Ai Apl#241334
Ai Apl#241334
1
Graduate School of Science and Engineering, Chiba University, 1-33 Yayoi-cho, Inage-ku, Chiba,
263-8522, Japan.
2
i-Powered Energy System Research Center, The University of Electro-Communications, 1-5-1
3
Graduate School of Informatics and Engineering, The University of Electro-Communications, 1-5-1
4
Graduate School of Engineering, Chiba University, 1-33 Yayoi-cho, Inage-ku, Chiba, 263-8522,
Japan.
1
5
Molecular Chirality Research Center, Chiba University, 1-33 Yayoi-cho, Inage-ku, Chiba, 263-8522,
Japan.
*a)
Corresponding Author to whom correspondence should be addressed: [email protected]
Abstract
selective layers (HSLs) in inverted perovskite solar cells. As an HSL, the electron-blocking
capability is important and directly related to electron affinity (EA). Low-energy inverse
photoelectron spectroscopy (LEIPS) is the most reliable method for EA measurement. However, the
intense electron-impact-induced fluorescence from the carbazole interferes with the measurement.
By improving the photon detector, we were able to measure 2PACz and MeO-2PACz LEIPS spectra
and determine their respective EAs of 1.73 eV and 1.48 eV. These small EA values ensure effective
Perovskite solar cells (PSCs) have attracted tremendous attention because they can be fabricated by a
low-cost solution process and achieve a power conversion efficiency (PCE) of 26.1% [%.1].
2,2′’,7,7′’-Tetrakis[N,N-di(4-methoxyphenyl)amino]-9,9′’-spirobifluorene (spiro-OMeTAD) is
exclusively used as a hole-selective layer (HSL) in a normal (n-–i-–p) structure PSC with high PCE.
However, spiro-OMeTAD requires doping and has stability issues [.2]. On the other hand, such stable
1
poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine] (PTAA) are used as an HSL in an inverted (p-–i-–
n) structure PSC [.3]. However, the polymer layer is usually around 10 to –40 nm thick, which
A self-assembled monolayer (SAM) has been developed as a substitute for these polymer
HSLs. As SAM-based HSL is very thin (approximately 1 nm) [5] and firmly binds to a transparent
electrode, it hardly absorbs light and has high stability [.4]. Among the various SAM materials,
carbazole derivatives are the most widely studied. A carbazole-based SAM molecule comprises a
carbazole backbone, an anchor moiety binding to the transparent electrode surface, and an alkyl
chain connecting the carbazole backbone and the anchor moiety. The SAM molecules with alkyl
chains CnH2n of various lengths (n = = 2, 4, and 6) and substituents such as methyl and methoxy
To achieve good performance as HSL, high hole collection efficiency and electron-blocking
capability are required, and these can be assessed from the HOMO and LUMO energy levels,
respectively. If the vacuum level alignment is assumed for simplicity, the ionization energy (IE;
HOMO energy level with reference to the vacuum level) should lie between the IE of the perovskite
surface and the work function of the transparent electrode. Conversely, the electron affinity (EA; the
LUMO energy level with reference to the vacuum level) should be smaller than that of the perovskite
layer for electron blocking. Therefore, precise and reliable IE and EA values of HSL are essential to
evaluate and choose a suitable SAM material. In addition, interfacial hole/electron recombination
1
crucially impacts PSC performance. The recombination rate depends on IE and EA as well as the trap
density [.[6,7].] To discuss the recombination dynamics in detail, accurate IE and EA values of HSL are
indispensable.
of a film. Conversely, EA is often predicted from the IE and the optical gap [4]gap4 assuming that the
transport gap (IE−EA) is equal to the optical gap. However, the transport gap and the optical gap
differ owing to the exciton binding energy (usually ranging from 0.2 eV to 1 eV) for organic
materials [.8-–10]. Thus, EA is often overestimated from the IE and the optical gap. IE and EA are also
estimated from the oxidation and reduction potentials derived from cyclic voltammetry (CV)
measurement. However, CV data are affected by the solvent and the supporting electrolyte, and the
measured values differ from those in the film [.11-–14]. In principle, EA can be best measured by inverse
photoelectron spectroscopy (IPES), which is regarded as the inversion process of UPS. However,
damage to organic materials by electron irradiation has hindered the application of conventional
IPES. Although IPES has provided unprecedent information about the unoccupied states of organic
semiconductors over 30 years [,15-–20], its applicability is limited. The development of low-energy
IPES (LEIPS) has made precise EA measurement possible [.[21,22].] In LEIPS, an electron with less
than 5 eV kinetic energy (i.e., the damage threshold of most organic materials) and a photon emitted
by the radiative transition from a free electron to the unoccupied states (including LUMO) are
detected. As the photon energy is in the near-ultraviolet or visible range, the photons are detected by
1
a combination of a bandpass filter and a photomultiplier. LEIPS has enabled thickness-dependent
measurements to study the interface energy level alignment of unoccupied states [,23], the precise
determination of exciton binding energy [8]energy8 and polarization energy [,24-–26], the electron
injection barrier in organic light emitting diode [,27], analysis of the charge separation mechanism in
organic solar cells [28,29] and measuring the on-site coulomb repulsion energy (Hubberd U) in
polymers [,30], etc. Recently, the angle-resolved measurement to observe conduction band structure
[31,32]
has been achieved, which has never been possible with the conventional IPES. However, the
because the carbazole derivatives generate intense electron-induced fluorescence that interferes with
the detection of the weak photon signal by LEIPS. To observe the LEIPS spectrum of the carbazole
We choose the two most widely used materials (2-(9H-Carbazol-9-yl)ethyl)phosphonic acid (2PACz)
we explain the problem arising from the EA measurement of the SAMs using LEIPS. Second, we
show an improved experimental setup to solve this problem. Third, we measure the LEIPS spectra of
2PACz and MeO-2PACz. The determined EAs corroborate the good electron-blocking capability of
1
2PACz and MeO-2PACz film samples were prepared by spin coating 1 mM solutions of
2PACz and MeO-2PACz in isopropyl alcohol (IPA) onto ITO substrates at the University of Electro-
Communications. After annealing the samples at 100 °C for 10 minutes, the unbonded 2PACz and
MeO-2PACz molecules on the ITO substrates were washed off with IPA, and the washed samples
The samples were kept in a nitrogen atmosphere and transported to Chiba University for
LEIPS measurement. The details of the LEIPS setup are described elsewhere [.33]. The samples were
irradiated with 1.0 to –6.5 eV electron beams. The emitted photons were detected by a bandpass
detector consisting of a bandpass filter and a photomultiplier. We have used the photomultiplier tube,
Hamamatsu Photonics R585s, since the development of LEIPS because of its low dark counts. We
find that the photomultiplier tube should be replaced in this work. Because the photon detector is key
to this work, we will discuss the details later. The work function was measured from the onset of the
Figure 1 shows the LEIPS spectra of 2PACz measured with a bandpass filter (center
wavelengths of 232 nm and 260 nm) and a photomultiplier tube (Hamamatsu Photonics R585s). The
spectral onset shifted from -−0.22 eV to 0.42 eV when the bandpass filter was changed from 232
nm to 260 nm. If the photon signal originated from the IPES process, the electron binding energy Eb
= h – Ek should be independent of the photon energy h. Because the onset difference of 0.6 eV
corresponds approximately to the photon energy difference, 5.3 eV (232 nm) and 4.8 eV (260 nm),
1
the process depends only on the electron energy Ek. Furthermore, the signal was observed even
below the Fermi level when the photon with the wavelength of 232 nm was detected. Likely, the
excitation [.34]. Carbazole is a highly emissive molecule that shows fluorescence by electron impact
[ 35]
. . Its spectrum is similar to photon-excited fluorescence due to the S1->S0 transition from 330 nm
to 400 nm. It was reported that a small amount of impurities in carbazole samples may cause
fluorescence [.36]. We used a bandpass filter with a center wavelength of 232 nm (5.28 eV) or 260
nm (4.79 eV), the energy of which seemed too high for fluorescence detection below 330 nm, and
inspected the sensitivity curves of our photon detector. Figure 2 shows the quantum efficiencies of
the R585s photomultiplier tube and of the photon detector combined with the bandpass filters. The
photon detector shows high sensitivity at the center wavelength of the bandpass filter (260 or 232
nm) and the long tail extending toward the long-wavelength region. Even around 330–400 nm where
carbazole shows strong fluorescence, the photon detector has low but non-negligible sensitivity.
Because the photon signal intensity is the sensitivity integrated over the wavelength, the signal
intensity of the photon detector tail region should be sufficiently high. Furthermore, the inverse
photoelectron process has a very small cross section [,37], and typical photon intensity in the present
measurement ranges from 10 to 100 cps, whereas the fluorescence induced by the electron impact is
so intense that it is sometimes visible to the naked eye. Therefore, we conclude that the fluorescence
1
from carbazole is so strong that a sufficient amount of photons are transmitted through the bandpass
The R585s photomultiplier tube uses a bi-alkali (Sb-–Rb-–Cs and Sb-–K-–Cs) photocathode
and has a sensitivity range of 700 nm to 160 nm (FigureFig. 2). Conversely, a photomultiplier tube
with a Cs-–Te “solar-blind” photocathode is not sensitive to wavelengths longer than 300 nm [.38].
Aiming to avoid the intense photon signal in the visible range, we combined a Cs-–Te
photomultiplier tube with near-ultraviolet bandpass filters. Unfortunately, the Cs-–Te photomultiplier
tube for photon counting use is not available, so we speciallyin particular, selected a low dark count
tube from the manufacture’s stock of Hamamatsu Photonics R821 and used it for photon counting.
Note that this problem is specific to LEIPS, because LEIPS analyzes visible/near-UV photons
whose energies are close to those of molecular fluorescence in the visible range. The high energy
selectivity is crucial to safely observe the LEIPS signal. Conversely, the conventional IPES using the
bandpass photon detector in the vacuum ultraviolet range has a sufficient energy difference to block
the molecular fluorescence. The photon detector developed by Dose, using a gas-filled Geiger-–
Mueller counter [39]counter39 is totally insensitive to photons below the ionization energy of the filled
gas (> 9 eV of I2, isopropanol, acetone, etc. [40]).) The solid-state bandpass detector using the CuBe
electron multiplier [,41], which was later improved by coating the first dynode with an alkali halide
[ 42]
, , also has negligible sensitivity in the visible range.
1
FigureFIG. 1.: LEIPS spectra of 2PACz measured with a bandpass filter (232 and 260 nm) and a
FigureFIG. 2.: Sensitivity curves of photon detector. (a) Quantum efficiencies of photomultiplier
tubes R585s [43]R585s43 and R821[.44]. (b) Comparison of quantum efficiencies of photon detector
comprised of the bandpass filter (260 nm) and photomultiplier tubes R585s and R821. (c) The same
as panel (b) except for the bandpass filter with a center wavelength of 232 nm.
Figure 3 shows the LEIPS spectra of 2PACz and MeO-2PACz measured with the photon
detector consisting of a bandpass filter with a center wavelength of 285 nm and the photomultiplier
tube R821. We observed clear onsets at 3.49 eV (2PACz) and 3.32 eV (MeO-2PACz) above the
Fermi level. The LEIPS spectrum of pristine ITO substrate exhibits a gradual rise around 1 eV (not
shown). We assigned the small increases in the background to the signal from the ITO substrate.
When the 285 nm bandpass filter was replaced with a 260 nm one, the onset energies in the Ek scale
shifted by 0.51 eV, but the electron binding energies Eb were unchanged (3.42 eV for 2PACz and
3.19 eV for MeO-2PACz), confirming that the spectral onsets originated from the IPES process,
namely, the transition from a free electron to the LUMO energy levels. The LEET spectra in Figure 3
show that the work functions (or the vacuum level) are 5.21 eV (2PACz) and 4.80 eV (MeO-
2PACz). Using these work functions, we determined EAs to be 1.73 eV (2PACz) and 1.48 eV
(MeO-2PACz) as the difference between the LEIPS onset and the vacuum level (FigureFig. 4).
1
FigureFIG. 3: . LEIPS (left) and LEET (right) spectra of 2PACz and MeO-2PACz measured by the
improved photon detector using the photomultiplier tube R821 and the bandpass filter with a center
The reported IEs are 5.6 eV [45]- eV45-–5.9 eV [46] eV45,46 for 2PACz and 5.1 eV [45] eV45 for
MeO-2PACz. The optical gaps are 3.5 eV (2PACz) and 3.2 eV (MeO-2PACz) [).[45,46].] These values
and the EAs determined in this work are summarized in FigureFig. 4. The transport gaps calculated
as the difference between IE and EA are 4.0–4.3 eV for 2PACz and 3.6 eV for MeO-2PACz. In
previous studies, the EA (or LUMO energy level) was estimated from the IE and the optical gap [.45-–
49]
. The estimated EA are systematically about 0.5 eV larger (or the LUMO level is lower) than the
LEIPS measurements. The transmission coefficient for electron tunneling or the carrier population at
the barrier is exponentially related to the energy difference [.50]. If we take the energy levels shown in
Fig. 4, the SAM thickness of 1 nm, the difference of 0.5 eV in barrier height leads to a four orders
of magnitude difference in the transmission coefficient for electron tunneling. The electron density in
the LUMO of the SAM layer can be up to 10 ten orders of magnitude different, assuming the
Boltzmann distribution (the values vary slightly depending on the energy parameters, but do not
affect the main conclusion). These estimates show that a difference of only 0.5 eV in barrier height
can make a significant difference in electron blocking capability, highlighting the importance of
1
We calculated the exciton binding energies from the difference between the transport and
optical gaps, and they were 0.5–0.8 eV for 2PACz and 0.4 eV for MeO-2PACz. We recently
reported that the exciton binding energy of organic semiconductors and polymers is approximately ¼
of the transport gap and 1/3 of the optical gap [.8]. Although we predicted an exciton binding energy
of approximately 1 eV, the observed exciton binding energies are smaller than this prediction
probably because a positive or negative charge is stabilized near the metallic electrodes owing to the
From the obtained EAs, we discuss the electron-blocking capability of SAMs. The energy
level alignment at the SAM–perovskite interface should lie between two limiting cases, namely, the
Fermi level alignment (the Schottky limit) and the vacuum level alignment (the Bardeen limit).
Figure 4 shows the conduction band (CB) minima and the valence band (VB) maxima with the Fermi
level alignment of representative perovskite CH3NH3PbI3 (MAPbI3) and wide bandwideband gap
(Cs0.2FA0.8PbI1.8Br1.2). The UPS and LEIPS of MAPbI3 measured at normal angles cannot access the
VB and CB edges (VB at -−1.55 eV and CB at 0.97 eV from the Fermi level, respectively
[51]
)respectively51) owing to band dispersion [.[52,53].] We show the corrected energy levels of MAPbI3
in Figure Fig. 4 [51].451.51 The LUMO energy levels of the SAMs are located more than 3 eV above
the CB of the both perovskite layers in the case of the Schottky limit. On the other hand, assuming a
vacuum level alignment, the energy barrier (difference between the LUMO energy level of SAM and
1
the CB of perovskite) becomes approximately 2 eV, which is still sufficiently high to efficiently
block the electrons at the SAM/perovskite also in the Bardeen limit. Because the band gaps of
perovskites for efficient PSCs are between 1 and 2 eV (theoretically, the best optical gap is 0.8–1.4
eV for the single cell and 1.7 eV and 1 eV for the double-layered tandem cell [56]),cell56) the LUMO
energy level of the present SAMs is expected to always be well above the CB of any perovskite layer
FigureFIG. 4: . Energy level diagram of carbazole-based SAMs 2PACz and MeO-2PACz, and
perovskites MAPbI3 [51-MAPbI351–53] and Cs0.2FA0.8PbI1.8Br1.2[.54,55]. The values (in eV) are with
reference to the Fermi level except for the electron affinities (numbers in white) and the optical
transition energies (numbers in black rotated upward by 90°°) and the ionization energies (numbers in
derivatives in a solid. The EA values of organic materials are best determined by LEIPS. However,
when the sample surface is irradiated with an electron beam for LEIPS measurement, the carbazole
derivatives emit strong fluorescence induced by electron impact in the visible range, which interferes
with the LEIPS measurement. We used a “solar-blind” photomultiplier tube with a Cs-Te
photocathode to separate the LEIPS signal from the fluorescence and observed the LEIPS spectra.
The present method applies to not only carbazole-based SAMs but also other emissive organic
semiconductors. We applied this technique to prototypical carbazole SAMs used as HSL and
1
determined their EAs to be 1.73 eV (2PACz) and 1.48 eV (MeO-2PACz). These EAs are smaller
than the previously reported values, implying that the carbazole-based SAMs have better electron
Acknowledgments
ACKNOWLEDGMENTS
We thank Mr. Daisuke Hamamura for his help in measuring UV-vis spectra and Prof.Professor
Kazuki Nakamura of Chiba University for allowing us to use the UV-vis spectrometer. This work
Data Availability
AUTHOR DECLARATIONS
Conflict of Interest
Author Contributions
1
Aruto Akatsuka: Conceptualization (supporting); Data curation (equal); Formal analysis (equal);
Writing – original draft (equal); Writing – review & editing (supporting). Makoto Miura: Data
curation (equal); Formal analysis (equal); Methodology (equal). Gaurav Kapil: Data curation
(lead); Writing – original draft (equal); Writing – review & editing (lead).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon
reasonable request
supporting data for this study are available from the corresponding author upon reasonable
request.
AUTHOR CONTRIBUTIONS
Author Contributions
HY, AA, and SH planned this work. MM and HY developed the photon detector. AA and MM
measured and analyzed the data. GK and SH prepared the samples. AA and HY wrote the
manuscript.
References
1
[1] NREL. , Best Research-Cell Efficiency Chart, 2024, see https://www.nrel.gov/pv/cell-
[2] [2]T. Zhang, T.,F. Wang, F.,H. Kim, H.,I. Choi, I.,C. Wang, C.,E. Cho, E.,R. Konefal, R.,Y.
Puttisong, Y.,K. Terado, K.,L. Kobera, L.,M. Chen, M.,. Yang, M.,S. Bai, S.,B. Yang, B.,J. Suo, J.,S.
Yang, S.,X. Liu, X.,F. Fu, F.,H. Yoshida, H.,W. M. Chen, W. M.,J. Brus, J.,V. Coropceanu, V.,A.
Hagfeldt, A.,J. Brédas, J.,M. Fahlman, M.,D. Kim, D.,Z. Hu, Z.,and F. Gao, F., “Ion-modulated
radical doping of spiro-OMeTAD for more efficient and stable perovskite solar cells.,” Science 377,
[3] [3]Y. Yao, Y.,C. Cheng, C.,. Zhang, C.,H. Hu, H.,K. Wang, K.,and S. De Wolf, S., “Organic hole-
transport layers for efficient, stable, and scalable inverted perovskite solar cells.,” Adv. Mater. 34,
[4] Wang, S.,. Wang, H. Guo, H.,and Y. Wu, Y., “Advantages and challenges of self-assembled
monolayer as a hole-selective contact for perovskite solar cells.,” Mater. Futures 2, 012105 (2023).
[5] A. Magomedov, A.,. Al-Ashouri, A.,E. Kasparavičius, E.,S. Strazdaite, S.,G. Niaura, G.,M. Jošt,
M.,T. Malinauskas, T.,S. Albrecht, S.,and V. Getautis, V., “Self-assembled hole transporting
monolayer for highly efficient perovskite solar cells,,” Adv. Energy Mater. 8, 1801892 (2018). DOI:
10.1002/aenm.201801892 [CrossRef][10.1002/aenm.201801892]
1
[6] I. Levine, I.,A. Al-Ashouri, A.,. Musiienko, A.,H. Hempel, H.,A. Magomedov, A.,.
Drevilkauskaite, A.,V. Getautis, V.,D. Menzel, D.,K. Hinrichs, K.,T. Unold, T.,S. Albrecht, S.,and T.
Dittrich, T., “Charge transfer rates and electron trapping at buried interfaces of perovskite solar
[10.1016/j.joule.2021.07.016][Mismatch] [InsertedFromOnline]
[7] [7]M. Stolterfoht, M.,P. Caprioglio, P.,C. M. Wolff, C. M.,J. A. Márquez, J. A., Nordmann, J.,S.
Zhang, S.,D. Rothhardt, D.,U. Hörmann, U.,Y. Amir, Y.,A. Redinger, A.,L. Kegelmann, L.,F. Zu,
F.,S. Albrecht, S.,N. Koch, N.,T. Kirchartz, T.,M. Saliba, M.,T. Unold, T.,and D. Neher, D., “The
impact of energy alignment and interfacial recombination on the internal and external open-circuit
voltage of perovskite solar cells,,” Energy Environ. Sci. 12, 2778-–2788 (2019).).[7],., -(). DOI:
10.1039/C9EE02020 [CrossRef][10.1039/C9EE02020A]
[8],.,. DOI: [8]A. Sugie, A.,K. Nakano, K.,. Tajima, K.,I. Osaka, I.,and H. Yoshida, H., “Dependence
of exciton binding energy on bandgap of organic semiconductors,,” J. Phys. Chem. Lett. 14, 11412–
[9] I. G. Hill, I.G.,A. Kahn, A.,Z. G. Soos, Z.G.,and R. A. Pascal, Jr, R.A., .,, “Charge-separation
energy in films of π-conjugated organic molecules,,” Chem. Phys. Lett.. 327, 181-–188 (2000). DOI:
10.1016/S0009-2614(00)00882-4 [CrossRef][10.1016/S0009-2614(00)00882-4][Mismatch]
1
[10] M. Knupfer, M., “Exciton binding energies in organic semiconductors,,” Appl. Phys. A 77,
[11] J. Sworakowski, J., “How accurate are energies of HOMO and LUMO levels in small-molecule
[10.1016/j.synthmet.2017.11.013]
[12] M. Kubo, M., and H. Yoshida, H., “Electron affinities of small-molecule organic
spectroscopy, and low-energy inverse photoelectron spectroscopy,,” Org. Electron. 108, 106551
[13] P. I. Djurovich, PE. I.,. Mayo, E. I.,S. R. Forrest, S. R.,and M. E. Thompson, M. E.,
[10.1016/j.orgel.2008.12.011][Mismatch]
[14] J. Sworakowski, J.,. Lipiński, J.,and K. Janus, K., “On the reliability of determination of
measurements. A simple picture based on the electrostatic model,,” Org. Electron. 33, 300-–310
1
[15] K. H. Frank, K. H.,P. Yannoulis, P.,R. Dudde, R.,and E. E. Koch, E. E., “Unoccupied molecular
orbitals of aromatic hydrocarbons adsorbed on Ag(111),),” J. Chem. Phys. 89, 7569- –7576 (1988).
[16] I. G. Hill, I.G.,A. Kahn, A.,J. Cornil, J.,D. A. dos Santos, D.A.,and J. L. Brédas, J.L., “Occupied
between experiment and theory,,” Chem. Phys. Lett. 317, 444-–450 (2000). DOI: 10.1016/S0009-
2614(99)01384-6 [CrossRef][10.1016/S0009-2614(99)01384-6]
[17] H. Yoshida, H.,K. Tsutsumi, K.,and N. Sato, N., “Unoccupied electronic states of 3d-transition
metal phthalocyanines (MPc: M=Mn, Fe, Co, Ni, Cu and Zn) studied by inverse photoemission
spectroscopy,,” J. Electron Spectros. Relat. Phenom. 121, 83-–91 (2001). DOI: 10.1016/S0368-
2048(01)00328-0 [CrossRef][10.1016/S0368-2048(01)00328-0]
[18] D. R. T. Zahn, D. R.T.,G. N. Gavrila, G. N.,and M. Gorgoi, M., “The transport gap of organic
semiconductors studied using the combination of direct and inverse photoemission,,” Chem. Phys.
[10.1016/j.chemphys.2006.02.003]
[19] S. Krause, S.,M. B. Casu, M. B.,A. Schöll, A.,and E. Umbach, E., “Determination of transport
levels of organic semiconductors by UPS and IPS,,” New J. Phys. 10, 085001 (2008). DOI:
10.1088/1367-2630/10/8/085001 [CrossRef][10.1088/1367-2630/10/8/085001][Mismatch]
1
[20] K. Kanai, K.,. Akaike, K.,. Koyasu, K.,. Sakai, K.,T. Nishi, T.,Y. Kamizuru, Y.,T. Nishi, T.,Y.
Ouchi, Y.,and K. Seki, K., “Determination of electron affinity of electron accepting molecules,,”
008-5021-1]
[21] H. Yoshida, H., “Near-ultraviolet inverse photoemission spectroscopy using ultra-low energy
[CrossRef][10.1016/j.cplett.2012.04.058]
[22] H. Yoshida, H., “Principle and application of low energy inverse photoemission spectroscopy: A
new method for measuring unoccupied states of organic semiconductors,,” J. Electron Spectrosc.
[23] T. Aihara, T.,S. Abd-rahman, S.,and H. Yoshida, H., “Metal screening effect on energy levels at
metal/organic interface: Precise determination of screening energy using photoelectron and inverse-
10.1103/PhysRevB.104.085305 [CrossRef][10.1103/PhysRevB.104.085305][Mismatch]
[24] K. Yamada, K.,S. Yanagisawa, S.,T. Koganezawa, T.,K. Mase, K.,N. Sato, N.,and H. Yoshida,
H., “Impact of the molecular quadrupole moment on ionization energy and electron affinity of
organic thin films: Experimental determination of electrostatic potential and electronic polarization
1
energies,,” Phys. Rev. B 97, 245206 (2018). DOI: 10.1103/PhysRevB.97.245206 [CrossRef]
[10.1103/PhysRevB.97.245206][Mismatch]
[25] Y. Uemura, Y.,S. A. Abd-Rahman, S. A., Yanagisawa, S.,and H. Yoshida, H., “Quantitative
analysis of the electrostatic and electronic polarization energies in molecularly mixed films of
[CrossRef][10.1103/PhysRevB.102.125302][Mismatch]
[26] S. A. Abd-Rahman, S. A.,T. Yamaguchi, T.,S. Kera, S.,and H. Yoshida, H., “Sample-shape
dependent energy levels in organic semiconductors,,” Phys. Rev. B 106, 075303 (2022). DOI:
10.1103/PhysRevB.106.075303 [CrossRef][10.1103/PhysRevB.106.075303][Mismatch]
[27] T. Sasaki, T.,M. Hasegawa, M.,K. Inagaki, K.,H. Ito, H.,K. Suzuki, K.,T. Oono, T.,K. Morii,
K.,T. Shimizu, T.,and H. Fukagawa, H., “Unravelling the electron injection/transport mechanism in
organic light-emitting diodes, Nature Comm.,” Nat. Commun. 12, 2706 (2021). DOI:
10.1038/s41467-021-23067-2 [CrossRef][10.1038/s41467-021-23067-2]
[28] K. Nakano, K.,Y. Chen, Y.,B. Xiao, B.,W. Han, W.,J. Huang, J.,H. Yoshida, H.,E. Zhou, E.,and
K. Tajima, K., “Anatomy of the energetic driving force for charge generation in organic solar cells,
Nature Comm.,” Nat. Commun. 10, 2520 (2019). DOI: 10.1038/s41467-019-10434-3 [CrossRef]
[10.1038/s41467-019-10434-3]
1
[29] J. Bertrandie, J.,. Han, C. S. PJ.,. De Castro, C. ES. P. Yengel, E.,J. Gorenflot, J.,T.
Anthopoulos, T.,F. Laquai, F.,A. Sharma, A.,and D. Baran, D., “The energy level conundrum of
organic semiconductors in solar cells,,” Adv. Mater. 34, 2202575 (2022). DOI:
10.1002/adma.202202575 [CrossRef][10.1002/adma.202202575][Mismatch]
[30] D. Lungwitz, D.,S. Joy, S.,A. E. Mansour, A. E., Opitz, A.,C. Karunasena, C.,H. Li, H.,N. A.
Panjwani, N. A., -K. Moudgil, K.,. Tang, K.,J. Behrends, J.,S. Barlow, S.,. R. Marder, S. R.,J.
Brédas, J.,K. Graham, K.,N. Koch, N.,and A. Kahn, A., “Spectral signatures of a negative polaron in
a doped polymer semiconductor: Energy levels and Hubbard U interactions,,” J. Phys. Chem. Lett.
[31] H. Sato, H.,S. A. Abd. Rahman, S. A.,Y. Yamada, Y.,H. Ishii, and H.,. Yoshida, H., “Conduction
band structure of high-mobility organic semiconductors and partially dressed polaron formation,
[32] J. Yang, J.,H. Sato, H.,. Orio, H.,X. Liu, X.,M. Fahlman, M.,N. Ueno, N.,H. Yoshida, H.,T.
Yamada, T.,and S. Kera, S., “Accessing the conduction band dispersion in CH3NH3PbI3 single
crystals,,” J. Phys. Chem. Lett. 12, 3773-–3778 (2021). DOI: 10.1021/acs.jpclett.1c00530 [CrossRef]
[10.1021/acs.jpclett.1c00530][Mismatch]
1
[33] H. Yoshida, H., “Low energy inverse photoemission spectroscopy apparatus,,” Rev. Sci.
[24517826]
[34] A. V. Kukhta, A. V., and S. M. Kazakov, S. M., “Spectroscopy of the interaction of a low-
energy electron beam with organic luminescent molecules,,” Phys.-Usp. 66, 173-–181 (2023). DOI:
10.3367/UFNe.2022.05.039197
phase in excitation by electrons and photons,,” J. Appl. Spectrosc. 68, 871-–876 (2001). DOI:
[36] C. Chen, C.,Z. Chi, Z.,K. C. Chong, K. CA. S. Batsanov, A. SZ. Yang, Z.,. Mao, Z.,. Yang,
Z.,and B. Liu, B., “Carbazole isomers induce ultralong organic phosphorescence,,” Nat. Mater. 20,
[37] J. B. Pendry, J. B., “New probe for unoccupied bands at surfaces,,” Phys. Rev. Lett. 45, 1356
[38] Hamamatsu Photonics K.K.,., Photomultiplier Tubes Basic and Applications, 4th ed.,
(Hamamatsu Photonics K.K. Electron Tube Division, Japan, 2017), pp. 246.
1
[39] V. Dose, “V., VUV isochromat spectroscopy,” Appl. Phys. 14, 117-–118 (1977). DOI:
10.1007/BF00882639
[40] C. Thiede, C.,I. Niehues, I.,A. B. Schmidt, A. B.,and M. Donath, M., “The acetone bandpass
6501/aab941]
[41] N. Babbe, N.,W. Drube, W.,I. Schafer, I.,and M. Skibowski, M., “A simple and compact system
for combined angular resolved inverse photoemission and photoemission in the vacuum ultraviolet,,”
[10.1088/0022-3735/18/2/014]
[42] I. Schäfer, I.,W. Drube, W.,M. Schlüter, M.,G. Plagemann, G.,and M. Skibowski, M.,
“Bandpass photon detector with high efficiency for inverse photoemission spectroscopy,,” Rev. Sci.
[43] Hamamatsu Photonics K.K., Photomultiplier Tubes R464, R585, 2016, see
https://www.hamamatsu.com/content/dam/hamamatsu-
photonics/sites/documents/99_SALES_LIBRARY/etd/R464_R585_TPMH1140E.pdf for
1
[44] Hamamatsu Photonics K.K., Photomultiplier Tubes R821, 1998, see
https://www.hamamatsu.com/content/dam/hamamatsu-photonics/sites/documents/
[45] A. Al-Ashouri, A.,. Magomedov, Ar.,M. Roß, M.,. Jošt, M.,. Talaikis, M.,G. Chistiakova, G.,T.
Bertram, T.,J. A. Márquez, J. A.,E. Köhnen, E.,. Kasparavičius, E.,S. Levcenco, S.,L. Gil-Escrig,
L.,C. J. Hages, C. J.,R. Schlatmann, R.,B. Rech, B.,T. Malinauskas, T.,. Unold, T.,C. A. Kaufmann,
C. A.,L. Korte, L.,G. Niaura, G.,V. Getautis, V.,and S. Albrecht, S., “Conformal monolayer contacts
with lossless interfaces for perovskite single junction and monolithic tandem solar cells,,” Energy
[10.1039/C9EE02268F]
[46] A. Al-Ashouri, A.,E. Köhnen, E.,B. Li, B.,A. Magomedov, A.,H. Hempel, H.,P. Caprioglio, P.,J.
A. Márquez, J. A.,. B. Morales Vilches, A. B.,E. Kasparavicius, E.,J. A. Smith, J. AN. Phung, N.,D.
Menzel, D.,M. Grischek, M.,L. Kegelmann, L.,D. Skroblin, D.,C. Gollwitzer, C.,T. Malinauskas,
T.,M. Jošt, M.,G. Matič, G.,B. Rech, B.,R. Schlatmann, R.,M. Topič, M.,L. Korte, L.,A. Abate, A.,B.
Stannowski, B.,D. Neher, D.,M. Stolterfoht, M.,T. Unold, T.,V. Getautis, V.,and S. Albrecht, S.,
“Monolithic perovskite/silicon tandem solar cell with >29% efficiency by enhanced hole
[10.1126/science.abd4016]
1
[47] T. H. Kim, TJ. H.,. Lee, JM. H.,. Jang, G. M. H., Lee, G. M.,E. S. Shim, E. S.,. Oh, S.,M. A.
Saeed, M. A.,J. Lee, M. J.,B. Yu, B.,D. K. Hwang, D. K.,C. W. Park, C. W.,S. Y. Lee, S. Y.,J. W. Jo,
and J. W.,. Shim, J. W., “Atto-scale noise near-infrared organic photodetectors enabled by controlling
interfacial energetic offset through enhanced anchoring ability,” Adv.,”. Mater. ▪, 2403647 (2024).
[48] M. Liu, M.,. Li, M.,Y. Li, Y.,. An, Y.,Z. Yao, Z.,B. Fan, B.,F. Qi, F.,K. Liu, K.,H. Yip, H.,F. R.
Lin, F. R., Jen,and A. K.-Y., . Jen, “Defect-passivating and stable benzothiophene-based self-
assembled monolayer for high-performance inverted perovskite solar cells,,” Adv. Energy Mater. 14,
[49] Zhang, H.,. Zhang, S.,. Zhang, X. Ji, X.,J. He, J.,H. Guo, H.,S. Wang, S.,W. Wu, W.,. H. Zhu,
W. H.,and Y. Wu, Y., “Formamidinium lead iodide-based inverted perovskite solar cells with
[50] S. M. Sze, S. M.,Y. Li, Y., Ng, Kwokand K.,. K. Ng, Physics of Semiconductor Devices, 4th ed.,
[51] A. Mirzehmet, A.,T. Ohtsuka, T.,S. A. Abd. Rahman, S. A.,T. Yuyama, T.,P. Krüger, P.,and H.
1
using electron spectroscopies,,” Adv. Mater. 33, 2004981 (2020). DOI: 10.1002/adma.202004981
[CrossRef][10.1002/adma.202004981]
[52] J. Endres, J.,D. A. Egger, D. A.,M. Kulbak, M.,R. A. Kerner, R. A.,L. Zhao, L.,S. H. Silver, S.
H.,G. Hodes, G.,B. P. Rand, B. P.,D. Cahen, D.,L. Kronik, L.,and A. Kahn, A., “Valence and
[CrossRef][10.1021/acs.jpclett.6b00946]
[53] F. Zu, F.,P. Amsalem, P.,D. A. Egger, D. AR. Wang, R.,C. M. Wolff, C. M.,H. Fang, H.,M. A.
Loi, M. A.,D. Neher, D.,L. Kronik, L.,S. Duhm, S.,and N. Koch, N., “Constructing the electronic
structure of CH3NH3PbI3 and CH3NH3PbBr3 perovskite thin films from single-crystal band structure
[CrossRef][10.1021/acs.jpclett.8b03728]
[54] H. Bi, H.,J. Liu, J.,Z. Zhang, Z.,L. Wang, L.,R. Beresneviciute, R.,D. Tavgeniene, D.,G. Kapil,
G.,C. Ding, C.,A. K. Baranwal, A. K.,S. R. Sahamir, S. R.,Y. Sanehira, Y.,H. Segawa, Hi.,S.
Grigalevicius, S.,Q. Shen, Q.,and S. Hayase, S., “All-perovskite tandem solar cells approach 26.5%
efficiency by employing wide bandgap lead perovskite solar cells with new monomolecular hole
[CrossRef][10.1021/acsenergylett.3c01275]
1
[55] H. Bi, H.,J. Liu, J.,R. Beresneviciute, R.,D. Tavgeniene, D.,Z. Zhang, Z.,L. Wang, L.,G. Kapil,
G.,C. Ding, C.,S. R. Sahamir, S. R.,Y. Sanehira, Y.,A. K. Baranwal, A. K.,T. Kitamura, T.,D. Wang,
D.,Y. Wei, Y.,. Yang, Y.,D. W. Kang, D. W.,S. Grigalevicius, S.,Q. Shen, Q.,and S. Hayase, S.,
“Efficiency enhancement of wide bandgap lead perovskite solar cells with PTAA surface-passivated
with monomolecular layer from the viewpoint of PTAA band bending,,” ACS Appl. Mater. Interfaces
[56] S. M. Sze, S. M.,Y. Li, Y., Ng, Kwokand K.,. K. Ng, Physics of Semiconductor Devices, 4th ed.,