated wavefunctions are possible within EquationEquation (1), but these can be expressed as linear combinations or extensions of the cases 1–4.

Figure 4. The multi-orbital nature of excitons can be identified in the photoemission spectroscopy experiment. Kern et al. (a) and Meneghini et al. (b) found that multi-orbital excitons lead to more complex trARPES fingerprints, with one peak for each unique final-state binding energy of the hole. Panel a is reproduced from ref.  [Citation83]. Copyright 2023 by the American Physical Society. Panel b is reproduced with permission from ref.  [Citation132]. Copyright 2023 American Chemical Society.

Figure 4. The multi-orbital nature of excitons can be identified in the photoemission spectroscopy experiment. Kern et al. (a) and Meneghini et al. (b) found that multi-orbital excitons lead to more complex trARPES fingerprints, with one peak for each unique final-state binding energy of the hole. Panel a is reproduced from ref.  [Citation83]. Copyright 2023 by the American Physical Society. Panel b is reproduced with permission from ref.  [Citation132]. Copyright 2023 American Chemical Society.

The next step is to calculate the characteristic photoemission signature for these individual cases. At first, it is instructive to consider energy conservation in the photoemission process, i.e. EquationEquation (3), which implies that the kinetic energy of the observed photoelectron is related to the final-state energy Ehole of the remaining hole. For case 1 (and 4), this energy is well defined: as the exciton is derived from a single orbital, it follows that the hole remains in this state after the exciton is broken. For cases 2 and 3, however, multiple valence orbitals contribute to the exciton wavefunction, and the hole may remain in either of these orbitals after photoemission. Depending on the hole final state, the photoelectron then acquires more or less kinetic energy, leading to one peak in the photoelectron spectrum for each unique hole final-state energy Ehole (see bottom panel of ).

Kern et al. went beyond this first step and calculated the complete photoelectron spectrum for weakly-interacting systems in the plane-wave approximation, including the amplitude of the individual peaks in the photoelectron spectrum and, crucially, the momentum distribution. For an exciton with energy Eexc that is photoemitted by a photon with energy ωprobe, the angle-resolved photoelectron spectrum is proportional to

(4) I(Ekin,k)|Ak|2v|cXvc(m)F[χc](k)|2×δ(ωprobeEkinEv+Eexc).(4)

Here A is the vector potential of the incident light field, F the Fourier transform, k the photoelectron momentum, Ev the final-state binding energy of the hole in the vth valence orbital, and Ekin the energy of the photoemitted electron. This description directly reproduces the appearance of multiple peaks in the photoelectron spectrum for excitons with multi-orbital hole contributions.

It is interesting to observe the effect of the electron (conduction) orbitals on the momentum distribution: in case 1, the momentum distribution is directly determined by the momentum distribution F[χc](k) of the electron orbital. We emphasize that the hole (valence) orbital ϕv has no contribution in the photoelectron momentum distribution (cf. EquationEquation (4)). The same happens in cases 2 and 3: here, each peak in the photoelectron spectrum is linked to electron-hole pairs derived from a single conduction orbital, and thus the momentum distribution is given by that conduction orbital. In case 4, however, the photoelectron spectrum contains signal at the same kinetic energy coming from multiple conduction orbitals, (χ1 and χ2 in ). As described by EquationEquation (4), the photoelectron momentum distribution is then given by the coherent sum of the individual electron orbitals. Thus, trARPES is sensitive to unique interferences related to the coherent sum of electron-hole pairs that builds up a single exciton, which promises a unique insight into the correlated nature of excitons.

Meneghini et al.  [Citation132] derive a complementary picture of the exciton photoemission spectrum and predict a very similar phenomenon for the case of bilayer TMDs. In such systems, hybridization between the top and bottom TMD layers leads to a splitting of the valence band at Γ into two bands with bonding and anti-bonding character, respectively. These bands can host so-called hybrid excitons, in which the hole-part of the exciton wavefunction is derived from the hybridized bands at Γ. Due to the Coulomb interaction between electron and hole, the exciton’s hole is in a state that is described by a superposition of band contributions from two valence bands from the neighboring layers, and after the breakup of the exciton, the exciton’s hole can either remain in the band with bonding or with anti-bonding character. In this manner, very similar to case 3 in , the model predicts a two-peak photoemission structure, where the energy difference between both photoemission signatures is given by the energy splitting of the valence bands at the Γ valley, i.e. the energy position where the single-particle holes remain in the sample ().

3. Fingerprints of excitons in time-resolved momentum microscopy

The last few years have shown tremendous progress in the characterization of excitons in energy- and momentum-resolved photoemission spectroscopy. In this section, selected examples on how photoemission signatures of excitons could be identified in two-dimensional and organic semiconductors are provided. Thereby, we highlight how the momentum-resolved photoelectron detection scheme has facilitated the direct characterization of the energy landscape of bright and dark excitons.

3.1. Spectroscopy of intralayer, interlayer, and hybrid excitons in TMDs

Excitons with a large oscillator strength are routinely studied in all-optical spectroscopies. Since 2020, momentum microscopy experiments could strongly enhance the current understanding as they provided direct access to excitons with a vanishing oscillator strength. In this section, we review the recent experimental results that reported the observation of intralayer, interlayer, and hybrid excitons in exfoliated mono-  [Citation46–48,Citation52] and twisted bilayer  [Citation50,Citation51,Citation53–55] TMDs.

3.1.1. Momentum-direct and -indirect excitons in monolayer TMDs

Madéo et al.  [Citation46] () and Wallauer et al.  [Citation47] () have reported the observation of photoemission spectral weight originating from excitons in momentum microscopy experiments performed on exfoliated monolayer WSe2 and monolayer WS2, respectively. In their work, single-particle photoelectrons being emitted from excitons have been detected at energies above the valence band maximum at the K and the Σ valley of the hexagonal Brillouin zone, and were identified as signatures of the A-excitons and the dark Σ-excitons (; Σ valley also termed Q or Λ in literature). illustrates the involved excitons in an electron-hole picture for monolayer WSe2: The photoemission yield at the Σ valley originates from intralayer Σ-excitons, where the exciton’s electron- and hole-components reside at the Σ valley and the K valley, respectively. Importantly, as the exciton’s components reside in different valleys of the Brillouin zone separated by momentum w, those excitons cannot recombine in a radiative process and are thus termed optically dark (cf ). For the case of the spectral weight at the K valley, two different excitons can contribute (): The optical excitation leads to the formation of excitons, where the electron- and the hole-component reside in the same K (or Kʹ) valley, i.e. the exciton’s components have the same valley index. However, subsequent intervalley thermalization processes can lead to the formation of the energetically more favorable K-Kʹ excitons, where the electron- and the hole-component have different valley indices  [Citation33,Citation139]. In consequence, these excitons are again termed optically dark due to the finite w that separates the electron from the hole-component. It has to be noted that, as the energy difference between the K-Kʹ (Kʹ-K) excitons and the K-K (Kʹ-Kʹ) excitons is small, it is not straightforwardly possible to spectroscopically differentiate these excitons in the photoemission experiment. If one is interested in these excitons, it is necessary to repeat the photoemission experiments in a valley-selective approach using circular polarized pump pulses  [Citation52,Citation84], as reported by Kunin et al. for the case of monolayer WS2  [Citation52].

Figure 5. Energy- and momentum-resolved photoemission spectra collected from exfoliated a monolayer WSe2 and b monolayer WS2. (a) Photoemission spectral weight originating from the breakup of excitons is detected at the K and the Σ (Q) valley within the single-particle band gap. (b) Pump-probe delay dependent momentum-maps indicating the optical excitation of K excitons at the K valley and the subsequent formation of dark Σ excitons at the Σ valley. Panel a from ref.  [Citation46]. Reprinted with permission from AAAS. Panel b reprinted with permission from ref.  [Citation47]. Copyright 2021 American Chemical Society.

Figure 5. Energy- and momentum-resolved photoemission spectra collected from exfoliated a monolayer WSe2 and b monolayer WS2. (a) Photoemission spectral weight originating from the breakup of excitons is detected at the K and the Σ (Q) valley within the single-particle band gap. (b) Pump-probe delay dependent momentum-maps indicating the optical excitation of K excitons at the K valley and the subsequent formation of dark Σ excitons at the Σ valley. Panel a from ref.  [Citation46]. Reprinted with permission from AAAS. Panel b reprinted with permission from ref.  [Citation47]. Copyright 2021 American Chemical Society.

Figure 6. Energy- and momentum-resolved photoemission data collected on monolayer WSe2 (a) and twisted heterobilayer WSe2/MoS2 (b). a in the occupied regime, the electronic band structure of monolayer WSe2 is characterized by the spin-split valence bands at the K valley and the energetically lower lying valence band maxima at the Γ valley (horizontal arrow). Photoemission signatures of optically bright and dark intralayer excitons are detected at higher photoelectron energies (cf. EDCs taken in the colored regions-of-interest). b Interlayer hybridization leads to a renormalization of the valence band structure and the formation of a split valence band at the Γ valley (horizontal arrows). At higher energies, photoemission yield from intralayer K- (orange) and hybrid hΣ (grey) excitons, as well as interlayer ILX (red) is detected. c, d Electron-hole picture of the low energy exciton landscape of monolayer WSe2 (c) and heterobilayer WSe2/MoS2 (d). Panels adopted from ref.  [Citation51] under Creative Commons Attribution License 4.0 (CC BY).

Figure 6. Energy- and momentum-resolved photoemission data collected on monolayer WSe2 (a) and twisted heterobilayer WSe2/MoS2 (b). a in the occupied regime, the electronic band structure of monolayer WSe2 is characterized by the spin-split valence bands at the K valley and the energetically lower lying valence band maxima at the Γ valley (horizontal arrow). Photoemission signatures of optically bright and dark intralayer excitons are detected at higher photoelectron energies (cf. EDCs taken in the colored regions-of-interest). b Interlayer hybridization leads to a renormalization of the valence band structure and the formation of a split valence band at the Γ valley (horizontal arrows). At higher energies, photoemission yield from intralayer K- (orange) and hybrid hΣ (grey) excitons, as well as interlayer ILX (red) is detected. c, d Electron-hole picture of the low energy exciton landscape of monolayer WSe2 (c) and heterobilayer WSe2/MoS2 (d). Panels adopted from ref.  [Citation51] under Creative Commons Attribution License 4.0 (CC BY).

Based on the data shown in , Madéo et al.  [Citation46] and Wallauer et al.  [Citation47] had the major task to unambiguously discriminate photoelectrons that result from the breakup of excitons and those that are emitted from the conduction band (cf. energy diagram in ). For this, the authors reported several arguments: First, it could be shown that photoelectrons originating from excitons are indeed detected at energies below the conduction band minimum, consistent with expectations from EquationEquation (3). Thereby, the energy of the conduction band was estimated by trARPES experiments where the pump photon energy was tuned to energies larger than the single-particle band gap, as has also been carried out in a report by Dong et al.  [Citation49] on bulk WSe2. Second, excitation-energy-dependent experiments showed a clear enhancement of photoemission yield for the case that the excitation energy is tuned into resonance with the exciton energy  [Citation46]. In a similar manner, Kunin et al. report on the observation of B excitons in monolayer WS2 [Citation52].

3.1.2. Hybrid excitons in heterobilayer TMDs

So far, we have focused on monolayer TMDs, where the exciton’s electron- and hole-components necessarily reside in the same layer. Intriguingly, the situation can become much more interesting by stacking two semiconducting TMDs on-top of each other. For example, it is possible to fabricate an atomically thin p-n junction  [Citation140,Citation141], i.e. a heterostructure with a type-II band alignment  [Citation142]. In such a heterostructure, the electronic band structure and the concomitant energy landscape of excitons depend on the composition of the heterostructure and other parameters such as the twist-angle and the emergent moiré potential. For example, interlayer hybridization between two neighboring TMDs in a bilayer, as compared to a monolayer, leads to an energetic splitting of the hybridized bands and, in consequence, to a transition from a direct to an indirect semiconductor  [Citation123,Citation124]. In static ARPES experiments, the formation of hybrid electronic bands can be observed, e.g. when comparing the valence bands at the Γ valley for the case of a mono- and bilayer TMD  [Citation98,Citation143], as visualized for the case of monolayer WSe2 and heterobilayer WSe2/MoS2 in (black horizontal arrows).

In the case of the corresponding exciton energy landscape, interlayer hybridization will also lead to a renormalization of the exciton energies and the formation of so-called hybrid excitons, whose electron- and/or hole-components can be delocalized over both monolayers (inset )  [Citation132,Citation144–146]. Depending on the degree of delocalization of the exciton’s components across the interface, it is insightful to distinguish between three types of excitons, based on the trARPES data obtained from monolayer WSe2 and heterobilayer WSe2/MoS2  [Citation51]. show energy- and momentum-resolved photoemission data along the Γ-Σ-KW crystal direction collected at a pump-probe delay of 1 ps. At energies EEVBM<0 eV, the valence bands of WSe2 and MoS2 are detected. At higher energies, photoemission signatures from the (i) intralayer K-excitons (orange), (ii) interlayer excitons (red) and (iii) hybrid hΣ excitons (gray) are found. Before describing the photoemission signatures in detail, it is important to note that the trARPES experiment performed with linear polarized pump pulses cannot differentiate between excitons where the electron- or the hole-component reside at high-symmetry points with the same or different valley indices  [Citation52,Citation84].

First, the excitation of WSe2/MoS2 resonant to the optical band gap leads to the formation of (i) intralayer A1s-excitons that are fully localized in a single TMD layer and where the electron- and the hole-components reside in the K (or Kʹ) valleys. The dominant intralayer character of these excitons is due to the fact that interlayer hybridization is negligible at the K (Kʹ) valley of semiconducting TMDs  [Citation87,Citation90,Citation145,Citation146]. In consequence, photoemission spectral weight of the intralayer excitons is found one exciton energy Eexc above the WSe2 valence band maximum, where the exciton’s hole remains in the sample after the breakup of the exciton. Intriguingly, a consequence of the vanishing degree of interlayer hybridization can be directly seen in experiment when comparing the photoemission spectra collected from monolayer WSe2 and heterobilayer WSe2/MoS2: The exciton energies of the K-excitons are identical in the mono- and in the heterobilayer sample (cf. orange energy distribution curves (EDCs) in ). Second, after the excitation of the A1s-excitons, intra- and intervalley scattering events lead to the formation of (ii) interlayer excitons (ILX) via the exciton cascade KhΣILX  [Citation50,Citation51,Citation144]. For ILX, the exciton’s electron- and hole-components reside in the K valleys of MoS2 and WSe2, respectively. As interlayer hybridization is negligible at the K valleys, the exciton’s electron- and hole-components must be fully localized in one and the other single TMD layer, respectively. Third, in-between the excitation of (i) intralayer excitons and the build-up of (ii) ILX, (iii) hybrid excitons are formed, where the exciton’s electron and/or hole components are delocalized between both layers. As interlayer hybridization goes along with a reduction of the exciton energy, the photoemission yield at the Σ valley is found at lower photoelectron energies for the case of the WSe2/MoS2 heterobilayer as compared to the WSe2 monolayer (grey EDCs in ). In other words, the exciton energy Eexc of the intralayer Σ excitons in monolayer WSe2 is larger than the exciton energy of hybrid hΣ excitons in heterobilayer WSe2, consistent, e.g. with photoluminescence experiments reported in refs.  [Citation147–150] and first principle calculations  [Citation51,Citation113,Citation144]. Notably, these hybrid excitons are particularly important to mediate an efficient charge-separation process between the layers  [Citation50,Citation51,Citation53,Citation87,Citation132,Citation144,Citation146,Citation147,Citation151–153].

3.2. Momentum-fingerprints of excitons in photoemission spectroscopy

Until now, we have mainly focused on the spectral information in the photoemission experiment. The momentum-coordinate of the photoelectrons has only been of interest to differentiate between momentum-direct and -indirect excitons, i.e. to discriminate excitons with a different w. Next, we review the recent efforts to evaluate the momentum-resolved photoemission intensity from a specific exciton in order to show that not only information on the exciton’s electron component but also on the exciton’s hole can be extracted.

3.2.1. Hole-like energy-momentum dispersion of the exciton’s electron photoemission yield

The multi-dimensional data collection scheme of the momentum microscopy experiment facilitates the evaluation of the energy-momentum dispersion of the spectral weight that originates from the breakup of excitons. shows such data collected from monolayer WSe2 by Man et al.  [Citation48]. In this work, the authors have been able to identify a photoemission signature that shows a hole-like energy-momentum dispersion, as indicated by the magenta line in . The data of this work were collected at a pump-probe delay of 500 fs and at liquid-nitrogen temperature, so that a sufficiently small exciton temperature was reached for excitons with a center-of-mass momentum of mainly Q = 0. Notably, the authors found a photoemission signature with a hole-like energy-momentum-dispersion, which cannot result from photoelectrons that result from the conduction band and would thus mimic its electron-like energy-momentum dispersion  [Citation49]. In contrast, consistent with the theoretical predictions  [Citation35,Citation83,Citation129–133], Man et al. explain the hole-like energy-momentum dispersion to result from the breakup of Q = 0 excitons and the conservation of energy and momentum as single-particle photoelectrons are detected and single-particle holes remain in the sample.

Figure 7. Hole-like energy-momentum fingerprint of photoemission spectral weight originating from the break-up of intralayer excitons in monolayer WSe2. Panel reproduced from ref.  [Citation48] under Creative Commons Attribution License 4.0 (CC BY).

Figure 7. Hole-like energy-momentum fingerprint of photoemission spectral weight originating from the break-up of intralayer excitons in monolayer WSe2. Panel reproduced from ref.  [Citation48] under Creative Commons Attribution License 4.0 (CC BY).

3.2.2. Interlayer excitons in the presence of the moiré potential

The rotational misalignment of multiple two-dimensional material sheets can be used to tailor an emergent potential that manipulates the energy landscape of a TMD heterostructure in real-space, i.e. the so-called moiré superlattice. Schematically, this is visualized in , where two hexagonal lattices are slightly twisted with respect to each other, leading to a well-known interference pattern. Depending on the lattice constant mismatch and the twist-angle Θ, the moiré wavelength can be as large as 20 nm and the potential energy well as deep as 100 meV  [Citation154,Citation155]. This additional modulation of the energy landscape in real-space enables further tailoring of the material’s properties and can be used, for example, to create and tune correlated electronic and excitonic phases  [Citation156–159]. For the sake of this review, the most interesting point is that excitons can interact with the additional periodic potential energy landscape. Specifically, depending on the Bohr radius and the moiré wavelength, excitons can either be localized (confined) in a single moiré potential well or they can be delocalized over multiple moiré unit cells. Exciton confinement has been experimentally demonstrated in optical spectroscopies where a fine structure in the fluorescence spectra could be attributed to the formation of so-called moiré minibands  [Citation160–162]. Naturally, the renormalization of the energy landscape of excitons due to the formation of moiré minibands is then a function of the twist-angle Θ  [Citation160,Citation163], and, hence, it is highly desirable to extract such information in a momentum-resolved photoemission experiment. In the following, we aim to review how a hallmark of the moiré potential is imprinted onto the momentum-resolved interlayer exciton photoemission signal.

Figure 8. Moiré superlattice hallmark imprinted onto the momentum-resolved photoemission signal from interlayer excitons. (a) The rotational misalignment of two hexagonal lattices leads to the formation of an emergent moiré superstructure in real-space. (b) Translated to momentum-space, the moiré superlattice results from the rotational misalignment of the hexagonal Brillouin zones by a twist-angle Θ. (c) Momentum-momentum-resolved photoemission data originating from the break-up of interlayer excitons taken on WSe2/MoS2 heterostructures with twist angles of 2.8, 9.8, and 28.8 (from left-to-right). The momentum map is taken at the energy of the interlayer exciton (cf. red EDC in ). The top row shows the full momentum-range that is accessible in the photoemission experiment, the bottom row shows blow-ups of the black circled areas. With increasing twist-angle, the length of the moiré lattice vectors (black arrows) and the size of the moiré mini Brillouin zone (red hexagon) becomes larger. The 9.8 and 28.8 data is reproduced and adopted from ref.  [Citation50] (Copyright 2022, Nature Publishing Group) and ref.  [Citation54], respectively. The 2.8 data was taken in the Göttingen ultrafast momentum microscopy laboratory by D. Schmitt, J.P. Bange, and W. Bennecke from a sample prepared by A. AlMutairi and S. Hofmann and has not been published elsewhere.

Figure 8. Moiré superlattice hallmark imprinted onto the momentum-resolved photoemission signal from interlayer excitons. (a) The rotational misalignment of two hexagonal lattices leads to the formation of an emergent moiré superstructure in real-space. (b) Translated to momentum-space, the moiré superlattice results from the rotational misalignment of the hexagonal Brillouin zones by a twist-angle Θ. (c) Momentum-momentum-resolved photoemission data originating from the break-up of interlayer excitons taken on WSe2/MoS2 heterostructures with twist angles of 2.8∘, 9.8∘, and 28.8∘ (from left-to-right). The momentum map is taken at the energy of the interlayer exciton (cf. red EDC in Figure 6b). The top row shows the full momentum-range that is accessible in the photoemission experiment, the bottom row shows blow-ups of the black circled areas. With increasing twist-angle, the length of the moiré lattice vectors (black arrows) and the size of the moiré mini Brillouin zone (red hexagon) becomes larger. The 9.8∘ and 28.8∘ data is reproduced and adopted from ref.  [Citation50] (Copyright 2022, Nature Publishing Group) and ref.  [Citation54], respectively. The 2.8∘ data was taken in the Göttingen ultrafast momentum microscopy laboratory by D. Schmitt, J.P. Bange, and W. Bennecke from a sample prepared by A. AlMutairi and S. Hofmann and has not been published elsewhere.

To set the stage, shows the consequence of the real-space moiré periodicity in momentum-space. The real-space misalignment of the monolayers translates to a rotational misalignment of the hexagonal Brillouin zones of both layers by the twist-angle Θ. Naturally, this misalignment of the hexagonal Brillouin zones can then be visualized in static ARPES experiments by the misalignment of the valence bands and, potentially, a renormalization of the electronic band structure  [Citation98,Citation99,Citation164,Citation165]. Recently, we and others extended this discussion to the excitonic response of a 9.8 twisted WSe2/MoS2 heterostructure  [Citation50]. While the energy-momentum-resolved data obtained from this sample are already described above in shows the corresponding momentum-momentum-resolved visualization of the photoemission data (data taken at the energy of the photoelectron energy of the interlayer excitons (ILX)). Before discussing the data in detail, one has to recapitulate what might be expected for the photoemission signal originating from the ILX residing in the twisted heterostructure. The ILX’s electron- and hole-components reside in the KMo and KW valley of MoS2 and WSe2, respectively. In the description of the energy-momentum-resolved photoemission data from excitons shown in , a rule of thumb was that the single-particle photoelectrons are detected at the in-plane momenta where the exciton’s electron component resides in the electron-hole picture (cf. ), as first shown by Madéo et al. and Wallauer et al. for intralayer excitons in monolayer TMDs  [Citation46,Citation47]. In consequence, for the ILX, one would expect to observe photoemission yield at the KMo (and KMo) valleys. However, the momentum-resolved photoemission data shown in are more complex and show a three-fold structure. To understand this ILX momentum fingerprint, the moiré mini Brillouin zone (mBz) that is built-up by the rotationally misaligned reciprocal lattice vectors of WSe2 and MoS2 (red hexagons in in combination with an extended zone scheme (see refs.  [Citation166,Citation167]) needs to be considered  [Citation50]. In such an analysis, it is found that photoemission yield of the ILX is not only detected at the momentum of the exciton’s electron location (i.e. the KMo() valley) but also at those momenta accessible via Umklapp processes with reciprocal lattice vectors of the layer where the exciton’s hole resides. Hence, the photoemission momentum-structure of the ILX shows a hallmark of the moiré superlattice, and it contains direct information on the exciton’s electron- and hole-components. In reciprocal space, an increasing twist-angle Θ elongates the moiré lattice vectors (black arrows in ), leading to an increasing size of the mBz (red hexagon). Naturally, by Fourier analogy, this goes along with a decreasing size of the real-space moiré wavelength.

From the twist-angle-dependent data in , two conclusions can be drawn: First, trARPES is capable to experimentally access ILXs in a nearly 30 misaligned heterostructure, i.e. from a sample where the momentum offset w of the exciton’s electron- and hole-component is maximal and only leads to vanishing fluorescence signal  [Citation30]. Even more important, second, the momentum-resolved photoelectron detection scheme can be used to directly visualize the transition from a twist-angle regime where the formation of moiré minibands can be neglected (large Θ, i.e. 9.8 and 28.8 in ) to the regime where the twist-angle is sufficiently small such that Umklapp processes with the reciprocal moiré superlattice vectors leads to overlap between the excitonic bands and thus the formation of moiré mini bands (small Θ, i.e. 2.8 in ). For such small twist-angles Θ, one gets into a regime where excitons can be confined in a single moiré potential well  [Citation160,Citation163]. In this context, we want to highlight the work of Karni et al.  [Citation55], who reported momentum microscopy experiments on a 2.2 twisted WSe2/MoS2 heterostructure and could reconstruct the full exciton wavefunction that is confined in a single moiré potential well (see subsection 4.2).

3.3. Photoelectron spectrum of multi-orbital excitons

In the last part of this section, we review the recent experimental progress concerning the probing of multi-orbital excitons in momentum microscopy. As discussed in subsection 2.5 and shown in , such excitons are predicted to yield multi-peaked signals in trARPES, appearing at kinetic energies separated by the difference in binding energy of the different hole orbitals that contribute to the exciton.

Bennecke et al.  [Citation56] have reported a time-resolved momentum microscopy study of excitons in multilayer C60, where the experimental data are compared to theoretical exciton spectra calculated within the many-body interaction frameworks of the GW approach and the Bethe-Salpeter equation. Specifically, it is predicted that the low-energy excitons, which have been studied in trARPES before  [Citation168,Citation169], are derived from the (multiply degenerate) HOMO and LUMO levels, while higher lying excitons with an exciton energy of 3 eV and more are of increasing multi-orbital nature (). For example, at an exciton energy of 3.6 eV, a band of excitons is identified which shows clear multi-orbital nature where the exciton’s hole-component is spread over the HOMO and HOMO-1 orbitals (S4-exciton). In the pump-probe delay evolution of momentum-integrated energy distribution curves collected in the trARPES experiment, the double-peak structure could indeed be identified (, grey shading), as predicted for an exciton of type (ii) as shown in and for the case of the C60 molecule in . In this manner, these results provide an experimental confirmation of the predictions in refs.  [Citation83] and [Citation132], and, more generally, showcase how trARPES can experimentally disentangle the multi-orbital contributions to a single exciton band.

Figure 9. Multi-orbital excitons in the organic semiconductor C60. (a) A calculation in the combined many-body interaction frameworks of the GW approach and the Bethe-Salpeter equation predicts that multi-orbital excitons are prevalent in C60. (a) In particular, an exciton band around 3.6 eV excitation energy (‘S4’) is identified, in which both the HOMO and the HOMO-1 contribute significantly to the hole character of the exciton (case 2 in ). This leads to the prediction of two photoemission features at 3.6 eV above the HOMO and 3.6 eV above the HOMO-1 ( = 2.2 eV above the HOMO), respectively. (c) A challenge in the measurement of this double signature, however, is the short lifetime of the S4 exciton band. By performing momentum microscopy measurements on the sub-100-fs timescale at two distinct pump energies, it is possible to identify an enhancement of the photoemission yield that is only present during excitation of the S4 exciton band (gray shading in c, right panel). Figure adopted from ref [Citation56] under Creative Commons Attribution License 4.0 (CC BY).

Figure 9. Multi-orbital excitons in the organic semiconductor C60. (a) A calculation in the combined many-body interaction frameworks of the GW approach and the Bethe-Salpeter equation predicts that multi-orbital excitons are prevalent in C60. (a) In particular, an exciton band around 3.6 eV excitation energy (‘S4’) is identified, in which both the HOMO and the HOMO-1 contribute significantly to the hole character of the exciton (case 2 in Figure 4a). This leads to the prediction of two photoemission features at 3.6 eV above the HOMO and 3.6 eV above the HOMO-1 ( = 2.2 eV above the HOMO), respectively. (c) A challenge in the measurement of this double signature, however, is the short lifetime of the S4 exciton band. By performing momentum microscopy measurements on the sub-100-fs timescale at two distinct pump energies, it is possible to identify an enhancement of the photoemission yield that is only present during excitation of the S4 exciton band (gray shading in c, right panel). Figure adopted from ref [Citation56] under Creative Commons Attribution License 4.0 (CC BY).

At this point, it is important to note that it is perhaps a bit surprising that such multi-orbital excitons were not studied in detail before. The general theoretical framework of excitons allows such excitations [Citation134–136], and following the recent results [Citation56,Citation57,Citation83,Citation132] it is clear that such excitons are expected to arise in TMDs as well as in organic semiconductors. Although research on this topic is still in a very early stage, it is possible to identify two potential reasons why such multi-orbital excitons have not been observed before: First, the work on organic semiconductors suggests that multi-orbital excitons are more likely to occur at higher photon energies above 3 eV  [Citation56,Citation83]. At these energies, absorption of two photons from the pump pulse commonly suffices to photoemit an electron above the vacuum level, which renders two-color trARPES studies rather challenging. In fact, precisely this process limited the applicable pump fluence in the C60 study discussed above  [Citation56]. On the other hand, the work by Meneghini et al.  [Citation132] illustrates another experimental challenge in the detection of multi-orbital excitons: the relative strength of the photoemission features is determined by the relative occupation of the different hole states, meaning that in many cases, one feature will be significantly brighter than the other (compare intensity difference of the higher- and lower-energy-peak in ). The observation of multi-orbital excitons therefore requires a good signal-to-noise ratio in trARPES, especially if the splitting of the hole states is smaller than or comparable to the energy resolution of the experiment.

4. Photoemission orbital tomography of excitons

In this section, we switch the analysis of exciton photoemission from momentum-space to real-space, and discuss how the real-space exciton wavefunction can be reconstructed with Ångström-level resolution from momentum-resolved photoemission data. This is possible by means of photoemission orbital tomography (POT), a technique that was first developed for the imaging of molecular orbitals by static (kx,ky)-resolved ARPES  [Citation78]. POT is based upon the plane-wave model of photoemission, which states that the final-state of the photo-excited electrons can be well approximated by a plane wave. For two-dimensional systems such as TMDs and also flat organic semiconductors, such as pentacene, this leads to the conclusion that the two-dimensional photoelectron momentum distribution I(kx,ky) is proportional to the Fourier transform of the single-particle real-space molecular orbital F[ψ(rx,ry)] up to a smoothly-varying factor Ak, i.e.

(5) I(kx,ky)|F[ψ(rx,ry)]|2.(5)

While the accuracy of the plane wave model has been topic of discussion  [Citation170–173], it has been very successfully applied in the last years. For instance, the plane-wave model enables to calculate accurate predictions of the photoemission momentum distribution of occupied molecular orbitals that have been calculated by density functional theory. In POT studies, the comparison of these predictions to experimental data allows an exceptional insight into the (electronic) structure of well-ordered molecular thin films and metal-organic hybrid interfaces  [Citation174,Citation175]. Perhaps, the most intriguing application of the plane-wave model is, however, to image the molecular orbital from experimental ARPES data alone. Similar to X-ray coherent diffractive imaging  [Citation176], numerical phase retrieval algorithms can be used to determine the phase of the photoelectron distribution, and thereby enable to calculate an image of the molecular orbital by an inverse Fourier transform  [Citation78,Citation104,Citation177–179].

4.1. Time-resolved photoemission orbital tomography

Although photoemission orbital tomography was first developed as a static technique, the potential to employ POT in a time-resolved experiment has been realized early on, and it has been a long-standing goal of the scientific community. Similar to the measurement of exciton dynamics in exfoliated 2D materials, it was the development of momentum microscopy in combination with high-repetition rate EUV beamlines that enabled time-resolved photoemission orbital tomography (trPOT) to be realized. Time-resolved momentum microscopy for POT was first harnessed by Wallauer et al.  [Citation76], who realized that trPOT not only requires an efficient measurement scheme but also an excited state that lives long enough to provide sufficient signal on top of the strong photoemission signal from occupied molecular orbitals. Wallauer et al. created such a long-lived state by decoupling the organic adsorbate layer (here perylene-3,4,9,10-tetracarboxylic dianhydride, PTCDA) from the metallic Cu(001) surface using a thin oxide layer.

As in the case of any seminal work in a new field of research, the work by Wallauer et al. also raises some important questions. For example, when applied to excitonic states, does trPOT have the potential to shed light on the dynamics and structure of the Coulomb-correlated electron-hole pairs? Given the two-particle nature of the excitons, can trPOT access the real-space properties of both the electron and the hole? Since ground-state density functional theory does not give access to the exciton spectrum, what level of theoretical description is necessary to access real-space properties in trPOT? In subsection 2.5, we have already seen that new theoretical frameworks are currently being developed that connect the real-space exciton wavefunction to the momentum-resolved photoemission experiment. In the following sections, we will review how this framework has recently been exploited to access the real-space properties of excitons in TMDs and organic semiconductors alike. Before going into detail, it has to be noted that the reviewed theoretical framework (EquationEquation (4)) is only valid for excitons consisting of a single electron and hole. The situation is much more complicated when composite excitons such as biexcitons or trions are considered  [Citation180–182], as in this case the left-behind state after photoemission can be excitonic in itself  [Citation57].

4.2. Two-dimensional semiconductors: reconstruction of the real-space wavefunction of excitons

In section 3, we have reviewed the energy- and momentum-resolved photoemission spectra that are collected after the breakup of excitons. Among these, the in-plane momentum-resolved photoemission intensity contains information on the electron-contribution to the exciton wavefunction, and, moreover, indirect information on the hole-component can be extracted from the momentum-fingerprint. In this section, we review how recent publications used such data and the framework of orbital tomography to reconstruct the real-space distribution of the electron-contribution to the exciton wavefunction. Since 2021, this concept has been applied to intralayer excitons in monolayer WSe2  [Citation48], intralayer excitons in bulk WSe2  [Citation49], and intra- and interlayer excitons in twisted heterobilayer WSe2/MoS2  [Citation50,Citation55].

show the real-space probability density to find the electron-contribution to intralayer excitons if the hole-component is fixed at the center of the coordinate system. Such data are obtained by evaluating the Fourier transform (EquationEquation (5)) that connects the real-space wavefunction Ψ(rx,ry) (main panels ) with the momentum-resolved photoemission intensity I(kx,ky) (inset ) within the plane wave approximation and the assumption of a constant phase profile  [Citation48,Citation50,Citation78,Citation94]. With the data at hand, it has then been possible to evaluate the exciton Bohr radius that was found to be in the order of 1–2 nm, and, therefore, extends over multiple lattice sites of the TMD (see hexagonal overlays in ).

Figure 10. Reconstruction of the real-space resolved exciton wavefunction from momentum-resolved photoemission data. (a–c) Application of the framework of orbital tomography to reconstruct the electron-contribution to the exciton wavefunction of intralayer excitons in bulk WSe2 (a), monolayer WSe2 (b), and heterobilayer WSe2/MoS2 (c). The reported Bohr radii are 1.72 nm, 1.4 nm, and 1.1±0.1 nm, respectively. (d) Calculated real-space envelope of the interlayer exciton wavefunction in the center-of-mass coordinate on 2.2 twisted WSe2/MoS2. (e) Analysis of the exciton Bohr radius of intralayer and interlayer excitons in WSe2/MoS2. (f) Reconstruction of the real-space distribution of the electron-contribution to the exciton wavefunction of interlayer excitons in 9.8 twisted WSe2/MoS2. Panel a is adopted from ref.  [Citation49] under Creative Commons Attribution License 4.0 (CC BY). Panel b is adopted from ref.  [Citation48] under Creative Commons Attribution License 4.0 (CC BY). Panels c,e,f are adopted from ref.  [Citation50] (Copyright by Springer Nature). Panel d is adopted from ref.  [Citation55] (Copyright by Springer Nature).

Figure 10. Reconstruction of the real-space resolved exciton wavefunction from momentum-resolved photoemission data. (a–c) Application of the framework of orbital tomography to reconstruct the electron-contribution to the exciton wavefunction of intralayer excitons in bulk WSe2 (a), monolayer WSe2 (b), and heterobilayer WSe2/MoS2 (c). The reported Bohr radii are 1.72 nm, 1.4 nm, and 1.1±0.1 nm, respectively. (d) Calculated real-space envelope of the interlayer exciton wavefunction in the center-of-mass coordinate on 2.2∘ twisted WSe2/MoS2. (e) Analysis of the exciton Bohr radius of intralayer and interlayer excitons in WSe2/MoS2. (f) Reconstruction of the real-space distribution of the electron-contribution to the exciton wavefunction of interlayer excitons in 9.8∘ twisted WSe2/MoS2. Panel a is adopted from ref.  [Citation49] under Creative Commons Attribution License 4.0 (CC BY). Panel b is adopted from ref.  [Citation48] under Creative Commons Attribution License 4.0 (CC BY). Panels c,e,f are adopted from ref.  [Citation50] (Copyright by Springer Nature). Panel d is adopted from ref.  [Citation55] (Copyright by Springer Nature).

Having reconstructed the electron-component to the wavefunction of intralayer excitons, in the next step, Schmitt et al.  [Citation50] and Karni et al.  [Citation55] focused on interlayer excitons in WSe2/MoS2 and the question of how far the exciton wavefunction is delocalized over multiple moiré unit cells or confined in a single moiré potential well. Schmitt et al.  [Citation50] focused on a 9.8 twisted heterostructure and found that the exciton Bohr radius is indeed – as expected from photoluminescence experiments and theory  [Citation160,Citation163] - larger than the size of a moiré unit cell (). Karni et al.  [Citation55] performed a similar experiment on a sample with a twist angle of 2.2 and found that the exciton wavefunction is confined in a single moiré potential well, in agreement with GW plus Bethe-Salpeter calculations (). Notably, in the work of Karni et al., the authors found a very distinctive intensity loss in the ARPES signature of the valence band that they attributed to the hole-component of the interlayer exciton. In this way, they succeeded in reconstructing both the electron- and the hole-coordinate of the exciton wavefunction (cf. EquationEquation (1)).

4.3. Organic semiconductors: probing the localization and delocalization of excitons in real-space

Finally, we review the analysis of real-space exciton wavefunctions in organic semiconductors. Here, a crucial difference with the TMD semiconductors should be noted: whereas TMDs are periodic systems in which the excitons derive their wavefunction from the Bloch states of the valence and conduction bands, the electrons in organic semiconductors are more localized. Consequently, excitons in TMDs are Wannier excitons [Citation183], while excitons in organic semiconductors are conventionally classified into Frenkel and charge-transfer (CT) excitons [Citation184–186]. Frenkel excitons have their electron and hole components in the same molecule, while the electron and hole components in a CT exciton are separated over multiple molecules.

The potential of momentum microscopy to shine light on the charge-transfer or Frenkel nature of an exciton is well illustrated by the recent research on crystalline C60  [Citation56]. Whereas CT excitons are commonly studied in multi-component systems such as hybrid organic/inorganic interfaces [Citation187–189], an indication for a CT exciton was recently found in momentum-integrated time-resolved photoemission spectroscopy experiments on C60  [Citation168,Citation169]. However, as these initial works on CT excitons in C60 were performed using a hemispherical analyzer and no full momentum distributions were available at EUV energies, it was not yet possible to apply the framework of POT to access the real-space wavefunction of the excitons. This task could only be later realized in a study based on momentum microscopy experiments  [Citation56]. Indeed, the ab-initio many-body interaction GW+BSE calculations in this work, plugged into EquationEquation (4) and combined with the full EUV momentum microscopy data, provide the opportunity to analyze the real-space wavefunction of these excitons. Such a comparison, and exemplary visualizations of the electron-hole wavefunction, are shown in . While this comparison was able to confirm the CT nature of the 2.8 eV exciton in multilayer C60, the results are still limited by the computational expense of a full GW and BSE calculation, which in this case was performed for a unit cell of two C60 molecules. It is to be expected that the field of time-resolved exciton POT (aka time-resolved”exPOT”  [Citation56,Citation83]) will benefit significantly from future improvements in the computational power and methods.

Figure 11. Real-space analysis of the exciton wavefunction in organic semiconductors. (a) in ref.  [Citation56], the authors directly compare the momentum microscopy data of a C60 multilayer to ab-initio calculations based on the GW+BSE framework and EquationEquation (4). Whereas the momentum fingerprints for the lower-energy S1 and S2 exciton bands are accurately predicted, the theory fails to fully explain the differences in the S3 momentum fingerprints. These limitations are attributed to the limited size of the unit cell in the calculations. Nevertheless, the comparison of experiment and theory supports a CT-nature for the S3 exciton band. (b) in pentacene crystals, the optically bright singlet exciton at 1.81 eV decays by singlet fission into a pair of triplet excitons. Neef et al.  [Citation57] exploit the Fourier relationship between orbital and trARPES data to probe the exciton localization. They observe that the optically excited singlet (S, top-left) shows well-defined peaks in momentum, while the bitriplet (TT, top-right) is spread out, i.e. The excited state transitions from a delocalized to a localized state. Panel a is adapted from ref.  [Citation56] and panel b is adapted from ref.  [Citation57] under Creative Commons Attribution License 4.0 (CC BY).

Figure 11. Real-space analysis of the exciton wavefunction in organic semiconductors. (a) in ref.  [Citation56], the authors directly compare the momentum microscopy data of a C60 multilayer to ab-initio calculations based on the GW+BSE framework and EquationEquation (4)(4) I(Ekin,k)∝|Ak|2∑v|∑cXvc(m)F[χc](k)|2×δ(ℏωprobe−Ekin−Ev+Eexc).(4) . Whereas the momentum fingerprints for the lower-energy S1 and S2 exciton bands are accurately predicted, the theory fails to fully explain the differences in the S3 momentum fingerprints. These limitations are attributed to the limited size of the unit cell in the calculations. Nevertheless, the comparison of experiment and theory supports a CT-nature for the S3 exciton band. (b) in pentacene crystals, the optically bright singlet exciton at 1.81 eV decays by singlet fission into a pair of triplet excitons. Neef et al.  [Citation57] exploit the Fourier relationship between orbital and trARPES data to probe the exciton localization. They observe that the optically excited singlet (S, top-left) shows well-defined peaks in momentum, while the bitriplet (TT, top-right) is spread out, i.e. The excited state transitions from a delocalized to a localized state. Panel a is adapted from ref.  [Citation56] and panel b is adapted from ref.  [Citation57] under Creative Commons Attribution License 4.0 (CC BY).

TrPOT can also provide highly valuable insight even in the absence of a full theoretical description of the exciton wavefunction, as was shown in a recent study by Neef et al.  [Citation57]. Turning their attention to exciton dynamics in pentacene crystals, Neef et al. applied trPOT to study singlet-fission in this material  [Citation190]. Here, the momentum-resolved detection scheme provided two crucial benefits: first, the different excitonic states, particularly the optically excited singlet state and the intermediate bitriplet exciton, each have their own unique momentum finger print. By using these fingerprints, Neef et al. are able to decompose the measurement signal, and thus measure state-resolved dynamics where a decomposition purely based upon the photoelectron kinetic energy would not be possible  [Citation39]. Second, the authors also make use of the Fourier relationship between measurement and wavefunction that is inherent to (tr)POT: as shown in , the observed momentum fingerprints show distinct differences, with the optically excited singlet exciton displaying more pronounced peaks and the bitriplet exciton showing a more slowly-varying amplitude. From these data, Neef et al. elucidate that the (electron-contribution to the) singlet exciton in pentacene is delocalized, while the (electron-contributions to the) bitriplet are localized to a single molecule. These results show that the framework of trPOT can provide detailed spatial information on the exciton wavefunction, even for complicated cases such as the bitriplet biexciton, where no general theoretical framework yet exists to predict the momentum-resolved photoemission signature.

5. Summary

We have reviewed the recent efforts to make use of state-of-the-art time-resolved momentum microscopy to study excitons in two-dimensional and organic semiconductors in real- and momentum-space. By evaluating the kinetic energy and the momentum of single-particle photoelectrons originating from excitons, it is possible to quantify the exciton energy Eexc, the exciton binding energy Ebin, and also information on the momentum-coordinate of the exciton wavefunction. On the one side, this facilitates direct access to the long-sought after dark exciton energy landscape in two-dimensional TMDs. On the other side, this approach enables the measurement of very distinct momentum-fingerprints of excitons that could be related to the emergent moiré potential. Moreover, building on the momentum-resolved detection scheme, it has become possible to reconstruct the real-space exciton wavefunction within the framework of orbital tomography, providing thus access to the real-space wavefunction of an excited state.

6. Outlook

In the last section, we aim to provide a brief outlook on possible future research directions by discussing further developments of momentum microscopy with regard to its time-, energy-, real-space, and spin-resolution. Likewise, we briefly discuss the potential impact of higher probe photon energies and more advanced driving schemes with pump laser pulses in the mid-infrared (MIR) to terahertz (THz) regime.

Energy-time uncertainty restricts the momentum microscopy experiments to either address femto- to attosecond dynamics or probe optical excitations in a high-resolution spectroscopy. Notably, future efforts can address both these regimes by optimizing the high-harmonic generation process for time- or energy-resolution  [Citation13,Citation191,Citation192]. In this way, the so far developed attosecond trARPES methodology  [Citation193–197] can be extended to momentum microscopy experiments, thus providing direct access to attosecond phenomena with (kx,ky)-momentum resolution. On the other hand, the generation of energetically narrow EUV pulses with a spectral width < 20 meV  [Citation191] has the potential to, for example, resolve not only the lowest energy A1s excitons but also the Rydberg-like series of higher energy excitons  [Citation23,Citation62]. Similarly, it might be possible to resolve energetic fingerprints of flat moiré minibands  [Citation160–162] and the dynamics therein  [Citation198].

Additional new insights can be expected from the combination of momentum microscopy with more advanced excitation schemes (pump photons) and higher-energy probe photons. For the latter, first experiments at free-electron lasers have demonstrated the potential of using > 100 eV probe photons to simultaneously address the dynamics of core- and valence levels after the excitation of excitons  [Citation199] and even the application of ultrafast photoelectron diffraction  [Citation200]. Likewise, tuneable and higher-energy probe photons will allow to extend three-dimensional orbital tomography  [Citation201,Citation202] into the time-domain in order to probe the full wavefunction of excited states in two-dimensional and organic semiconductors. Complementary, on the side of the pump laser pulses, it can be envisioned that establishing THz driving schemes  [Citation203,Citation204] in the momentum microscope will facilitate the study of lightwave electronics  [Citation205] with full (kx,ky)-momentum resolution. Moreover, the combination of momentum microscopy with NIR/MIR excitation  [Citation206,Citation207] or even coherent multi-pump experiments  [Citation82,Citation208–215] can provide new avenues to study coherent light-matter interaction and coherent material control  [Citation216,Citation217] with full momentum resolution.

In addition to the study of excitons in pristine two-dimensional and organic semiconductors, future research can address hybrid heterostructures that are assembled of organic and inorganic materials  [Citation45,Citation188,Citation218–223], lateral two-dimensional heterostructures  [Citation224,Citation225], nanostructured heterostructures  [Citation226,Citation227], electrostatically contacted samples  [Citation101,Citation102,Citation228–230], and magnetic two-dimensional materials  [Citation231,Citation232]. Specifically for the case of a hybrid organic/inorganic heterostructures, we envision that excitons localized in the TMD or the organic layer can be straightforwardly discriminated based on their unique photoemission momentum-fingerprint (cf. versus ). In this way, it might become possible to elaborate on the contribution of dark and hybrid excitons to the charge-transfer dynamics across the hybrid interface. The combination of the time-resolved momentum microscopy experiment with dark-field imaging techniques  [Citation110,Citation111,Citation233] might facilitate the measurement of ultrafast exciton dynamics with ≈ 100 nm spatial-resolution  [Citation54]. Notably, such dynamical processes on the nano-femto scale are then not only accessible for optically bright but also for optically dark excitons  [Citation54]. Moreover, following on the pioneering, but static, nanoARPES experiments on electrostatically contacted samples  [Citation101,Citation102,Citation228–230], the implementation of such samples into the momentum microscope will allow to study the non-equilibrium response of, for example, optically bright and dark trions  [Citation89,Citation234] and even correlated phases  [Citation159,Citation235–238] with energy and (kx,ky)-momentum-resolution. Finally, the extension of momentum microscopy with a spin-filter  [Citation65,Citation239,Citation240] will facilitate to perform the full photoemission experiment, and thereby, for example, the study of exciton-magnon coupling  [Citation231,Citation232].

Acknowledgments

At this point, we would like to thank all colleagues that contributed to the research reviewed in the present article. First, we would like to thank the members of the Göttingen photoemission laboratory (in alphabetical order): Jan Philipp Bange, Wiebke Bennecke, Marius Keunecke, Mattis Langendorf, Marco Merboldt, Christina Möller, Hannah Strauch, Paul Werner, David Schmitt, Daniel Steil, Sabine Steil, and Bent van Wingerden. Second, we would like to thank Ermin Malic (University of Marburg), his team members Samuel Brem and Giuseppe Meneghini, as well as Peter Puschnig (University of Graz) and his team members Christian S. Kern and Andreas Windischbacher for the fruitful experiment-theory collaborations. And third, we would like to thank Stephan Hofmann (University of Cambridge) and his team members AbdulAziz AlMutairi and Irina Chirca, Thomas Weitz (University of Göttingen) and his team members Jonas F. Pöhls and Anna Seiler, and Martin Aeschlimann, Benjamin Stadtmüller and their team member Ralf Hemm for the fruitful collaboration on sample fabrication and characterization. Finally, we thank the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) who funded this research via 432680300/SFB 1456 (project B01) and 217133147/SFB 1073 (projects B07 and B10).

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Hüfner S. Photolectron spectroscopy – principles and applications. 3rd ed. Berlin: Springer; 2003.
  • Damascelli A, Hussain Z, Shen Z-X. Angle-resolved photoemission studies of the cuprate superconductors. Rev Mod Phys. 2003;75:473–541. doi: 10.1103/RevModPhys.75.473
  • Bovensiepen U, Petek H, Wolf M. Dynamics at solid state surfaces and interfaces, 1st ed vol. 2: fundamentals. Weinheim: Wiley-VCH Verlag Gmbh & Co.; 2012.
  • Sobota JA, He Y, Shen Z-X. Angle-resolved photoemission studies of quantum materials. Rev Mod Phys. 2021;93:025006. doi: 10.1103/RevModPhys.93.025006
  • Boschini F, Zonno M, Damascelli A. Time-resolved ARPES studies of quantum materials. Rev Mod Phys. 2024;96:015003. doi: 10.1103/RevModPhys.96.015003
  • Echenique PM, Berndt R, Chulkov EV, et al. Decay of electronic excitations at metal surfaces. Surf Sci Rep. 2004;52:219–317. doi: 10.1016/j.surfrep.2004.02.002
  • Bauer M, Marienfeld A, Aeschlimann M. Hot electron lifetimes in metals probed by time-resolved two-photon photoemission. Prog Surf Sci. 2015;90:319–376. doi: 10.1016/j.progsurf.2015.05.001
  • Petek H, Li A, Li X, et al. Plasmonic decay into hot electrons in silver. Prog Surf Sci. 2023;98:100707. doi: 10.1016/j.progsurf.2023.100707
  • Zhu XY. Electronic structure and electron dynamics at molecule-metal interfaces: implications for molecule-based electronics. Surf Sci Rep. 2004;56:1–83. doi: 10.1016/j.surfrep.2004.09.002
  • Johannsen JC, Ulstrup S, Cilento F, et al. Direct view of hot carrier dynamics in graphene. Phys Rev Lett. 2013;111:027403. doi: 10.1103/PhysRevLett.111.027403
  • Gierz I, Petersen JC, Mitrano M, et al. Snapshots of non-equilibrium Dirac carrier distributions in graphene. Nat Mater. 2013;12:1119–1124. doi: 10.1038/nmat3757
  • Na MX, Mills AK, Boschini F, et al. Direct determination of mode-projected electron-phonon coupling in the time domain. Science. 2019;366:1231–1236. doi: 10.1126/science.aaw1662
  • Düvel M, Merboldt M, Bange JP, et al. Far-from-equilibrium electron–phonon interactions in optically excited graphene. Nano Lett. 2022;22:4897–4904. doi: 10.1021/acs.nanolett.2c01325
  • Sobota JA, Yang S, Analytis JG, et al. Ultrafast optical excitation of a persistent surface-state population in the topological insulator Bi2Se3. Phys Rev Lett. 2012;108:117403. doi: 10.1103/PhysRevLett.108.117403
  • Niesner D, Otto S, Hermann V, et al. Bulk and surface electron dynamics in a p-type topological insulator SnSb2Te4. Phys Rev B. 2014;89:081404. doi: 10.1103/PhysRevB.89.081404
  • Reimann J, Güdde J, Kuroda K, et al. Spectroscopy and dynamics of unoccupied electronic states of the topological insulators Sb2Te3 and Sb2Te2S. Phys Rev B. 2014;90:081106. doi: 10.1103/PhysRevB.90.081106
  • Mahmood F, Chan C-K, Alpichshev Z, et al. Selective scattering between Floquet–Bloch and Volkov states in a topological insulator. Nat Phys. 2016;12:306–310. doi: 10.1038/nphys3609
  • Schmitt F, Kirchmann PS, Bovensiepen U, et al. Transient electronic structure and melting of a charge density wave in TbTe3. Science. 2008;321:1649–1652. doi: 10.1126/science.1160778
  • Rohwer T, Hellmann S, Wiesenmayer M, et al. Collapse of long-range charge order tracked by time-resolved photoemission at high momenta. Nature. 2011;471:490–493. doi: 10.1038/nature09829
  • Haug H, Koch SW. Quantum theory of the optical and electronic properties of semiconductors. Singapur: World Scientific Publishing Company; 2009.
  • Kira M, Koch SW. Semiconductor quantum optics. Cambridge: Cambridge University Press; 2011.
  • Huber R, Tauser F, Brodschelm A, et al. How many-particle interactions develop after ultrafast excitation of an electron–hole plasma. Nature. 2001;414:286–289. doi: 10.1038/35104522
  • Wang G, Chernikov A, Glazov MM, et al. Colloquium: excitons in atomically thin transition metal dichalcogenides. Rev Mod Phys. 2018;90:021001. doi: 10.1103/RevModPhys.90.021001
  • Valencia AM, Bischof D, Anhäuser S, et al. Excitons in organic materials: revisiting old concepts with new insights. Electron Struct. 2023;5:033003. doi: 10.1088/2516-1075/acf2d4
  • Kaindl RA, Carnahan MA, Hagele D, et al. Ultrafast terahertz probes of transient conducting and insulating phases in an electron–hole gas. Nature. 2003;423:734–738. doi: 10.1038/nature01676
  • Poellmann C, Steinleitner P, Leierseder U, et al. Resonant internal quantum transitions and femtosecond radiative decay of excitons in monolayer WSe2. Nat Mater. 2015;14:889–893. doi: 10.1038/nmat4356
  • Merkl P, Mooshammer F, Steinleitner P, et al. Ultrafast transition between exciton phases in van der waals heterostructures. Nat Mater. 2019;18:691–696. doi: 10.1038/s41563-019-0337-0
  • Kira M, Jahnke F, Hoyer W, et al. Quantum theory of spontaneous emission and coherent effects in semiconductor microstructures. Prog Quantum Electron. 1999;23:189–279. doi: 10.1016/S0079-6727(99)00008-7
  • Koch SW, Kira M, Khitrova G, et al. Semiconductor excitons in new light. Nat Mater. 2006;5:523–531. doi: 10.1038/nmat1658
  • Nayak PK, Horbatenko Y, Ahn S, et al. Probing evolution of twist-angle-dependent interlayer excitons in MoSe2/WSe2 van der waals heterostructures. ACS Nano. 2017;11:4041–4050. doi: 10.1021/acsnano.7b00640
  • Malic E, Selig M, Feierabend M, et al. Dark excitons in transition metal dichalcogenides. Phys Rev Mater. 2018;2:014002. doi: 10.1103/PhysRevMaterials.2.014002
  • Wu F, Lovorn T, MacDonald AH. Theory of optical absorption by interlayer excitons in transition metal dichalcogenide heterobilayers. Phys Rev B. 2018;97:035306. doi: 10.1103/PhysRevB.97.035306
  • Lindlau J, Selig M, Neumann A, et al. The role of momentum-dark excitons in the elementary optical response of bilayer WSe2. Nat Commun. 2018;9:2586. doi: 10.1038/s41467-018-04877-3
  • Brem S, Ekman A, Christiansen D, et al. Phonon-assisted photoluminescence from indirect excitons in monolayers of transition-metal dichalcogenides. Nano Lett. 2020;20:2849–2856. doi: 10.1021/acs.nanolett.0c00633
  • Rustagi A, Kemper AF. Photoemission signature of excitons. Phys Rev B. 2018;97:235310. doi: 10.1103/PhysRevB.97.235310
  • Weinelt M, Kutschera M, Fauster T, et al. Dynamics of exciton formation at the Si(100)c (4 × 2) surface. Phys Rev Lett. 2004;92:126801. doi: 10.1103/PhysRevLett.92.126801
  • Varene E, Bogner L, Bronner C, et al. Ultrafast exciton population, relaxation, and decay dynamics in thin oligothiophene films. Phys Rev Lett. 2012;109:207601. doi: 10.1103/PhysRevLett.109.207601
  • Deinert JC, Wegkamp D, Meyer M, et al. Ultrafast exciton formation at the ZnO (101¯0) surface. Phys Rev Lett. 2014;113:057602. doi: 10.1103/PhysRevLett.113.057602
  • Chan W-L, Ligges M, Jailaubekov A, et al. Observing the multiexciton state in singlet fission and ensuing ultrafast multielectron transfer. Science. 2011;334:1541–1545. doi: 10.1126/science.1213986
  • Cui X, Wang C, Argondizzo A, et al. Transient excitons at metal surfaces. Nat Phys. 2014;10:505–509. doi: 10.1038/nphys2981
  • Tanimura H, Tanimura K, van Loosdrecht PHM. Dynamics of incoherent exciton formation in Cu2O: time- and angle-resolved photoemission spectroscopy. Phys Rev B. 2019;100:115204. doi: 10.1103/PhysRevB.100.115204
  • Tanimura H, Tanimura K, Kanasaki J. Surface exciton formation on InP(110)− (1×1) studied by time- and angle-resolved photoemission spectroscopy. Phys Rev B. 2023;107:075304. doi: 10.1103/PhysRevB.107.075304
  • Zhu XY. Photoemission from excitons in organic semiconductors. J Electron Spectrosc Relat Phenom. 2015;204:75–79. doi: 10.1016/j.elspec.2015.02.007
  • Mori R, Ciocys S, Takasan K, et al. Spin-polarized spatially indirect excitons in a topological insulator. Nature. 2023;614:249–255. doi: 10.1038/s41586-022-05567-3
  • Gierster L, Turkina O, Deinert J-C, et al. How hybrid excitons suppress charge separation: ultrafast, but delayed. arXiv preprint arXiv:2307.05257. 2023.
  • Madéo J, Man MKL, Sahoo C, et al. Directly visualizing the momentum-forbidden dark excitons and their dynamics in atomically thin semiconductors. Science. 2020;370:1199–1204. doi: 10.1126/science.aba1029
  • Wallauer R, Perea-Causin R, Münster L, et al. Momentum-resolved observation of exciton formation dynamics in monolayer WS2. Nano Lett. 2021;21:5867–5873. doi: 10.1021/acs.nanolett.1c01839
  • Man MKL, Madéo J, Sahoo C, et al. Experimental measurement of the intrinsic excitonic wave function. Sci Adv. 2021;7:eabg0192. doi: 10.1126/sciadv.abg0192
  • Dong S, Puppin M, Pincelli T, et al. Direct measurement of key exciton properties: energy, dynamics, and spatial distribution of the wave function. Nat Sci. 2021;1:e10010. doi: 10.1002/ntls.10010
  • Schmitt D, Bange JP, Bennecke W, et al. Formation of moiré interlayer excitons in space and time. Nature. 2022;608:499–503. doi: 10.1038/s41586-022-04977-7
  • Bange JP, Werner P, Schmitt D, et al. Ultrafast dynamics of bright and dark excitons in monolayer WSe2 and heterobilayer WSe2/MoS2. 2D Mater. 2023;10:035039. doi: 10.1088/2053-1583/ace067
  • Kunin A, Chernov S, Bakalis J, et al. Momentum-resolved exciton coupling and valley polarization dynamics in monolayer WS2. Phys Rev Lett. 2023;130:046202. doi: 10.1103/PhysRevLett.130.046202
  • Bange JP, Schmitt D, Bennecke W, et al. Probing electron-hole coulomb correlations in the exciton landscape of a twisted semiconductor heterostructure. Sci Adv. 2024;10:eadi1323. doi: 10.1126/sciadv.adi1323
  • Schmitt D, Bange JP, Bennecke W, et al. Ultrafast nano-imaging of dark excitons. arXiv:2305.18908 [cond-mat.mes-hall]. 2023.
  • Karni O, Barré E, Pareek V, et al. Structure of the moiré exciton captured by imaging its electron and hole. Nature. 2022;603:247–252. doi: 10.1038/s41586-021-04360-y
  • Bennecke W, Windischbacher A, Schmitt D, et al. Disentangling the multiorbital contributions of excitons by photoemission exciton tomography. Nat Commun. 2024;15:1804. doi: 10.1038/s41467-024-45973-x
  • Neef A, Beaulieu S, Hammer S, et al. Orbital-resolved observation of singlet fission. Nature. 2023;616:275–279. doi: 10.1038/s41586-023-05814-1
  • Beaulieu S, Dong S, Christiansson V, et al. Berry curvature signatures in chiroptical excitonic transitions. arXiv preprint arXiv:2308.09634. 2023.
  • Karni O, Esin I, Dani KM. Through the lens of a momentum microscope: viewing light-induced quantum phenomena in 2D materials. Adv Mater. 2023;35:2204120. doi: 10.1002/adma.202204120
  • Liu F. Time- and angle-resolved photoemission spectroscopy (TR-ARPES) of TMDC monolayers and bilayers. Chem Sci. 2023;14:736–750. doi: 10.1039/D2SC04124C
  • Geim AK, Grigorieva IV. Van der Waals heterostructures. Nature. 2013;499:419–425. doi: 10.1038/nature12385
  • Chernikov A, Berkelbach TC, Hill HM, et al. Exciton binding energy and nonhydrogenic rydberg series in monolayer WS2. Phys Rev Lett. 2014;113:076802. doi: 10.1103/PhysRevLett.113.076802
  • He K, Kumar N, Zhao L, et al. Tightly bound excitons in monolayer WSe2. Phys Rev Lett. 2014;113:026803. doi: 10.1103/PhysRevLett.113.026803
  • Krömker B, Escher M, Funnemann D, et al. Development of a momentum microscope for time resolved band structure imaging. Rev Sci Instrum. 2008;79:053702–7. doi: 10.1063/1.2918133
  • Tusche C, Krasyuk A, Kirschner J. Spin resolved bandstructure imaging with a high resolution momentum microscope. Ultramicroscopy. 2015;159:520–529. doi: 10.1016/j.ultramic.2015.03.020
  • Medjanik K, Fedchenko O, Chernov S, et al. Direct 3D mapping of the fermi surface and fermi velocity. Nat Mater. 2017;16:615–621. doi: 10.1038/nmat4875
  • Kutnyakhov D, Xian RP, Dendzik M, et al. Time- and momentum-resolved photoemission studies using time-of-flight momentum microscopy at a free-electron laser. Rev Sci Instruments. 2020;91:013109. doi: 10.1063/1.5118777
  • Heyl CM, Güdde J, L’Huillier A, et al. High-order harmonic generation with μJ laser pulses at high repetition rates. J Phys B: At Mol Optical Phys. 2012;45:074020. doi: 10.1088/0953-4075/45/7/074020
  • Carstens H, Högner M, Saule T, et al. High-harmonic generation at 250 MHz with photon energies exceeding 100 eV. Optica. 2016;3:366–369. doi: 10.1364/OPTICA.3.000366
  • Chiang C-T, Huth M, Trützschler A, et al. Boosting laboratory photoelectron spectroscopy by megahertz high-order harmonics. New J Phys. 2015;17:013035. doi: 10.1088/1367-2630/17/1/013035
  • Keunecke M, Möller C, Schmitt D, et al. Time-resolved momentum microscopy with a 1 MHz high-harmonic extreme ultraviolet beamline. Rev Sci Instruments. 2020;91:063905. doi: 10.1063/5.0006531
  • Möller C, Probst H, Otto J, et al. Ultrafast element-resolved magneto-optics using a fiber-laser-driven extreme ultraviolet light source. Rev Sci Instruments. 2021;92:065107. doi: 10.1063/5.0050883
  • Puppin M, Deng Y, Nicholson CW, et al. Time- and angle-resolved photoemission spectroscopy of solids in the extreme ultraviolet at 500 kHz repetition rate. Rev Sci Instruments. 2019;90:023104. doi: 10.1063/1.5081938
  • Mathias S, Miaja-Avila L, Murnane MM, et al. Angle-resolved photoemission spectroscopy with a femtosecond high harmonic light source using a two-dimensional imaging electron analyzer. Rev Sci Instruments. 2007;78:083105. doi: 10.1063/1.2773783
  • Baumgärtner K, Reuner M, Metzger C, et al. Ultrafast orbital tomography of a pentacene film using time-resolved momentum microscopy at a FEL. Nat Commun. 2022;13:2741. doi: 10.1038/s41467-022-30404-6
  • Wallauer R, Raths M, Stallberg K, et al. Tracing orbital images on ultrafast time scales. Science. 2021;371:1056–1059. doi: 10.1126/science.abf3286
  • Adamkiewicz A, Raths M, Stettner M, et al. Coherent and incoherent excitation pathways in time-resolved photoemission orbital tomography of CuPc/Cu (001)-2O. J Phys Chem C. 2023;127:20411–20418. doi: 10.1021/acs.jpcc.3c04859
  • Puschnig P, Berkebile S, Fleming AJ, et al. Reconstruction of molecular orbital densities from photoemission data. Science. 2009;326:702–706. doi: 10.1126/science.1176105
  • Werner P. Time-resolved momentum microscopy of dark hybrid excitons in TMD bilayer systems [ master-thesis]. Georg August University Göttingen; 2023.
  • Zhang H, Pincelli T, Jozwiak C, et al. Angle-resolved photoemission spectroscopy. Nat Rev Methods Primers. 2022;2:54. doi: 10.1038/s43586-022-00133-7
  • Güdde J, Rohleder M, Meier T, et al. Time-resolved investigation of coherently controlled electric currents at a metal surface. Science. 2007;318:1287–1291. doi: 10.1126/science.1146764
  • Reutzel M, Li A, Wang Z, et al. Coherent multidimensional photoelectron spectroscopy of ultrafast quasiparticle dressing by light. Nat Commun. 2020;11:2230. doi: 10.1038/s41467-020-16064-4
  • Kern CS, Windischbacher A, Puschnig P. Photoemission orbital tomography for excitons in organic molecules. Phys Rev B. 2023;108:085132. doi: 10.1103/PhysRevB.108.085132
  • Bertoni R, Nicholson CW, Waldecker L, et al. Generation and evolution of spin-, valley-, and layer-polarized excited carriers in inversion-symmetric WSe2. Phys Rev Lett. 2016;117:277201. doi: 10.1103/PhysRevLett.117.277201
  • Wallauer R, Reimann J, Armbrust N, et al. Intervalley scattering in MoS2 imaged by two-photon photoemission with a high-harmonic probe. Appl Phys Lett. 2016;109:162102. doi: 10.1063/1.4965839
  • Hein P, Stange A, Hanff K, et al. Momentum-resolved hot electron dynamics at the 2H − MoS 2 surface. Phys Rev B. 2016;94:205406. doi: 10.1103/PhysRevB.94.205406
  • Wallauer R, Marauhn P, Reimann J, et al. Momentum-resolved observation of ultrafast interlayer charge transfer between the topmost layers of MoS2. Phys Rev B. 2020;102:125417. doi: 10.1103/PhysRevB.102.125417
  • Grubišić Čabo A, Miwa JA, Grønborg SS, et al. Observation of ultrafast free carrier dynamics in single layer MoS2. Nano Lett. 2015;15:5883–5887. doi: 10.1021/acs.nanolett.5b01967
  • Liu F, Ziffer ME, Hansen KR, et al. Direct determination of band-gap renormalization in the photoexcited monolayer MoS2. Phys Rev Lett. 2019;122:246803. doi: 10.1103/PhysRevLett.122.246803
  • Liu F, Li Q, Zhu XY. Direct determination of momentum-resolved electron transfer in the photoexcited van der Waals heterobilayer WS2/MoS2. Phys Rev B. 2020;101:201405. doi: 10.1103/PhysRevB.101.201405
  • Lin Y, Chan Y-H, Lee W, et al. Exciton-driven renormalization of quasiparticle band structure in monolayer MoS2. Phys Rev B. 2022;106:L081117. doi: 10.1103/PhysRevB.106.L081117
  • Lee W, Lin Y, Lu L-S, et al. Time-resolved ARPES determination of a quasi-particle band gap and hot electron dynamics in monolayer MoS2. Nano Lett. 2021;21:7363–7370. doi: 10.1021/acs.nanolett.1c02674
  • Aeschlimann S, Rossi A, Chávez-Cervantes M, et al. Direct evidence for efficient ultrafast charge separation in epitaxial WS2/graphene heterostructures. Sci Adv. 2020;6:eaay0761. doi: 10.1126/sciadv.aay0761
  • Dong S, Beaulieu S, Selig M, et al. Observation of ultrafast interfacial Meitner-Auger energy transfer in a Van der Waals heterostructure. Nat Commun. 2023;14:5057. doi: 10.1038/s41467-023-40815-8
  • Potočnik T, Burton O, Reutzel M, et al. Fast twist angle mapping of bilayer graphene using spectroscopic ellipsometric contrast microscopy. Nano Lett. 2023;23:5506–5513. doi: 10.1021/acs.nanolett.3c00619 PMID: 37289669.
  • Cattelan M, Fox NA. A perspective on the application of spatially resolved ARPES for 2D materials. Nanomaterials. 2018;8:284. doi: 10.3390/nano8050284
  • Rotenberg E, Bostwick A. microARPES and nanoARPES at diffraction-limited light sources: opportunities and performance gains. J Synchrotron Radiat. 2014;21:1048–1056. doi: 10.1107/S1600577514015409
  • Wilson NR, Nguyen PV, Seyler K, et al. Determination of band offsets, hybridization, and exciton binding in 2D semiconductor heterostructures. Sci Adv. 2017;3:e1601832. doi: 10.1126/sciadv.1601832
  • Ulstrup S, Koch RJ, Singh S, et al. Direct observation of minibands in a twisted graphene/WS2 bilayer. Sci Adv. 2020;6:eaay6104. doi: 10.1126/sciadv.aay6104
  • Lisi S, Lu X, Benschop T, et al. Observation of flat bands in twisted bilayer graphene. Nat Phys. 2021;17:189–193. doi: 10.1038/s41567-020-01041-x
  • Nguyen PV, Teutsch NC, Wilson NP, et al. Visualizing electrostatic gating effects in two-dimensional heterostructures. Nature. 2019;572:220–223. doi: 10.1038/s41586-019-1402-1
  • Hofmann P. Accessing the spectral function of in operando devices by angle-resolved photoemission spectroscopy. AVS Quantum Sci. 2021;3:021101. doi: 10.1116/5.0038637
  • Li A, James NA, Wang T, et al. Towards full surface brillouin zone mapping by coherent multi-photon photoemission. New J Phys. 2020;22:073035. doi: 10.1088/1367-2630/ab98d6
  • Jansen GSM, Keunecke M, Düvel M, et al. Efficient orbital imaging based on ultrafast momentum microscopy and sparsity-driven phase retrieval. New J Phys. 2020;22:063012. doi: 10.1088/1367-2630/ab8aae
  • Decking W, Abeghyan S, Abramian P, et al. A MHz-repetition-rate hard X-ray free-electron laser driven by a superconducting linear accelerator. Nat Photonics. 2020;14:391–397. doi: 10.1038/s41566-020-0607-z
  • Faatz B, Plönjes E, Ackermann S, et al. Simultaneous operation of two soft x-ray free-electron lasers driven by one linear accelerator. New J Phys. 2016;18:062002. doi: 10.1088/1367-2630/18/6/062002
  • Maklar J, Dong S, Beaulieu S, et al. A quantitative comparison of time-of-flight momentum microscopes and hemispherical analyzers for time- and angle-resolved photoemission spectroscopy experiments. Rev Sci Instruments. 2020;91:123112. doi: 10.1063/5.0024493
  • Keunecke M, Reutzel M, Schmitt D, et al. Electromagnetic dressing of the electron energy spectrum of Au(111) at high momenta. Phys Rev B. 2020;102:161403. doi: 10.1103/PhysRevB.102.161403
  • Li A, Reutzel M, Wang Z, et al. Multidimensional multiphoton momentum microscopy of the anisotropic Ag(110) surface. Phys Rev B. 2022;105:075105. doi: 10.1103/PhysRevB.105.075105
  • Barrett N, Conrad E, Winkler K, et al. Dark field photoelectron emission microscopy of micron scale few layer graphene. Rev Sci Instruments. 2012;83:083706. doi: 10.1063/1.4746279
  • Fujikawa Y, Sakurai T, Tromp RM. Micrometer-scale band mapping of single silver islands in real and reciprocal space. Phys Rev B. 2009;79:121401–121404. doi: 10.1103/PhysRevB.79.121401
  • Rohlfing M, Louie SG. Electron-hole excitations and optical spectra from first-principles. Phys Rev B. 2000;62:4927. doi: 10.1103/PhysRevB.62.4927
  • Hagel J, Brem S, Linderälv C, et al. Exciton landscape in van der Waals heterostructures. Phys Rev Res. 2021;3:043217. doi: 10.1103/PhysRevResearch.3.043217
  • Katsch F, Selig M, Carmele A, et al. Theory of exciton–exciton interactions in monolayer transition metal dichalcogenides. Physica Status Solidi (B). 2018;255:1800185. doi: 10.1002/pssb.201800185
  • Blase X, Duchemin I, Jacquemin D, et al. The Bethe–Salpeter equation formalism: from physics to chemistry. J Phys Chem Lett. 2020;11:7371–7382. doi: 10.1021/acs.jpclett.0c01875 pMID: 32787315.
  • Quintela MFCM, Henriques JACG, Tenório LGM, et al. Theoretical methods for excitonic physics in 2D materials. Physica Status solidi (B). 2022;259:2200097. doi: 10.1002/pssb.202200097
  • Casida ME. Time-dependent density functional response theory for molecules. In: Recent advances in density functional methods, recent advances in computational chemistry. Vol. 1. WORLD SCIENTIFIC; 1995. pp. 155–192. doi: 10.1142/9789812830586_0005
  • Andrade X, Strubbe D, Giovannini UD, et al. Real-space grids and the Octopus code as tools for the development of new simulation approaches for electronic systems. Phys Chem Chem Phys. 2015;17:31371–31396. doi: 10.1039/C5CP00351B
  • Tancogne-Dejean N, Oliveira MJT, Andrade X, et al. Octopus, a computational framework for exploring light-driven phenomena and quantum dynamics in extended and finite systems. J Chem Phys. 2020;152:124119. doi: 10.1063/1.5142502
  • Varsano D, Marini A, Rubio A. Optical saturation driven by exciton confinement in molecular chains: a time-dependent density-functional theory approach. Phys Rev Lett. 2008;101:133002. doi: 10.1103/PhysRevLett.101.133002
  • Cocchi C, Draxl C. Optical spectra from molecules to crystals: insight from many-body perturbation theory. Phys Rev B. 2015;92:205126. doi: 10.1103/PhysRevB.92.205126
  • Lin K-Q, Ong CS, Bange S, et al. Narrow-band high-lying excitons with negative-mass electrons in monolayer WSe2. Nat Commun. 2021;12:5500. doi: 10.1038/s41467-021-25499-2
  • Mak KF, Lee C, Hone J, et al. Atomically thin MoS2: a new direct-gap semiconductor. Phys Rev Lett. 2010;105:136805. doi: 10.1103/PhysRevLett.105.136805
  • Splendiani A, Sun L, Zhang Y, et al. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010;10:1271–1275. doi: 10.1021/nl903868w
  • Mueller T, Malic E. Exciton physics and device application of two-dimensional transition metal dichalcogenide semiconductors. Npj 2D Mater Appl. 2018;2:1–12. doi: 10.1038/s41699-018-0074-2
  • Zhang X-X, You Y, Zhao SYF, et al. Experimental evidence for dark excitons in monolayer WSe2. Phys Rev Lett. 2015;115:257403. doi: 10.1103/PhysRevLett.115.257403
  • Rivera P, Seyler KL, Yu H, et al. Valley-polarized exciton dynamics in a 2D semiconductor heterostructure. Science. 2016;351:688–691. doi: 10.1126/science.aac7820
  • Rivera P, Yu H, Seyler KL, et al. Interlayer valley excitons in heterobilayers of transition metal dichalcogenides. Nat Nanotechnol. 2018;13:1004–1015. doi: 10.1038/s41565-018-0193-0
  • Ohnishi H, Tomita N, Nasu K. Direct determination of exciton wavefunction amplitudes by the momentum-resolved photo-electron emission experiment. Int J Mod Phys B. 2018;32:1850094. doi: 10.1142/s0217979218500947
  • Christiansen D, Selig M, Malic E, et al. Theory of exciton dynamics in time-resolved ARPES: intra- and intervalley scattering in two-dimensional semiconductors. Phys Rev B. 2019;100:205401. doi: 10.1103/PhysRevB.100.205401
  • Steinhoff A, Florian M, Rösner M, et al. Exciton fission in monolayer transition metal dichalcogenide semiconductors. Nat Commun. 2017;8:1166. doi: 10.1038/s41467-017-01298-6
  • Meneghini G, Reutzel M, Mathias S, et al. Hybrid exciton signatures in ARPES spectra of van der Waals materials. ACS Photonics. 2023;10:3570–3575. doi: 10.1021/acsphotonics.3c00599
  • Perfetto E, Sangalli D, Marini A, et al. First-principles approach to excitons in time-resolved and angle-resolved photoemission spectra. Phys Rev B. 2016;94:245303. doi: 10.1103/PhysRevB.94.245303
  • Krylov AI. From orbitals to observables and back. J Chem Phys. 2020;153:080901. doi: 10.1063/5.0018597
  • Martin RL. Natural transition orbitals. J Chem Phys. 2003;118:4775–4777. doi: 10.1063/1.1558471
  • Hammon S, Kümmel S. Pump-probe photoemission simulated in real time: revealing many-particle signatures. Phys Rev A. 2021;104:012815. doi: 10.1103/PhysRevA.104.012815
  • Cheiwchanchamnangij T, Lambrecht WRL. Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2. Phys Rev B. 2012;85:205302. doi: 10.1103/PhysRevB.85.205302
  • De Giovannini U, Hübener H, Rubio A. A first-principles time-dependent density functional theory framework for spin and time-resolved angular-resolved photoelectron spectroscopy in periodic systems. J Chem Theory Comput. 2017;13:265–273. doi: 10.1021/acs.jctc.6b00897
  • Selig M, Berghäuser G, Richter M, et al. Dark and bright exciton formation, thermalization, and photoluminescence in monolayer transition metal dichalcogenides. 2D Mater. 2018;5:035017. doi: 10.1088/2053-1583/aabea3
  • Hong X, Kim J, Shi S-F, et al. Ultrafast charge transfer in atomically thin MoS2/WS2 heterostructures. Nat Nanotechnol. 2014;9:682–686. doi: 10.1038/nnano.2014.167
  • Lee C-H, Lee G-H, van der Zande AM, et al. Atomically thin p–n junctions with van der waals heterointerfaces. Nat Nanotechnol. 2014;9:676–681. doi: 10.1038/nnano.2014.150
  • Jin C, Ma EY, Karni O, et al. Ultrafast dynamics in van der waals heterostructures. Nat Nanotechnol. 2018;13:994–1003. doi: 10.1038/s41565-018-0298-5
  • Jin W, Yeh P-C, Zaki N, et al. Direct measurement of the thickness-dependent electronic band structure of MoS2 using angle-resolved photoemission spectroscopy. Phys Rev Lett. 2013;111:106801. doi: 10.1103/PhysRevLett.111.106801
  • Meneghini G, Brem S, Malic E. Ultrafast phonon-driven charge transfer in van der Waals heterostructures. Nat Sci. 2022;2:e20220014. doi: 10.1002/ntls.20220014
  • Brem S, Lin K-Q, Gillen R, et al. Hybridized intervalley moiré excitons and flat bands in twisted WSe2 bilayers. Nanoscale. 2020;12:11088–11094. doi: 10.1039/D0NR02160A
  • Wang Y, Wang Z, Yao W, et al. Interlayer coupling in commensurate and incommensurate bilayer structures of transition-metal dichalcogenides. Phys Rev B. 2017;95:115429. doi: 10.1103/PhysRevB.95.115429
  • Kunstmann J, Mooshammer F, Nagler P, et al. Momentum-space indirect interlayer excitons in transition-metal dichalcogenide van der Waals heterostructures. Nat Phys. 2018;14:801–805. doi: 10.1038/s41567-018-0123-y
  • Merkl P, Mooshammer F, Brem S, et al. Twist-tailoring coulomb correlations in van der waals homobilayers. Nat Commun. 2020;11:2167. doi: 10.1038/s41467-020-16069-z
  • van der Zande AM, Kunstmann J, Chernikov A, et al. Tailoring the electronic structure in bilayer molybdenum disulfide via interlayer twist. Nano Lett. 2014;14:3869–3875. doi: 10.1021/nl501077m
  • Liu K, Zhang L, Cao T, et al. Evolution of interlayer coupling in twisted molybdenum disulfide bilayers. Nat Commun. 2014;5:4966. doi: 10.1038/ncomms5966
  • Zheng Q, Saidi WA, Xie Y, et al. Phonon-assisted ultrafast charge transfer at van der waals heterostructure interface. Nano Lett. 2017;17:6435–6442. doi: 10.1021/acs.nanolett.7b03429
  • Zimmermann JE, Axt M, Mooshammer F, et al. Ultrafast charge-transfer dynamics in twisted MoS2/WSe2 heterostructures. ACS Nano. 2021;15:14725–14731. doi: 10.1021/acsnano.1c04549
  • Policht VR, Mittenzwey H, Dogadov O, et al. Time-domain observation of interlayer exciton formation and thermalization in a MoSe2/WSe2 heterostructure. Nat Commun. 2023;14:7273. doi: 10.1038/s41467-023-42915-x
  • Zhang C, Chuu C-P, Ren X, et al. Interlayer couplings, Moiré patterns, and 2D electronic superlattices in MoS 2 /WSe 2 hetero-bilayers. Sci Adv. 2017;3:e1601459. doi: 10.1126/sciadv.1601459
  • Wu F, Lovorn T, MacDonald AH. Topological exciton bands in moiré heterojunctions. Phys Rev Lett. 2017;118:147401. doi: 10.1103/PhysRevLett.118.147401
  • Wu F, Lovorn T, Tutuc E, et al. Hubbard model physics in transition metal dichalcogenide moiré bands. Phys Rev Lett. 2018;121:026402. doi: 10.1103/PhysRevLett.121.026402
  • Cao Y, Fatemi V, Fang S, et al. Unconventional superconductivity in magic-angle graphene superlattices. Nature. 2018;556:43–50. doi: 10.1038/nature26160
  • Wang L, Shih E-M, Ghiotto A, et al. Correlated electronic phases in twisted bilayer transition metal dichalcogenides. Nat Mater. 2020;19:861–866. doi: 10.1038/s41563-020-0708-6
  • Regan EC, Wang D, Jin C, et al. Mott and generalized Wigner crystal states in WSe2/WS2 moiré superlattices. Nature. 2020;579:359–363. doi: 10.1038/s41586-020-2092-4
  • Alexeev EM, Ruiz-Tijerina DA, Danovich M, et al. Resonantly hybridized excitons in moiré superlattices in van der waals heterostructures. Nature. 2019;567:81–86. doi: 10.1038/s41586-019-0986-9
  • Tran K, Moody G, Wu F, et al. Evidence for moiré excitons in van der waals heterostructures. Nature. 2019;567:71–75. doi: 10.1038/s41586-019-0975-z
  • Seyler KL, Rivera P, Yu H, et al. Signatures of moiré-trapped valley excitons in MoSe2/WSe2 heterobilayers. Nature. 2019;567:66–70. doi: 10.1038/s41586-019-0957-1
  • Brem S, Linderälv C, Erhart P, et al. Tunable phases of Moiré excitons in van der Waals heterostructures. Nano Lett. 2020;20:8534–8540. doi: 10.1021/acs.nanolett.0c03019
  • Stansbury CH, Utama MIB, Fatuzzo CG, et al. Visualizing electron localization of WS2/WSe2 moiré superlattices in momentum space. Sci Adv. 2021;7:eabf4387. doi: 10.1126/sciadv.abf4387
  • Gatti G, Issing J, Rademaker L, et al. Flat Γ moiré bands in twisted bilayer WSe2. Phys Rev Lett. 2023;131:046401. doi: 10.1103/PhysRevLett.131.046401
  • Ahn SJ, Moon P, Kim T-H, et al. Dirac electrons in a dodecagonal graphene quasicrystal. Science. 2018;361:782–786. doi: 10.1126/science.aar8412
  • Koshino M. Interlayer interaction in general incommensurate atomic layers. New J Phys. 2015;17:015014. doi: 10.1088/1367-2630/17/1/015014
  • Stadtmüller B, Emmerich S, Jungkenn D, et al. Strong modification of the transport level alignment in organic materials after optical excitation. Nat Commun. 2019;10:1470. doi: 10.1038/s41467-019-09136-7
  • Emmerich S, Hedwig S, Arnoldi B, et al. Ultrafast charge-transfer exciton dynamics in C60 thin films. J Phys Chem C. 2020;124:23579–23587. doi: 10.1021/acs.jpcc.0c08011
  • Bradshaw AM, Woodruff DP. Molecular orbital tomography for adsorbed molecules: is a correct description of the final state really unimportant? New J Phys. 2015;17:013033. doi: 10.1088/1367-2630/17/1/013033
  • Egger L, Kollmann B, Hurdax P, et al. Can photoemission tomography be useful for small, strongly-interacting adsorbate systems? New J Phys. 2019;21:043003. doi: 10.1088/1367-2630/ab0781
  • Kern CS, Haags A, Egger L, et al. Simple extension of the plane-wave final state in photoemission: bringing understanding to the photon-energy dependence of two-dimensional materials. Phys Rev Res. 2023;5:033075. doi: 10.1103/PhysRevResearch.5.033075
  • Dauth M, Körzdörfer T, Kümmel S, et al. Orbital density reconstruction for molecules. Phys Rev Lett. 2011;107:193002. doi: 10.1103/PhysRevLett.107.193002
  • Yang X, Jugovac M, Zamborlini G, et al. Momentum-selective orbital hybridisation. Nat Commun. 2022;13:5148. doi: 10.1038/s41467-022-32643-z
  • Kliuiev P, Zamborlini G, Jugovac M, et al. Combined orbital tomography study of multi-configurational molecular adsorbate systems. Nat Commun. 2019;10:5255. doi: 10.1038/s41467-019-13254-7
  • Miao J, Charalambous P, Kirz J, et al. Extending the methodology of X-ray crystallography to allow imaging of micrometre-sized non-crystalline specimens. Nature. 1999;400:342–344. doi: 10.1038/22498
  • Lüftner D, Ules T, Reinisch EM, et al. Imaging the wave functions of adsorbed molecules. Proc Natl Acad Sci. 2014;111:605–610. doi: 10.1073/pnas.1315716110
  • Weiß S, Lüftner D, Ules T, et al. Exploring three-dimensional orbital imaging with energy-dependent photoemission tomography. Nat Commun. 2015;6:8287. doi: 10.1038/ncomms9287
  • Dinh TL, Jansen GSM, Luke DR, et al. A minimalist approach to 3D photoemission orbital tomography: algorithms and data requirements. New J Phys. 2024;26:043024. doi: 10.1088/1367-2630/ad3e22
  • Pei J, Yang J, Yildirim T, et al. Many-body complexes in 2D semiconductors. Adv Mater. 2019;31:1706945. doi: 10.1002/adma.201706945
  • Barbone M, Montblanch A-P, Kara DM, et al. Charge-tuneable biexciton complexes in monolayer WSe2. Nat Commun. 2018;9:3721. doi: 10.1038/s41467-018-05632-4
  • Vektaris G. A new approach to the molecular biexciton theory. J Chem Phys. 1994;101:3031–3040. doi: 10.1063/1.467616
  • Wannier GH. The structure of electronic excitation levels in insulating crystals. Phys Rev. 1937;52:191–197. doi: 10.1103/PhysRev.52.191
  • Cudazzo P, Gatti M, Rubio A, et al. Frenkel versus charge-transfer exciton dispersion in molecular crystals. Phys Rev B. 2013;88:195152. doi: 10.1103/PhysRevB.88.195152
  • Knoester J, Agranovich VM. Frenkel and charge-transfer excitons in organic solids. In: Electronic excitations in organic nanostructures, thin films and nanostructures. Vol. Vol. 31. Academic Press; 2003. p. 1–96. doi: 10.1016/S1079-4050(03)31001-4
  • Agranovich VM. Excitations in organic solids. Vol. 142. Oxford: OUP Oxford; 2009.
  • Zhu X-Y, Yang Q, Muntwiler M. Charge-transfer excitons at organic semiconductor surfaces and interfaces. Acc Chem Res. 2009;42:1779–1787. doi: 10.1021/ar800269u
  • Tanda Bonkano B, Palato S, Krumland J, et al. Evidence for hybrid inorganic–organic transitions at the ws2/terrylene interface. Physica Status Solidi (A). 2024;221:2300346. doi: 10.1002/pssa.202300346
  • Gonzalez Oliva I, Caruso F, Pavone P, et al. Hybrid excitations at the interface between a MoS2 monolayer and organic molecules: a first-principles study. Phys Rev Mater. 2022;6:054004. doi: 10.1103/PhysRevMaterials.6.054004
  • Smith MB, Michl J. Singlet fission. Chem Rev. 2010;110:6891–6936. doi: 10.1021/cr1002613
  • Guo Q, Dendzik M, Grubišić-Čabo A, et al. A narrow bandwidth extreme ultra-violet light source for time- and angle-resolved photoemission spectroscopy. Struct Dyn. 2022;9:024304. doi: 10.1063/4.0000149
  • Pareek V, Bacon DR, Zhu X, et al. Driving non-trivial quantum phases in conventional semiconductors with intense excitonic fields. arXiv preprint arXiv:2403.08725. 2024.
  • Cavalieri AL, Müller N, Uphues T, et al. Attosecond spectroscopy in condensed matter. Nature. 2007;449:1029–1032. doi: 10.1038/nature06229
  • Neppl S, Ernstorfer R, Cavalieri AL, et al. Direct observation of electron propagation and dielectric screening on the atomic length scale. Nature. 2015;517:342–346. doi: 10.1038/nature14094
  • Tao Z, Chen C, Szilvási T, et al. Direct time-domain observation of attosecond final-state lifetimes in photoemission from solids. Science. 2016;353:62–67. Available from: https://science.sciencemag.org/content/353/6294/62.full.pdf
  • Chen C, Tao Z, Carr A, et al. Distinguishing attosecond electron–electron scattering and screening in transition metals. Proc Natl Acad Sci USA. 2017;114:E5300–E5307. doi: 10.1073/pnas.1706466114
  • Siek F, Neb S, Bartz P, et al. Angular momentum–induced delays in solid-state photoemission enhanced by intra-atomic interactions. Science. 2017;357:1274–1277. doi: 10.1126/science.aam9598
  • Meneghini G, Brem S, Malic E. Excitonic thermalization bottleneck in twisted TMD heterostructures. Nano Lett. 2024;24:4505–4511. doi: 10.1021/acs.nanolett.4c00450
  • Dendzik M, Xian RP, Perfetto E, et al. Observation of an excitonic Mott transition through ultrafast core-cum-conduction photoemission spectroscopy. Phys Rev Lett. 2020;125:096401. doi: 10.1103/PhysRevLett.125.096401
  • Curcio D, Pakdel S, Volckaert K, et al. Ultrafast electronic linewidth broadening in the C 1s core level of graphene. Phys Rev B. 2021;104:L161104. doi: 10.1103/PhysRevB.104.L161104
  • Graus M, Metzger C, Grimm M, et al. Three-dimensional tomographic imaging of molecular orbitals by photoelectron momentum microscopy. The Eur Phys J B. 2019;92:80. doi: 10.1140/epjb/e2019-100015-x
  • Weiß S, Lüftner D, Ules T, et al. Exploring three-dimensional orbital imaging with energy-dependent photoemission tomography. Nat Commun. 2015;6:8287. doi: 10.1038/ncomms9287
  • Reimann J, Schlauderer S, Schmid CP, et al. Subcycle observation of lightwave-driven dirac currents in a topological surface band. Nature. 2018;562:396–400. doi: 10.1038/s41586-018-0544-x
  • Ito S, Schüler M, Meierhofer M, et al. Build-up and dephasing of floquet–Bloch bands on subcycle timescales. Nature. 2023;616:696–701. doi: 10.1038/s41586-023-05850-x
  • Borsch M, Meierhofer M, Huber R, et al. Lightwave electronics in condensed matter. Nat Rev Mater. 2023;8:668–687. doi: 10.1038/s41578-023-00592-8
  • Merboldt M, Schüler M, Schmitt D, et al. Observation of Floquet states in graphene. arXiv:2404.12791 [cond-mat.mes-hall]. 2024.
  • Choi D, Mogi M, De Giovannini U, et al. Direct observation of Floquet-Bloch states in monolayer graphene. arXiv:2404.14392 [cond-mat.mes-hall]. 2024.
  • Petek H, Ogawa S. Femtosecond time-resolved two-photon photoemission studies of electron dynamics in metals. Prog Surf Sci. 1997;56:239–310. doi: 10.1016/S0079-6816(98)00002-1
  • Aeschlimann M, Brixner T, Fischer A, et al. Coherent two-dimensional nanoscopy. Science. 2011;333:1723–1726. doi: 10.1126/science.1209206
  • Reutzel M, Li A, Petek H. Coherent two-dimensional multiphoton photoelectron spectroscopy of metal surfaces. Phys Rev X. 2019;9:011044. doi: 10.1103/PhysRevX.9.011044
  • Uhl D, Bangert U, Bruder L, et al. Coherent optical 2D photoelectron spectroscopy. Optica. 2021;8:1316–1324. doi: 10.1364/OPTICA.434853
  • Eul T, Prinz E, Hartelt M, et al. Coherent response of the electronic system driven by non-interfering laser pulses. Nat Commun. 2022;13:3324. doi: 10.1038/s41467-022-30768-9
  • Aeschlimann M, Brixner T, Cinchetti M, et al. Observation of optical coherence in a disordered metal-molecule interface by coherent optical two-dimensional photoelectron spectroscopy. Phys Rev B. 2022;105:205415. doi: 10.1103/PhysRevB.105.205415
  • Reutzel M, Li A, Wang Z, et al. Probing nonlinear light–matter interaction in momentum space: coherent multiphoton photoemission spectroscopy. In: Matsuda I Arafune R, editors. Nonlinear X-ray spectroscopy for materials science. Singapore: Springer Nature Singapore; 2023. p. 57–82. doi: 10.1007/978-981-99-6714-8_3
  • Reutzel M, Li A, Petek H. Above-threshold multiphoton photoemission from noble metal surfaces. Phys Rev B. 2020;101:075409. doi: 10.1103/PhysRevB.101.075409
  • Oka T, Kitamura S. Floquet engineering of quantum materials. Annu Rev Condens Matter Phys. 2019;10:387–408. doi: 10.1146/annurev-conmatphys-031218-013423
  • de la Torre A, Kennes DM, Claassen M, et al. Colloquium: nonthermal pathways to ultrafast control in quantum materials. Rev Mod Phys. 2021;93:041002. doi: 10.1103/RevModPhys.93.041002
  • Bettis Homan S, Sangwan VK, Balla I, et al. Ultrafast exciton dissociation and long-lived charge separation in a photovoltaic Pentacene–MoS2 van der waals heterojunction. Nano Lett. 2017;17:164–169. doi: 10.1021/acs.nanolett.6b03704
  • Rijal K, Rudayni F, Kafle TR, et al. Collective effects of band offset and wave function dimensionality on impeding electron transfer from 2D to organic crystals. J Phys Chem Lett. 2020;11:7495–7501. doi: 10.1021/acs.jpclett.0c01796
  • Kafle TR, Kattel B, Lane SD, et al. Charge transfer exciton and spin flipping at organic–transition-metal dichalcogenide interfaces. ACS Nano. 2017;11:10184–10192. doi: 10.1021/acsnano.7b04751
  • Zhong C, Sangwan VK, Wang C, et al. Mechanisms of ultrafast charge separation in a PTB7/monolayer MoS2 van der waals heterojunction. J Phys Chem Lett. 2018;9:2484–2491. doi: 10.1021/acs.jpclett.8b00628
  • Günder D, Axt M, Witte G. Heteroepitaxy in organic/TMD hybrids and challenge to achieve it for TMD monolayers: the case of pentacene on WS2 and WSe2. ACS Appl Mater Interface. 2023;16:1911–1920. doi: 10.1021/acsami.3c15829
  • Arnoldi B, Zachritz SL, Hedwig S, et al. Revealing hidden spin polarization in centrosymmetric van der waals materials on ultrafast timescales. Nat Commun. 2024;15:3573. doi: 10.1038/s41467-024-47821-4
  • Taghinejad H, Eftekhar AA, Adibi A. Lateral and vertical heterostructures in two-dimensional transition-metal dichalcogenides [Invited]. Optical Mater Express. 2019;9:1590. doi: 10.1364/ome.9.001590
  • Rosati R, Paradisanos I, Huang L, et al. Interface engineering of charge-transfer excitons in 2D lateral heterostructures. Nat Commun. 2023;14:2438. doi: 10.1038/s41467-023-37889-9
  • Gonçalves PAD, Bertelsen LP, Xiao S, et al. Plasmon-exciton polaritons in two-dimensional semiconductor/metal interfaces. Phys Rev B. 2018;97:041402. doi: 10.1103/PhysRevB.97.041402
  • Wang X, Lin Z, Watanabe K, et al. Near-field coupling of interlayer excitons in MoSe2/WSe2 heterobilayers to surface plasmon polaritons. J Chem Phys. 2024;160. doi: 10.1063/5.0201383
  • Jiang Z, Hsieh K, Jones AJH, et al. Revealing flat bands and hybridization gaps in a twisted bilayer graphene device with microARPES. 2D Mater. 2023;10:045027. doi: 10.1088/2053-1583/acf775
  • Dale N, Utama MIB, Lee D, et al. Layer-dependent interaction effects in the electronic structure of twisted bilayer graphene devices. Nano Lett. 2023;23:6799–6806. doi: 10.1021/acs.nanolett.3c00253
  • Zhang H, Li Q, Park Y, et al. Observation of dichotomic field-tunable electronic structure in twisted monolayer-bilayer graphene. Nat Commun. 2024;15:3737. doi: 10.1038/s41467-024-48166-8
  • Wilson NP, Lee K, Cenker J, et al. Interlayer electronic coupling on demand in a 2D magnetic semiconductor. Nat Mater. 2021;20:1657–1662. doi: 10.1038/s41563-021-01070-8
  • Bae YJ, Wang J, Scheie A, et al. Exciton-coupled coherent magnons in a 2D semiconductor. Nature. 2022;609:282–286. doi: 10.1038/s41586-022-05024-1
  • Danz T, Domröse T, Ropers C. Ultrafast nanoimaging of the order parameter in a structural phase transition. Science. 2021;371:371–374. doi: 10.1126/science.abd2774
  • Perea-Causin R, Brem S, Schmidt O, et al. Trion photoluminescence and trion stability in atomically thin semiconductors. Phys Rev Lett. 2024;132:036903. doi: 10.1103/PhysRevLett.132.036903
  • Li H, Li S, Regan EC, et al. Imaging two-dimensional generalized Wigner crystals. Nature. 2021;597:650–654. doi: 10.1038/s41586-021-03874-9
  • Smoleński T, Dolgirev PE, Kuhlenkamp C, et al. Signatures of Wigner crystal of electrons in a monolayer semiconductor. Nature. 2021;595:53–57. doi: 10.1038/s41586-021-03590-4
  • Zhou Y, Sung J, Brutschea E, et al. Bilayer wigner crystals in a transition metal dichalcogenide heterostructure. Nature. 2021;595:48–52. doi: 10.1038/s41586-021-03560-w
  • Arsenault EA, Li Y, Yang B, et al. Two-dimensional moiré polaronic electron crystals. Phys Rev Lett. 2024;132:126501. doi: 10.1103/PhysRevLett.132.126501
  • Schönhense G, Medjanik K, Elmers H-J. Space-, time- and spin-resolved photoemission. J Electron Spectrosc Relat Phenom. 2015;200:94–118. special Anniversary Issue: Volume 200. doi: 10.1016/j.elspec.2015.05.016
  • Schönhense G, Medjanik K, Chernov S, et al. Spin-filtered time-of-flight k-space microscopy of Ir – Towards the “complete” photoemission experiment. Ultramicroscopy. 2017;183:19–29. lEEM/PEEM-10. doi: 10.1016/j.ultramic.2017.06.025