Hybrid organic-inorganic materials are among the latest class of materials proposed for thermoelectric applications. The organic-inorganic interface is critical in determining the effective transport properties of the hybrid material. We study the thermoelectric properties of the tetrafluoro-tetracyanoquinodimethane (F4TCNQ)–silicon interface. Transfer of electrons from silicon to F4TCNQ results in holes trapped within the screening length of the interface that can move parallel to the interface. We measure the response of these trapped charges to applied temperature differential and compare the thermoelectric transport properties of the silicon with and without F4TCNQ. The results confirm the presence of interface charges and demonstrate an enhanced interface thermoelectric power factor. These outcomes of this study could be used in designing 3D hybrid structures with closely packed interfaces to replicate a bulk thermoelectric material.

Interfaces play a key role in semiconductor devices. The electron transport at conducting interfaces is of importance in the design of high-electron-mobility-transistors,1–3 superlattice structure devices,4–6 and energy conversion materials such as photovoltaics7–9 and thermoelectric materials.10–12 In particular, novel hybrid organic-inorganic materials have drawn increasing attention in research for diverse range of applications in energy, environment, and healthcare.13 

This work focuses on the charge transfer at an organic-inorganic interface, which shows promise in designing transfer-doped semiconductor nano-devices. Conventional doping methods (e.g., ion implantation) are oftentimes destructive processes introducing excess defects. The activation process ubiquitously requires solid-state thermal diffusion, which cannot be applied to materials that are heat sensitive. Devices with ultrafine structures would also suffer from non-negligible statistical variation or deactivation of dopants.14,15 The other main drawback of conventional doping methods is that they all involve incorporating aliovalent impurity ions, which create long range Coulomb potentials that scatter conduction electrons and lower their mobility. Modulation doping or remote doping has been proposed for 2D structures to overcome these shortcomings and has been implemented in transistor technology.16 Over 22 years (from 1978 to 2000), mobility enhancement by 4 orders of magnitude has been reported in GaAs, which was achieved by spatially separating parent atoms from conduction electrons.17 Another alternative doping strategy is surface doping of inorganic semiconducting materials by organic molecules, which was proposed fairly recently.18–21 Both in modulation doping and surface doping, charge carriers are confined to the interface of organic-inorganic materials and form a 2D electron gas (2DEG). To extend the advantages of 2DEG to 3D devices, one can envision forming superlattice structures by alternating organic-inorganic layers. Such hybrid superlattices have been theoretically evaluated for thermoelectric applications.22 Alternative 3D structures such as nanoparticles, nanowires, and holes embedded in host matrices, specifically for thermoelectric applications, have been proposed and discussed in detail.23–28 

In this work, we fabricated a device to investigate the charge transfer of the tetrafluoro-tetracyanoquinodimethane (F4TCNQ)–silicon interface. F4TCNQ is a fluorinated TCNQ derivative and has an exceptionally high electron affinity of Ea = 5.24 eV.29 It is well known as a strong electron acceptor, to dope molecules p-type by forming a charge transfer organic complex, as well as to enhance hole injection by energy level alignment at organic–metal interfaces.30,31 Therefore, F4TCNQ has been widely applied in organic light emitting diodes (OLEDs),32,33 photovoltaics,34,35 organic field-effect transistors (OFETs),36 etc. P-type surface doping was also achieved at inorganic semiconductor surfaces, such as diamond,37 graphene,38 and VO2,39 and investigated by photoemission spectroscopy (PES) techniques. However, as the most pervasive semiconductor material for electronic devices, Si has only sparsely been examined experimentally as the active layer for transfer doping by F4TCNQ. Existing reports mainly focus on electronic structure study by PES40–42 without details on the resultant transport properties. Therefore, we seek to investigate the electrical transport properties of the F4TCNQ-Si interface. The lowest unoccupied state (LUMO) of F4TCNQ lies below the silicon valence band maximum (VBM); hence, the electron affinity of F4TCNQ is larger than the ionization energy of Si [Fig. 1(a)]. This favors electron-transfer from silicon to F4TCNQ molecules, and Si is thus p-type doped near the interface. The holes are confined in the direction normal to the interface due to the established electrostatic potential but free to move in the parallel direction. XPS study shows that the shift of Si 2p binding energy peak saturates when the in situ deposited F4TCNQ thickness is 2 nm,43 meaning the F4TCNQ molecules farther to the interface do not contribute to the charge transfer, as illustrated in Fig. 1(b). Our previous calculation indicated that physisorbed F4TCNQ self-assembled monolayers can efficiently dope silicon, achieving hole concentrations as high as 1013 cm−2.44 

FIG. 1.

(a) The band structure of silicon and F4TCNQ before contact and (b) the device configuration studied in this work: buried oxide/Si/F4TCNQ. The charge transfer is only effective in proximity to the interface.

FIG. 1.

(a) The band structure of silicon and F4TCNQ before contact and (b) the device configuration studied in this work: buried oxide/Si/F4TCNQ. The charge transfer is only effective in proximity to the interface.

Close modal

A combination of enhanced mobility by transfer doping and low thermal conductivity of F4TCNQ could be utilized to design a high performance F4TCNQ–nanostructured Si (e.g., nanopatterned holey Si or nanowires) hybrid thermoelectric materials. However, this concept relies on a fundamental understanding of charge transfer and thermoelectric transport at the interface of F4TCNQ–silicon, which is thereby the main focus of this work.

In order to make a device to measure the interface properties, we built a silicon channel, as shown in Fig. 2. The device Si(100) layer (4″ SOI with 2 µm BOX) was thinned down to ∼100 nm by reactive ion etching (RIE). The subsequent blanket boron implantation yielded a mild hole concentration of ∼1016 cm−3 in the Si layer. Four side contact (5 µm × 20 µm) areas were then exposed by resist and underwent another boron implantation process, which yielded ∼1020 cm−3 only in the contact areas. The Si layer was etched down to BOX to form arrays of 30 µm × 200 µm Si mesas for the following metallization. 1 µm Al with a 50 nm Au cap was deposited as contacts for transport measurements. The device configuration is shown in Fig. 2. The heater, thermometers, and contacts are annotated. Finally, rapid thermal annealing was conducted at 500 °C for 30 s to achieve Ohmic contact.

FIG. 2.

(a) Device configuration for transport measurements and (b) 90° zoomed-in view of (a) with the Si device layer, heater, and thermometers annotated.

FIG. 2.

(a) Device configuration for transport measurements and (b) 90° zoomed-in view of (a) with the Si device layer, heater, and thermometers annotated.

Close modal

Next, we deposited F4TCNQ on the silicon device by thermal evaporation. The thermometers and the electrodes are then in between the silicon and F4TCNQ, allowing us to characterize the interface. F4TCNQ is insulating. Despite withdrawing electrons from Si upon contact, the electrical conduction of the F4TCNQ film is negligible compared to the Si layer. Therefore, it was feasible to directly deposit F4TCNQ onto the metallized device, without shorting the metal lines. Prior to the deposition, the device was dipped in buffered oxide etch (BOE) for 25 s to create a passive H-terminated Si surface, in order to prevent charge confinement by the surface states. The metal pads were then covered by a shadow mask for the ease of wire bonding after deposition. The F4TCNQ crystal powder (Sigma-Aldrich, 99.9%) and the device were loaded into a homebuilt miniature thermal evaporation chamber with a base pressure of <10−6 Torr. F4TCNQ was degassed below 100 °C and then sublimated at 140 °C for deposition. Surface morphology was characterized by atomic force microscopy (AFM). The XTEM sample with an F4TCNQ/Si interface was prepared by a focused ion beam (FIB) lift-out process, and the interface morphology was then characterized by transmission electron microscopy (TEM) along Si [110]. The electrical resistance was measured using the four-point probe method, both before and after depositing F4TCNQ. The Seebeck coefficient was measured in a homebuilt 2D transport measurement system incorporated in a Janis cryostat.

It was found that the F4TCNQ has a relatively low sticking probability on H–Si(100), which is commonly observed for molecules physically adsorbed on the inorganic substrate.45 Furthermore, due to the limited wettability, F4TCNQ adsorbates tend to coalesce into densely packed islands on Si. In order to achieve full coverage and maximize the charge transfer, a relatively high vapor flux is required to create a high nucleation rate. Therefore, we used a higher F4TCNQ temperature of ∼140 °C compared to other PES studies.40–42 The net deposition rate was calibrated as ∼1.5 Å/s. Figure 3 shows the surface morphology of F4TCNQ with a root mean square (RMS) roughness of ∼17 nm under this deposition condition.

FIG. 3.

Surface morphology of F4TCNQ with (a) 10 µm × 10 µm and (b) 2 µm × 2 µm by AFM.

FIG. 3.

Surface morphology of F4TCNQ with (a) 10 µm × 10 µm and (b) 2 µm × 2 µm by AFM.

Close modal

The interface is investigated by XTEM, as shown in Fig. 4. The active Si layer is ∼100 nm in thickness, on a 2 µm buried oxide layer, so that the subsequent transport measurements are minimally affected by the substrate. As shown in Fig. 4(a), continuous coverage of F4TCNQ on Si was achieved, with an average thickness of ∼150 nm. In essence, only the first few layers of the molecule would contribute to the charge transfer, leaving the majority as neutral F4TCNQ bulk. Figure 4(b) is the high-resolution image along the Si [110] axis. It exhibits a smooth interface without observable voids or interlayers. Therefore, an effective charge transfer can be expected across the interface. Nevertheless, the orientation of molecules on the Si surface is still unclear, which could affect the transfer doping efficiency.44 Future work using STM46 or PES41 techniques may shed light on orientation of the molecules.

FIG. 4.

(a) XTEM of the device by FIB lift-out, with each layer annotated; (b) high-resolution TEM image of the F4TCNQ–Si interface along Si [110]. The inset shows the convergent-beam electron diffraction (CBED) pattern.

FIG. 4.

(a) XTEM of the device by FIB lift-out, with each layer annotated; (b) high-resolution TEM image of the F4TCNQ–Si interface along Si [110]. The inset shows the convergent-beam electron diffraction (CBED) pattern.

Close modal

In order to evaluate the charge transfer, and its effect on power factor, we performed electrical transport measurements. I-V measurements by the 4-probe method were conducted first on the mildly p-doped, uncoated Si device and then on the same device deposited with F4TCNQ. As shown in Fig. 5, Ohmic IV curves were achieved in both cases. The measured resistance of F4TCNQ–Si (401 kΩ) is 10 times lower compared to the resistance of a plain Si device (4.36 MΩ). The result provides direct evidence that charge transfer at the interface occurred and that silicon is surface doped by the physically adsorbed F4TCNQ.

FIG. 5.

Comparison of IV curves for Si and F4TCNQ–Si devices. The inset is the optical image of the device partially deposited with F4TCNQ using a shadow mask.

FIG. 5.

Comparison of IV curves for Si and F4TCNQ–Si devices. The inset is the optical image of the device partially deposited with F4TCNQ using a shadow mask.

Close modal

The Seebeck coefficient was then measured using a microheater and two calibrated thermometers, as shown in Fig. 2. When applying a current to the heater, a temperature difference is built up across the two ends of the Si mesa as a result of Joule heating inside the heater. Therefore, an electrical potential is produced due to the Seebeck effect. Since the thermometers are in direct contact with the Si surface, the Seebeck voltage could thereby be measured. We observe a large drop in the Seebeck voltage from −208.9 ± 12.3 µV for mildly doped, uncoated Si to −73.9 ± 3.7 µV after deposition of F4TCNQ [see the steady state voltage shown in Fig. 6(a)]. In order to extrapolate the actual Seebeck coefficient (S = VSΔT), the corresponding temperature differential must be measured. The temperature of the device could be extracted from the temperature-dependent resistance of the thermometers. Separate temperature calibrations were conducted for each thermometer. The sample’s temperature was raised using an external heater in the sample holder, and the resistance of the thermometers was measured versus temperature. A minor increase in thermometer resistance after depositing F4TCNQ was noted due to the presence of the molecules on the gold lines, and therefore their resistances were recalibrated afterwards.

FIG. 6.

(a) Comparison of the Seebeck voltage for Si and F4TCNQ–Si devices and [(b)–(e)] resistance curves of each thermometer when the heater is on/off.

FIG. 6.

(a) Comparison of the Seebeck voltage for Si and F4TCNQ–Si devices and [(b)–(e)] resistance curves of each thermometer when the heater is on/off.

Close modal

During the Seebeck measurement, when the heater turns on, the resistance (by the 4-point method) of the thermometers ramps up, as shown in Figs. 6(b)6(e). Using the calibration curves, we could then extrapolate ΔT(TC1TC2), which reached to a steady state after ∼20 s, corresponding to the Seebeck voltage plateau. The Seebeck coefficient determined by this approach for the uncoated Si device is 594.6 ± 38.0 µV/K and for the F4TCNQ–Si device is 243 ± 12.3 µV/K. The sign of Seebeck coefficients confirms that both devices are p-type doped. It was found that the power factor (S2/R) of silicon had a 75% enhancement after depositing F4TCNQ.

Since the screening length of surface-doped holes is estimated to be less than 10 nm, the majority of the coated device is not contributing to the conductance. Given that the resistance had 10 times’ reduction, and carrier mobility should be unchanged in F4TCNQ–Si, an ∼100 times enhancement in hole concentration can be estimated in the modulation doped region in proximity to the interface. The obtained Seebeck value of 243 ± 12.3 µV/K for the F4TCNQ–Si sample could be interpreted as the Seebeck coefficient of the 2D hole gas since the transferred charge accumulated near the surface is the dominant contributor to the Seebeck voltage. The thermal conductivity is not affected, and we estimated a similar thermal resistance with and without F4TCNQ since in this case, measurements are performed for the interface and the lattice thermal conductance of silicon is the dominant heat transfer mechanism.

These results serve as a proof of concept of the potential of organic–Si interfaces with large charge transfer for power factor optimization for thermoelectric device application. The issue of limited conductance enhancement due to the short screening length brings in the necessity of having nanostructured Si with a high surface-to-volume ratio (e.g., nanopatterned holey Si), in which case the transfer doping can be maximized, and the necking size is comparable to the screening length of transferred charges so as to freely transport throughout the material. Future study on manipulating the molecule orientation and surface treatment to minimize surface states as a charge trap is expected to further enhance the power factor.

The authors gratefully acknowledge Petra Reinke, Nathan Swami, Arthur Lichtenberger, and Kyusang Lee for insightful discussions. This work was supported by the National Science Foundation under Grant No. 1723353 (N.L. and M.Z.).

1.
M.
Asif Khan
,
A.
Bhattarai
,
J. N.
Kuznia
, and
D. T.
Olson
, “
High electron mobility transistor based on a GaN–AlxGa1–xN heterojunction
,”
Appl. Phys. Lett.
63
,
1214
1215
(
1993
).
2.
Y.
Zhang
and
J.
Singh
, “
Charge control and mobility studies for an AlGaN/GaN high electron mobility transistor
,”
J. Appl. Phys.
85
,
587
594
(
1999
).
3.
Y.
Hori
,
Z.
Yatabe
, and
T.
Hashizume
, “
Characterization of interface states in Al2O3/AlGaN/GaN structures for improved performance of high-electron-mobility transistors
,”
J. Appl. Phys.
114
,
244503
(
2013
).
4.
R. M.
Feenstra
,
D. A.
Collins
,
D. Z. Y.
Ting
,
M. W.
Wang
, and
T. C.
McGill
, “
Interface roughness and asymmetry in InAs/GaSb superlattices studied by scanning tunneling microscopy
,”
Phys. Rev. Lett.
72
,
2749
2752
(
1994
).
5.
S. J.
Haigh
,
A.
Gholinia
,
R.
Jalil
,
S.
Romani
,
L.
Britnell
,
D. C.
Elias
,
K. S.
Novoselov
,
L. A.
Ponomarenko
,
A. K.
Geim
, and
R.
Gorbachev
, “
Cross-sectional imaging of individual layers and buried interfaces of graphene-based heterostructures and superlattices
,”
Nat. Mater.
11
,
764
767
(
2012
).
6.
I.
Chowdhury
,
R.
Prasher
,
K.
Lofgreen
,
G.
Chrysler
,
S.
Narasimhan
,
R.
Mahajan
,
D.
Koester
,
R.
Alley
, and
R.
Venkatasubramanian
, “
On-chip cooling by superlattice-based thin-film thermoelectrics
,”
Nat. Nanotechnol.
4
,
235
238
(
2009
).
7.
C.
Goh
,
S. R.
Scully
, and
M. D.
McGehee
, “
Effects of molecular interface modification in hybrid organic-inorganic photovoltaic cells
,”
J. Appl. Phys.
101
,
114503
(
2007
).
8.
H.
Zhou
,
Q.
Chen
,
G.
Li
,
S.
Luo
,
T.-b.
Song
,
H.-S.
Duan
,
Z.
Hong
,
J.
You
,
Y.
Liu
, and
Y.
Yang
, “
Interface engineering of highly efficient perovskite solar cells
,”
Science
345
,
542
546
(
2014
).
9.
P.
Schulz
,
E.
Edri
,
S.
Kirmayer
,
G.
Hodes
,
D.
Cahen
, and
A.
Kahn
, “
Interface energetics in organo-metal halide perovskite-based photovoltaic cells
,”
Energy Environ. Sci.
7
,
1377
1381
(
2014
).
10.
D. L.
Medlin
and
G. J.
Snyder
, “
Interfaces in bulk thermoelectric materials. A review for current opinion in colloid and interface science
,”
Curr. Opin. Colloid Interface Sci.
14
,
226
235
(
2009
).
11.
Y.
Sungtaek Ju
and
U.
Ghoshal
, “
Study of interface effects in thermoelectric microrefrigerators
,”
J. Appl. Phys.
88
,
4135
(
2000
).
12.
M. S.
Dresselhaus
,
G.
Chen
,
M. Y.
Tang
,
R. G.
Yang
,
H.
Lee
,
D. Z.
Wang
,
Z. F.
Ren
,
J.-P.
Fleurial
, and
P.
Gogna
, “
New directions for low-dimensional thermoelectric materials
,”
Adv. Mater.
19
,
1043
1053
(
2007
).
13.
Hybrid Organic-Inorganic Interfaces : Towards Advanced Functional Materials
, edited by
M. H.
Delville
and
A.
Taubert
(
Wiley-VCH Verlag GmBh
,
2017
).
14.
K. J.
Rietwyk
,
Y.
Smets
,
M.
Bashouti
,
S. H.
Christiansen
,
A.
Schenk
,
A.
Tadich
,
M. T.
Edmonds
,
J.
Ristein
,
L.
Ley
, and
C. I.
Pakes
, “
Charge transfer doping of silicon
,”
Phys. Rev. Lett.
112
,
155502
(
2014
).
15.
M. T.
Björk
,
H.
Schmid
,
J.
Knoch
,
H.
Riel
, and
W.
Riess
, “
Donor deactivation in silicon nanostructures
,”
Nat. Nanotechnol.
4
,
103
107
(
2009
).
16.
R.
Dingle
,
H. L.
Störmer
,
A. C.
Gossard
, and
W.
Wiegmann
, “
Electron mobilities in modulation-doped semiconductor heterojunction superlattices
,”
Appl. Phys. Lett.
33
,
665
(
1978
).
17.
L.
Pfeiffer
and
K.
West
, “
The role of MBE in recent quantum Hall effect physics discoveries
,”
Physica E
20
,
57
64
(
2003
).
18.
J.
Ristein
, “
Surface transfer doping of semiconductors
,”
Science
313
,
1057
1058
(
2006
).
19.
W.
Chen
,
D.
Qi
,
X.
Gao
, and
A. T. S.
Wee
, “
Surface transfer doping of semiconductors
,”
Prog. Surf. Sci.
84
,
279
321
(
2009
).
20.
D.
Lee
,
S. Y.
Sayed
,
S.
Lee
,
C. A.
Kuryak
,
J.
Zhou
,
G.
Chen
, and
Y.
Shao-Horn
, “
Quantitative analyses of enhanced thermoelectric properties of modulation-doped PEDOT:PSS/undoped Si (001) nanoscale heterostructures
,”
Nanoscale
8
,
19754
19760
(
2016
).
21.
D.
Lee
,
J.
Zhou
,
G.
Chen
, and
Y.
Shao-Horn
, “
Enhanced thermoelectric properties for PEDOT:PSS/undoped Ge thin-film bilayered heterostructures
,”
Adv. Electron. Mater.
(published online).
22.
J.
Carrete
,
N.
Mingo
,
G.
Tian
,
H.
Ågren
,
A.
Baev
, and
P. N.
Prasad
, “
Thermoelectric properties of hybrid organic–inorganic superlattices
,”
J. Phys. Chem. C
116
,
10881
10886
(
2012
).
23.
B.
Yu
,
M.
Zebarjadi
,
H.
Wang
,
K.
Lukas
,
H.
Wang
,
D.
Wang
,
C.
Opeil
,
M.
Dresselhaus
,
G.
Chen
, and
Z.
Ren
, “
Enhancement of thermoelectric properties by modulation-doping in silicon germanium alloy nanocomposites
,”
Nano Lett.
12
,
2077
2082
(
2012
).
24.
M.
Zebarjadi
,
G.
Joshi
,
G.
Zhu
,
B.
Yu
,
A.
Minnich
,
Y.
Lan
,
X.
Wang
,
M.
Dresselhaus
,
Z.
Ren
, and
G.
Chen
, “
Power factor enhancement by modulation doping in bulk nanocomposites
,”
Nano Lett.
11
,
2225
2230
(
2011
).
25.
M.
Zebarjadi
,
G.
Chen
,
Z.
Ren
,
S.
Shin
,
R.
Chen
,
J. P.
Heremans
,
B.
Wiendlocha
,
H.
Jin
,
B.
Wang
, and
Q.
Zhang
, “
Engineering of materials
,” in
Advanced Thermoelectrics: Materials, Contacts, Devices and Systems
, edited by
Z.
Ren
,
Y.
Lan
, and
Q.
Zhang
(
CRC Press
,
2018
).
26.
E.
Song
,
Q.
Li
,
B.
Swartzentruber
,
W.
Pan
,
G. T.
Wang
, and
J. A.
Martinez
, “
Enhanced thermoelectric transport in modulation-doped GaN/AlGaN core/shell nanowires
,”
Nanotechnology
27
,
15204
(
2016
).
27.
H.
Wang
,
X.
Cao
,
Y.
Takagiwa
, and
G. J.
Snyder
, “
Higher mobility in bulk semiconductors by separating the dopants from the charge-conducting band–a case study of thermoelectric PbSe
,”
Mater. Horiz.
2
,
323
329
(
2015
).
28.
J.-H.
Bahk
,
Z.
Bian
,
M.
Zebarjadi
,
P.
Santhanam
,
R.
Ram
, and
A.
Shakouri
, “
Thermoelectric power factor enhancement by ionized nanoparticle scattering
,”
Appl. Phys. Lett.
99
,
72118
(
2011
).
29.
W.
Gao
and
A.
Kahn
, “
Electronic structure and current injection in zinc phthalocyanine doped with tetrafluorotetracyanoquinodimethane: Interface versus bulk effects
,”
Org. Electron.
3
,
53
63
(
2002
).
30.
N.
Koch
,
S.
Duhm
,
J. P.
Rabe
,
A.
Vollmer
, and
R. L.
Johnson
, “
Optimized hole injection with strong electron acceptors at organic-metal interfaces
,”
Phys. Rev. Lett.
95
,
237601
(
2005
).
31.
S.
Braun
,
W. R.
Salaneck
, and
M.
Fahlman
, “
Energy-level alignment at organic/metal and organic/organic interfaces
,”
Adv. Mater.
21
,
1450
1472
(
2009
).
32.
J.
Blochwitz
,
M.
Pfeiffer
,
T.
Fritz
, and
K.
Leo
, “
Low voltage organic light emitting diodes featuring doped phthalocyanine as hole transport material
,”
Appl. Phys. Lett.
73
,
729
731
(
1998
).
33.
J.
Huang
,
M.
Pfeiffer
,
A.
Werner
,
J.
Blochwitz
,
K.
Leo
, and
S.
Liu
, “
Low-voltage organic electroluminescent devices using pin structures
,”
Appl. Phys. Lett.
80
,
139
141
(
2002
).
34.
K.-H.
Yim
,
G. L.
Whiting
,
C. E.
Murphy
,
J. J. M.
Halls
,
J. H.
Burroughes
,
R. H.
Friend
, and
J.-S.
Kim
, “
Controlling electrical properties of conjugated polymers via a solution-based P-type doping
,”
Adv. Mater.
20
,
3319
3324
(
2008
).
35.
D.
Liu
,
Y.
Li
,
J.
Yuan
,
Q.
Hong
,
G.
Shi
,
D.
Yuan
,
J.
Wei
,
C.
Huang
,
J.
Tang
, and
M. K.
Fung
, “
Improved performance of inverted planar perovskite solar cells with F4-TCNQ doped PEDOT:PSS hole transport layers
,”
J. Mater. Chem. A
5
,
5701
5708
(
2017
).
36.
J.
Soeda
,
Y.
Hirose
,
M.
Yamagishi
,
A.
Nakao
,
T.
Uemura
,
K.
Nakayama
,
M.
Uno
,
Y.
Nakazawa
,
K.
Takimiya
, and
J.
Takeya
, “
Solution-crystallized organic field-effect transistors with charge-acceptor layers: High-mobility and low-threshold-voltage operation in air
,”
Adv. Mater.
23
,
3309
3314
(
2011
).
37.
D.
Qi
,
W.
Chen
,
X.
Gao
,
L.
Wang
,
S.
Chen
,
P. L.
Kian
, and
A. T. S.
Wee
, “
Surface transfer doping of diamond (100) by tetrafluoro-tetracyanoquinodimethane
,”
J. Am. Chem. Soc.
129
,
8084
8085
(
2007
).
38.
W.
Chen
,
S.
Chen
,
C. Q.
Dong
,
Y. G.
Xing
, and
A. T. S.
Wee
, “
Surface transfer P-type doping of epitaxial graphene
,”
J. Am. Chem. Soc.
129
,
10418
10422
(
2007
).
39.
K.
Wang
,
W.
Zhang
,
L.
Liu
,
P.
Guo
,
Y.
Yao
,
C. H.
Wang
,
C.
Zou
,
Y. W.
Yang
,
G.
Zhang
, and
F.
Xu
, “
Holes doping effect on the phase transition of VO2 film via surface adsorption of F4TCNQ molecules
,”
Appl. Surf. Sci.
447
,
347
354
(
2018
).
40.
K.
Mukai
and
J.
Yoshinobu
, “
Observation of charge transfer states of F4-TCNQ on the 2-methylpropene chemisorbed Si(100)(2 × 1) surface
,”
J. Electron Spectrosc. Relat. Phenom.
174
,
55
58
(
2009
).
41.
M.
Furuhashi
and
J.
Yoshinobu
, “
Charge transfer and molecular orientation of tetrafluoro-tetracyanoquinodimethane on a hydrogen-terminated Si(111) surface prepared by a wet chemical method
,”
J. Phys. Chem. Lett.
1
,
1655
1659
(
2010
).
42.
S.
Yoshimoto
,
M.
Furuhashi
,
T.
Koitaya
,
Y.
Shiozawa
,
K.
Fujimaki
,
Y.
Harada
,
K.
Mukai
, and
J.
Yoshinobu
, “
Quantitative analysis of chemical interaction and doping of the Si(111) native oxide surface with tetrafluoro-tetracyanoquinodimethane
,”
J. Appl. Phys.
115
,
143709
(
2014
).
43.
G. D.
Yuan
,
T. W.
Ng
,
Y. B.
Zhou
,
F.
Wang
,
W. J.
Zhang
,
Y. B.
Tang
,
H. B.
Wang
,
L. B.
Luo
,
P. F.
Wang
,
I.
Bello
 et al., “
p-type conductivity in silicon nanowires induced by heterojunction interface charge transfer
,”
Appl. Phys. Lett.
97
,
153126
(
2010
).
44.
X.
Wang
,
K.
Esfarjani
, and
M.
Zebarjadi
, “
First-principles calculation of charge transfer at the silicon-organic interface
,”
J. Phys. Chem. C
121
,
15529
(
2017
).
45.
M. A.
Leitch-Devlin
and
D. A.
Williams
, “
Sticking coefficients for atoms and molecules at the surfaces of interstellar dust grains
,”
Mon. Not. R. Astron. Soc.
213
,
295
306
(
1985
).
46.
H. Z.
Tsai
,
A. A.
Omrani
,
S.
Coh
,
H.
Oh
,
S.
Wickenburg
,
Y. W.
Son
,
D.
Wong
,
A.
Riss
,
H. S.
Jung
,
G. D.
Nguyen
 et al., “
Molecular self-assembly in a poorly screened environment: F4TCNQ on graphene/BN
,”
ACS Nano
9
,
12168
12173
(
2015
).