Functionalized Nickel Oxide Hole Contact Layers: Work Function

Functionalized Nickel Oxide Hole Contact Layers: Work Function versus Conductivity. Sebastian Hietzschold. 1,2,3,*. , Sabina Hillebrandt. 1,3. , Flori...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Florida | Smathers Libraries

Article

Functionalized Nickel Oxide Hole Contact Layers: Work Function versus Conductivity Sebastian Hietzschold, Sabina Hillebrandt, Florian Ullrich, Jakob Bombsch, Valentina Rohnacher, Shuangying Ma, Wenlan Liu, Andreas Köhn, Wolfram Jaegermann, Annemarie Pucci, Wolfgang Kowalsky, Eric Mankel, Sebastian Beck, and Robert Lovrincic ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.7b12784 • Publication Date (Web): 20 Oct 2017 Downloaded from http://pubs.acs.org on October 21, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Functionalized Nickel Oxide Hole Contact Layers: Work Function versus Conductivity Sebastian Hietzschold1,2,3,*, Sabina Hillebrandt1,3, Florian Ullrich1,4, Jakob Bombsch1,3, Valentina Rohnacher1,3, Shuangying Ma1,6, Wenlan Liu1,6, Andreas Köhn1,6, Wolfram Jaegermann1,4, Annemarie Pucci1,3,5, Wolfgang Kowalsky1,2, Eric Mankel1,4, Sebastian Beck1,3, and Robert Lovrincic1,2,* 1

InnovationLab, Speyerer Str. 4, 69115 Heidelberg, Germany

2

Institute for High-Frequency Technology, TU Braunschweig, Schleinitzstr. 22, 38106 Braunschweig, Germany 3

Kirchhoff-Institute for Physics, Heidelberg University, Im Neuenheimer Feld 227, 69120 Heidelberg, Germany 4

Surface Science Division, TU Darmstadt, Jovanka-Bontschits-Str. 2, 64287 Darmstadt, Germany 5

Centre for Advanced Materials, Heidelberg University, Im Neuenheimer Feld, 69120 Heidelberg, Germany 6

Institute for Theoretical Chemistry, University of Stuttgart, Pfaffenwaldring 55, 70569 Stuttgart, Germany *Corresponding authors, E-mail: [email protected], [email protected]

Keywords: metal oxides, self-assembled monolayers, hybrid interfaces, solar cells, density functional theory

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 27

Abstract Nickel oxide is a widely-used material for efficient hole extraction in opto-electronic devices. However, its surface characteristics strongly depend on the processing history and exposure to adsorbates. To achieve controllability of the electronic and chemical properties of solutionprocessed nickel oxide (sNiO), we functionalize its surface with a self-assembled monolayer (SAM) of 4-cyanophenylphosphonic acid. A detailed analysis of infrared and photoelectron spectroscopy shows the chemisorption of the molecules with a nominal layer thickness of around one monolayer and gives insight into the chemical composition of the SAM. DFT calculations reveal possible binding configurations. By the application of the SAM, we increase the sNiO work function by up to 0.8 eV. When incorporated in organic solar cells, the increase in work function and improved energy level alignment to the donor does not lead to a higher fill factor of these cells. Instead, we observe the formation of a transport barrier, which can be reduced by increasing the conductivity of the sNiO through doping with copper oxide. We conclude that the widespread assumption of maximizing the fill factor by only matching the work function of the oxide charge extraction layer with the energy levels in the active material is a too narrow approach. Successful implementation of interface modifiers is only possible with a sufficiently high charge carrier concentration in the oxide interlayer to support efficient charge transfer across the interface.

ACS Paragon Plus Environment

2

Page 3 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Introduction Transition metal oxides (TMOs) are commonly used in their non-stoichiometric phase as efficient charge selective transport materials in opto-electronic devices.1,2 Nickel oxide (NiO) is a particularly promising candidate for hole injection and extraction,3–8 as it combines p-type conductivity, high transparency, and a low electron affinity (EA).9–21 However, the work function of NiO films depends on their crystalline orientation, surface composition, processing history, and is not always sufficiently high to align well with the frontier orbitals of organic donor materials with high ionization potential (IP) for ohmic contact. It has previously been shown that a mixed NiO/MoO3 layer can overcome this problem and increase the built-in voltage and thereby the fill factor (FF) in organic solar cells.22 Another convenient way to increase the work function of NiO is an oxygen plasma treatment,23 but its adjustment range is limited. A more variable approach is the use of dipolar self-assembled monolayers (SAMs),24–29 with which the work function30 and also the surface energy (wettability) can be regulated by tailoring the backbone and functional group of the interface modifier.31–33 Phosphonic acids (PA) are known to chemisorb and form highly dense monolayers on hydroxylated oxide surfaces via a mono-, bior tridentate bonding mode.26,34,35 While there are reports on molecular modification with phosphonic acids on nickel foil36, e-beam evaporated NiO37 and pulsed laser deposited Ni:CoO38, we are aware of only one short report of SAM deposition on solution-processed NiO (sNiO).39 For these reasons it is of high interest to investigate the formation of SAMs on sNiO surfaces and its impact on device performance in more detail. Here, we show that phosphonic acids indeed chemisorb on sNiO surfaces and form a SAM. We use the commercially available 4-cyanophenylphosphonic acid (CYNOPPA) as SAM precursor, as it possesses a strong permanent dipole moment to increase the work function.31 Furthermore,

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 27

its well conjugated backbone is thought to enable efficient charge transport40 and the cyano group allows to improve the wettability for printing applications (see Figure S8). Through the growth of a CYNOPPA monolayer on the sNiO, we successfully increase the work function by up to 0.8 eV. However, despite the proper energy level alignment to the donor, the low intrinsic charge carrier density in the sNiO in combination with the SAM leads to a transport barrier in organic solar cells. This barrier can be reduced by increasing the conductivity of the sNiO HTL with copper oxide (CuO). While oxide contact layers are often treated like metals in terms of energy level alignment,13,41 our work highlights that the semiconducting properties of transition metal oxides, e.g. doping level and band bending, are still important for proper interface engineering. This differentiated consideration is crucial for the enhancement of charge transport across hybrid interfaces in general.

Results and discussion sNiO films of ~25 nm thickness were fabricated according to Manders et al.42 by spin-coating nickel acetate tetrahydrate from an ethanol solution and subsequent annealing in air at 325 °C and 400 °C, respectively (see methods for details). At these temperatures the precursor was shown to be decomposed to a high degree.43 The resulting films show excellent substrate coverage and consist of small grains with a typical size of 10 nm. High resolution AFM and SEM images revealing this grainy surface structure are attached in the supporting information (see Figure S14). To gain insight into the growth mechanism of CYNOPPA on the sNiO surface a complementary set of infrared and photoelectron spectroscopy measurements were performed. CYNOPPA was

ACS Paragon Plus Environment

4

Page 5 of 27

applied to the surface as suggested elsewhere.31 Figure 1 shows infrared transmission spectra of CYNOPPA on sNiO films for sNiO annealed at two different temperatures. As reference, the exact same substrates were measured before and after the deposition of the SAM to get as detailed information as possible. The angle of incidence (AOI) was set to 10°, which we correspond to as normal incidence. Various absorption bands of CYNOPPA arise in the spectra on both sNiO layers. Modes marked with black dashed lines are assigned to vibrational modes of the molecule with the help of DFT calculations (see Figure S1 and Table S1).

α-Ni(OH)2

C N

rel. transmission

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

β-Ni(OH)2

325°C-sNiO/CYN 400°C-sNiO/CYN 1000

1250

1500

1750

2000

0.2% 2250

wave number [cm-1]

3500

4000

Figure 1. Mid-infrared measurements of CYNOPPA on 325 °C (blue) and 400 °C (red) annealed sNiO. The sNiO substrate prior treatment was used as a reference. The characteristic absorption bands of CYNOPPA are marked with black dashed lines. At 833 cm-1, 966 cm-1, and 1250 cm-1 absorption bands of the phosphonate group, at 1130 cm-1 the stretching vibration of the phenyl ring and at 2231 cm-1 the absorption of the (C≡N) cyano group stretching vibration are observed. The reduction of β-Ni(OH)2 through the SAM treatment is marked with a light grey dashed line. Characteristic modes of α-Ni(OH) have been highlighted with grey diagonally striped regions. At 1130 cm-1 the stretching vibration of the phenyl ring is present. Another characteristic mode can be seen at 2231 cm-1, which is assigned to the stretching vibration of the cyano group (C≡N). The weak vibrational modes at 833 cm-1, 966 cm-1, and 1250 cm-1 are attributed to absorption bands of the phosphonate group. The existence of all these characteristic peaks is a clear evidence that CYNOPPA is present at the surface. Additional positive absorption bands are observed in the spectral range between 1280 and 1700 cm-1 and between 3500 and 3700 cm-1.

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 27

These features are assigned to changes in the sNiO layer, as discussed in the following. Incomplete conversion of the precursor leads to residual nickel hydroxide (Ni(OH)2), which exhibits absorption bands in this range.44,45 The very pronounced upwards peak at 3670 cm-1 can be attributed to a reduction of β-Ni(OH)2 and the broad mode around 3500 cm-1 implies a change in α-Ni(OH)2. The weak signals of the phosphonate group, especially the broadened P=O mode, and the reduction of Ni(OH)2 indicate a chemical interaction of the molecules with the sNiO. Furthermore, the overall intensity of the vibrational modes of the molecule, which lies below 0.4%, is in good accordance with a monolayer formation. The C≡N stretching vibration can directly be related to the amount of molecules if there is no orientation change in the two compared samples. As the peak height does not vary under different angle of incidence (see Figure S2), a smaller amount of CYNOPPA is present on sNiO films annealed at higher temperatures. This goes along with a decreasing amount of Ni(OH)2 in the films with increasing annealing temperature of the NiO precursor.43,46 Both experimental observations allow us to conclude that, here, the phosphonic acid binds mostly to the nickel hydroxide at the sNiO surface. To evaluate possible binding configurations, we employed density functional theory (DFT) calculations on a fully hydroxylated ideal NiO(111) surface considering a series of adsorption modes. A comparison of the adsorption energies (see Figure S15) shows that the reactions leading to adsorption modes with more than one water molecule involved in the dehydration process are energetically unfavorable. This is supported by ab initio molecular dynamics simulations which demonstrate that only mono- and bidentate binding modes are stable under the conditions of the surface treatment (ethanol solution). The purely hydrogen bonded mode will not be stable under the conditions of the CYNOPPA annealing step, which is introduced to

ACS Paragon Plus Environment

6

Page 7 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

activate the phosphonate bonding. In the absence of protic solvent molecules to reverse the bonding process the other binding modes become increasingly stable. Hence, the most probable adsorption mode is a monodentate binding between the dehydrated CYNOPPA and the NiO surface. The formation of a bidentate binding mode might be possible, too, but the tridentate bonding, which would lead to a strong net orientation towards the surface normal, is not accessible.

Figure 2. Background corrected and normalized XP detail spectra a) Ni 2p3/2 and b) O 1s core level spectra for 325 °C (blue) and 400 °C (red) annealed sNiO films with and without CYNOPPA treatment. Spectra were shifted to the maximum of the 325 °C sample. In both spectra, the intensity of the peak shoulder is altered after PA adsorption, highlighted in the zoomed inset of a). c) The spectral components P 2s, C 1s and N 1s are fitted with Voigt functions, exemplary for a substrate annealed at 325 °C. The C 1s signal exhibits two different species, and N 1s and P 2s only one.

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 27

We used X-ray photoelectron spectroscopy to determine the stoichiometry of the molecules, layer thickness and electrostatic properties of CYNOPPA on sNiO. In Figure 2a the Ni 2p3/2 core level spectra for both annealing temperatures are given with and without CYNOPPA modification, normalized to the intensity of the main peak to enable a clear comparison of the spectral shape. The overall spectral features of sNiO are in good agreement with previous results.42,20,47,48 Unanimously, the maximum at 854.1 eV is addressed to the Ni2+, while the shoulder around 856 eV is commonly assigned to a superposition of Ni bulk hydroxide (Ni(OH)2), surface hydroxide (-OH) and oxy-hydroxide (NiOOH).42,47–51 Several shake-up satellites appear at higher binding energies. The spectra for both temperatures show small but significant changes of the peak shapes after CYNOPPA adsorption, which indicates a chemical reaction of sNiO in the surface region. To prove that the peak changes are not only caused by the exposure to ethanol, the sNiO films were immersed into pure anhydrous ethanol without PA for the same process parameters and investigated afterwards. No variation in the peak shapes could be observed, which clearly evidences that the peak changes are due to a chemisorption of CYNOPPA on sNiO. The alteration is weaker for sNiO annealed at 400 °C. This finding is in good agreement with the IR measurements revealing a smaller amount of CYNOPPA for higher annealing temperatures. A similar trend can be observed in the according O 1s core levels (see Figure 2b). It must be emphasized that unambiguous conclusions about substrate changes cannot be drawn from the O 1s spectra, since the signals of the different PA oxygen species overlap with the side-peak of the sNiO spectra. We take a closer look at the C 1s, N 1s and P 2s core level spectra depicted exemplarily for a 325 °C sample in Figure 2c. The P 2s and N 1s signals exhibit single distinct features at binding

ACS Paragon Plus Environment

8

Page 9 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

energies of 191 eV and 400 eV, indicating that none of both species underwent any chemical reaction. The carbon spectral response is split into two different species: a major component at 285 eV belonging to the aromatic carbon with hydrogen termination27 and a minor response from contributions of the C≡N group52 as well as from the carbon connected to the phosphonic anchor.53 By integrating the spectral areas of Voigt fits and dividing by the respective atomic sensitivity factors, a carbon ratio of 1:2.6 is determined, which well aligns with the carbon ratio in the molecule of 1:2.5. An overall molecular stoichiometry of 8.4 carbon to 1 nitrogen to 1.2 phosphor is found (reference: N 1s). This adopts the molecular structure of CYNOPPA with a  ratio of 7:1:1 also quite well. The nominal layer thickness  = − ln  can be determined by

measuring the intensity of the background-corrected Ni 2p spectrum before (I0) and after (Id) PAtreatment. The inelastic mean free path  (1.84 nm) of electrons contributing to the Ni 2p signal (  ≈ 617 eV) was calculated according to the equation of Tanuma, Powell and Penn.54 This yields nominal layer thicknesses for CYNOPPA of around 0.75 nm (325 °C) and 0.50 nm (400 °C), respectively. These qualitative values are close to the molecular length of a single CYNOPPA molecule (0.81 nm). Their variation presumably results from the lower coverage of molecules at 400 °C. At this point the chemical analysis of IR, DFT and XPS results allows us to assume that the chemisorbed CYNOPPA layer is indeed a SAM, with a chemical bonding of the phosphonic acid anchor, predominantly to the Ni(OH)2. Furthermore, we can deduce that a layer of superposed mono- and bidentate binding modes is formed and that the net orientation of the SAM molecules to the surface normal is dominated by the accessible binding configurations as well as the grainy microstructure of the sNiO surface.

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 27

Table 1. Work functions and valence band maxima of differently annealed and treated sNiO films. WF values were measured with XPS using the secondary electron cut-off edge and the VBM was determined via the valence band onset relative to the Fermi level. Complementary Kelvin probe data can be found in Figure S7.

T

WF

VBM

[°C]

[eV]

[eV]

325

4.81 ± 0.19

0.86

400

4.40 ± 0.05

0.97

325

5.21 ± 0.12

0.94

400

5.09 ± 0.18

1.00

325

5.55 ± 0.08

0.68

400

5.63 ± 0.13

0.75

325

5.20 ± 0.11

0.92

400

5.18 ± 0.11

0.88

sNiO:CuO

325

4.53 ± 0.12

0.34

sNiO:CuO/CYN

325

5.05 ± 0.11

0.44

sNiO

sNiO/CYN

sNiO/OP

sNiO/OP/CYN

In a next step, the electronic properties, especially work function (WF) and valence band maximum (VBM), of the sNiO surface before and after the SAM growth are investigated with XPS and compared to the treatment with an oxygen plasma (OP). All values are summarized in Table 1. For higher sNiO annealing temperatures the WF of untreated films decreases from (4.81 ± 0.19) eV down to (4.40 ± 0.05) eV. This is probably the result of a higher degree of precursor conversion and thus, NiO stoichiometry.55 As expected, the OP treatment leads to a strong increase in the work function and to a decrease in the valence band maximum of around 0.2 eV for both annealing temperatures. Such an onset decrease can be ascribed to a higher hole carrier concentration due to the oxidation of the surface.

ACS Paragon Plus Environment

10

Page 11 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

More importantly, the CYNOPPA SAM successfully increases the work function to 5.1-5.2 eV on the as-deposited sNiO. We note that, as a further result from our DFT calculations according to the method shown in previous reports,56,57 the main change in the work function upon CYNOPPA chemisorption is probably not due to the net dipole moment of CYNOPPA, but due to a charge transfer between the hydroxylated NiO surface and the phosphonate group (see Figure S15). Interestingly, this final value of around 5.2 eV is almost independent of the pretreatment (i.e. oxygen plasma) and initial WF. A possible explanation could be that EF of the sNiO becomes pinned to surface states that are not present until the chemical bonding of the molecules takes place. The WF change of 0.8 eV seen on sNiO annealed at 400 °C fits well to results reported for CYNOPPA on ITO by Koh and coworkers.31 To study the influence of work function change on device performance, F4ZnPc:C60 flatheterojunction (FHJ) solar cells with sNiO annealed at 325 °C on pre-structured ITO substrates were built and tested under AM 1.5G illumination. Obviously, this is a rather low efficiency photo-active system, but it is very reproducible and well-defined, which is crucial for our investigated interface phenomenon. Devices with 400 °C sNiO could not be fabricated, as the high temperature would reduce the ITO conductivity and thus, the overall device performance.

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 27

Figure 3. a) Linear J-V characteristics of F4ZnPc:C60 FHJ devices with untreated sNiO (square), oxygen plasma treated (dot), CYNOPPA modified (triangle), and CYNOPPA modified oxygen plasma treated (diamond) sNiO HTL. The inset depicts the used solar cell stack. Dark currents are displayed in Figure S9. b) Schematic energy band situation before contact of the different materials utilized in the solar cell device stack. The Fermi level of the functionalized sNiO hole contact layer (green) should perfectly align with the HOMO of F4ZnPc through the dipole of the CYNOPPA SAM. Figure 3a and Table 2 sum up the representative J-V characteristics and corresponding device performance data (average values can be found in Table S2). The plasma treated sNiO with a strongly increased work function shows a significantly improved FF compared to the untreated sNiO HTL.26,61 The lower series resistance (Rs) might be a consequence of the oxidative growth of

a

thin

NiOOH

surface

layer

with

increased

hole

carrier

concentration

and

conductivity.13,15,46,58 This would agree with the decrease of the VBM value mentioned before. The Voc around 640 mV stays unchanged within statistical errors for different treatments, presumably because it is already dominated by the effective bandgap (∆EDA) of the donoracceptor system.59,60 All devices with CYNOPPA modified sNiO show a second photodiode behavior (s-kink) in the operating regime. This also applies to co-evaporated bulk-heterojunction (BHJ) cells built in comparison to the FHJ devices (see Figure S10). An oxygen plasma treatment prior to the SAM deposition cannot hinder the development of the s-kink.

ACS Paragon Plus Environment

12

Page 13 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Often the occurrence of an s-kink is ascribed to hole extraction barriers caused by the energetic misalignment of the HTL’s IP or Fermi level to the donor’s HOMO. Figure 3b depicts the energy band situation before the sNiO and the F4ZnPc (HOMO = 5.3 eV) are brought into contact. Fermi levels and the IPs are adapted from the values in Table 1. The electron affinities were estimated by the according bandgaps given in literature.47 If the sNiO surface is not further treated, the Fermi level of the as-deposited HTL exhibits an offset of around 0.5 eV to the HOMO of the donor. When tuned with CYNOPPA, we succeed to minimize the offset to only 0.1 eV. The increased sNiO work function and decrease in energy offset does not lead to an increase in fill factor compared to the OP treated sNiO. Theoretically, the oxygen plasma treatment shifts the work function so far that an energetic hole extraction barrier is formed. However, this is not reflected in the J-V characteristics. If the s-kink here was solely a consequence of energetic mismatch of the modified HTL’s Fermi level with the HOMO of F4ZnPc, the oxygen plasma treated sNiO devices should exhibit an s-kink as well. This is clearly not the case. To understand this apparent contradiction, we performed a set of different experimental approaches. Table 2. Device performance parameters for FHJ cells including differently post-treated sNiO hole transport layers annealed at 325 °C. HTL

FF

Voc

Jsc

Rs

(325 °C)

[%]

[V]

[mA cm-2]

[Ω cm2]

sNiO

52

0.634

3.93

39

sNiO/OP

61

0.640

4.01

15

sNiO/CYN

34

0.646

3.60

166

sNiO/OP/CYN

36

0.645

3.85

140

sNiO:CuO

56

0.633

3.87

25

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

sNiO:CuO/CYN

49

0.641

4.06

Page 14 of 27

59

The first approach refers to a simple test introduced by Tress et al. to distinguish between extraction and injection barriers at the electrodes of hybrid solar cells with s-shaped J-V characteristics by varying the illumination intensity during the measurement.61 Following this procedure, we analyzed the s-kink under different light intensities (see Figure S11). At first glance, the characteristics are very similar to the authors model of an energetic extraction barrier. The s-kink strongly depends on the illumination intensity, because the more photo-carriers are generated, the more charges pile up at the HTL/donor interface barrier. The Voc nearly stays the same for high enough intensities. When the intensity is drastically reduced, the Voc cannot be maintained and the open-circuit voltage gradually decreases over the light intensity with a slope of around 0.07 V per decade (see Figure S13). This is a typical indicator that the device is mainly limited by non-geminate recombination in the active layer. The J-V data show that recombination at the sNiO/CYN interface is not limiting the Voc here. We were also able to reproduce the s-kink by using 1H,1H,2H,2H-perfluorooctanephosphonic acid (FHOPA), which shifts the work function of the sNiO surface to 5.1 eV (see Figure S11). The hydrophobic character of its backbone and thus, change in surface energy, has no significant influence on the

ACS Paragon Plus Environment

14

Page 15 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

s-shape. Hence, the s-shaped J-V is not correlated to a possible change of the molecular orientation of the donor.62 A further method to probe the nature of the barrier is to vary the thickness of the underlying HTL. Reducing the layer thickness behind the barrier impacts the electric field and leads to a higher probability for charges to be extracted. This explains why the s-kink is reduced for thinner sNiO layers, depicted in Figure S12, again confirming the existence of an extraction barrier. As the energy levels seem to be perfectly aligned after the functionalization, we propose that the combination of the SAM with the low intrinsic carrier density in the as-deposited sNiO turns into a bottleneck. The considerable increase of Rs for all sNiO/CYN devices (see Table 2) indicates the creation of a poorly conductive interface, which we see as the reason for the appearance of the s-kink. To support this hypothesis, we intentionally increased the charge carrier density and the conductivity of the sNiO by blending the sNiO with copper acetate to form copper oxide (CuO, see experimental details) doped NiO. Adding CuO shifts the VBM of the sNiO 0.5 eV closer to EF (see Table 1 and Figure S4). This trend reminds of the typical EF shift induced by pdoping. However, from the location of Cu relative to Ni in the periodic table, doping a NiO crystal with Cu should lead to an n-doping. Therefore, we do not understand the doping here as a replacement of Ni vacancies by Cu,63,64 but rather as a charge transfer due to the higher work function of CuO (WF = 5.4 eV).65 Nevertheless, we stick to the terminology of doping in this context. The doping is accompanied by a decreasing specific resistance from >30 Ωm (sNiO) down to 6.6 Ωm (sNiO:CuO) measured by the four-point probe method on planar films on glass.

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 27

Figure 4. a) Linear J-V characteristics of F4ZnPc:C60 FHJ devices with CuO doped sNiO (square), CYNOPPA treated CuO doped sNiO (dot), and CYNOPPA modified sNiO (triangle). The CYNOPPA reference device is always built in the same batch of investigation. The inset depicts the solar cell stack with sNiO:CuO interlayer. b) Linear I-V characteristics of F4ZnPc hole-only devices incorporating the different sNiO hole contact layers. The active area was the same as in the solar cells (4 mm²) and the layer thicknesses were 50 nm for F4ZnPc and 10 nm for the MoO3, respectively. As a consequence of the doping and decreasing specific resistance of the HTL films, the series resistance of the sNiO:CuO device decreases slightly and the FF increases from 52% to 56% compared to the as-deposited sNiO device (see Figure 4 a)), which is consistent with earlier observations by Kim and coworkers.64,66,67 More importantly, the s-kink gets repressed such that the FF is almost fully recovered. We can rule out that this is due to a lower CYNOPPA coverage on the sNiO:CuO as the characteristic IR absorption modes show similar intensities and the thickness estimated from the damping of the Ni 2p3/2 and Cu 2p3/2 core levels are almost identical to the results on undoped sNiO (see Figure S3 and S5). To further prove the transport barrier, F4ZnPc hole-only devices incorporating the functionalized undoped and doped sNiO hole contact layers were fabricated and evaluated. The corresponding I-V characteristics are shown in Figure 4 b). It can clearly be seen that the CYNOPPA SAM hinders hole extraction at negative bias, but also impedes charge injection to a certain extent at

ACS Paragon Plus Environment

16

Page 17 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

positive bias. However, the doping with CuO facilitates charge transport over the barrier, which is in very good agreement with the previous observations in the solar cells. These results lead us to the conclusion that the combination of low doped sNiO together with the SAM causes a kinetic transport barrier, similar to the observations by Cowan et al. for phosphonic acid SAMs on ZnO.68 Such a barrier can lead to a reduced field-dependent surface recombination velocity of photo-generated majority carriers at the anode.69 In other words, photo-generated holes are not able to pass the sNiO/CYNOPPA at sufficiently high rates. This also explains why the s-kink in the BHJ architecture (see Figure S10) with a higher interfacial area to the acceptor is more pronounced. We reveal that this barrier can be overcome by doping the oxide HTL and increasing its conductivity.66 This might also reduce the width of a possible space charge region, which rises due to the pile up of carriers in front of the transport barrier.70 Conclusion In summary, we demonstrate the phosphonate based SAM formation on sNiO and successfully tune the work function of sNiO. CYNOPPA increases the sNiO work function by up to 800 meV to 5.2 eV, which is among the highest reported values for SAM induced WF changes on this material. While IR and XP-spectroscopy clearly reveal the SAM formation, the IR results give further evidence that the molecules preferably attach to the Ni(OH)2. DFT calculations show that the monodentate binding is energetically more favorable than a bidentate configuration and that the tridentate binding mode is not accessible. We apply these monolayers in devices and find that the poorly conductive sNiO/SAM interface induces a kinetic transport barrier, which is not a result of energy level mismatch. This barrier can be overcome by increasing the charge carrier density in the sNiO through doping. From a more general perspective, our results emphasize that for device optimization with interface modifiers it is not sufficient to concentrate only on the

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 27

energy levels of charge extracting layers. Especially in the case of low conductivity oxides great care must be taken to avoid the creation of transport barriers. With our work, we open a pathway for future surface functionalization of hydroxylated oxides, especially sNiO, and their application in hybrid opto-electronic devices.

ACS Paragon Plus Environment

18

Page 19 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Supporting Information. The following files are available free of charge. Experimental methods, DFT calculations, device structures, IR, XPS, AFM and device data. (PDF) AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] ORCID, Sebastian Hietzschold, 0000-0003-4781-9251 *E-mail: [email protected] ORCID, Robert Lovrincic, 0000-0001-5429-5586 Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes The authors declare no competing financial interest. ACKNOWLEDGMENT We acknowledge the German Federal Ministry of Education and Research (BMBF) for financial support within the INTERPHASE project (N. 13N13656, 13N13657, 13N13658, 13N13663). We also acknowledge the support by the state of Baden-Württemberg through bwHPC and the German Research Foundation (DFG) through grant no. INST 40/467-1 FUGG. The authors also acknowledge Paul Heimel for the support during fabrication of the hole-only devices.

ACS Paragon Plus Environment

19

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 27

REFERENCES (1)

Meyer, J.; Hamwi, S.; Kröger, M.; Kowalsky, W.; Riedl, T.; Kahn, A. Transition Metal Oxides for Organic Electronics: Energetics, Device Physics and Applications. Adv. Mater. 2012, 24 (40), 5408–5427.

(2)

Greiner, M. T.; Lu, Z.-H. Thin-Film Metal Oxides in Organic Semiconductor Devices: Their Electronic Structures, Work Functions and Interfaces. NPG Asia Mater. 2013, 5 (7), e55.

(3)

Ratcliff, E. L.; Zacher, B.; Armstrong, N. R. Selective Interlayers and Contacts in Organic Photovoltaic Cells. J. Phys. Chem. Lett. 2011, 2 (11), 1337–1350.

(4)

Liu, Z.; Zhu, A.; Cai, F.; Tao, L.; Zhou, Y.; Zhao, Z.; Chen, Q.; Cheng, Y.-B.; Zhou, H. Nickel Oxide Nanoparticles for Efficient Hole Transport in P-I-N and N-I-P Perovskite Solar Cells. J. Mater. Chem. A 2017, 5 (14), 6597–6605.

(5)

Wang, K.-C.; Jeng, J.-Y.; Shen, P.-S.; Chang, Y.-C.; Diau, E. W.-G.; Tsai, C.-H.; Chao, T.-Y.; Hsu, H.-C.; Lin, P.-Y.; Chen, P.; Guo, T.-F.; Wen, T.-C. P-Type Mesoscopic Nickel Oxide/Organometallic Perovskite Heterojunction Solar Cells. Sci. Rep. 2015, 4 (1), 4756.

(6)

Jeng, J. Y.; Chen, K. C.; Chiang, T. Y.; Lin, P. Y.; Tsai, T. Da; Chang, Y. C.; Guo, T. F.; Chen, P.; Wen, T. C.; Hsu, Y. J. Nickel Oxide Electrode Interlayer in CH3NH3PbI 3 perovskite/PCBM Planar-Heterojunction Hybrid Solar Cells. Adv. Mater. 2014, 26 (24), 4107–4113.

(7)

Yin, X.; Chen, P.; Que, M.; Xing, Y.; Que, W.; Niu, C.; Shao, J. Highly Efficient Flexible Perovskite Solar Cells Using Solution-Derived NiOx Hole Contacts. ACS Nano 2016, 10 (3), 3630–3636.

(8)

Ciro, J.; Ramirez, D.; Mejia Escobar, M. A.; Montoya, J. F.; Mesa, S.; Betancur, R.; Jaramillo, F. Self-Functionalization behind a Solution-Processed NiOx Film Used As Hole Transporting Layer for Efficient Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2017, 9 (14), 12348–12354.

(9)

Pizzini, S.; Morlotti, R. Thermodynamic and Transport Properties of Stoichiometric and Nonstoichiometric Nickel Oxide. J. Electrochem. Soc. 1967, 114 (11), 1179.

(10)

Fujimori, A.; Minami, F. Valence-Band Photoemission and Optical Absorption in Nickel Compounds. Phys. Rev. B 1984, 30 (2), 957–971.

(11)

Hüfner, S.; Steiner, P.; Sander, I.; Reinert, F.; Schmitt, H.; Neumann, M.; Witzel, S. The Electronic Structure of NiO Investigated by Photoemission Spectroscopy. Solid State Commun. 1991, 80 (10), 869–873.

(12)

Greiner, M. T.; Helander, M. G.; Wang, Z. Bin; Tang, W. M.; Lu, Z. H. Effects of Processing Conditions on the Work Function and Energy-Level Alignment of NiO Thin Films. J. Phys. Chem. C 2010, 114 (46), 19777–19781.

ACS Paragon Plus Environment

20

Page 21 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(13)

Steirer, K. X.; Chesin, J. P.; Widjonarko, N. E.; Berry, J. J.; Miedaner, A.; Ginley, D. S.; Olson, D. C. Solution Deposited NiO Thin-Films as Hole Transport Layers in Organic Photovoltaics. Org. Electron. 2010, 11 (8), 1414–1418.

(14)

Steirer, K. X.; Ndione, P. F.; Widjonarko, N. E.; Lloyd, M. T.; Meyer, J.; Ratcliff, E. L.; Kahn, A.; Armstrong, N. R.; Curtis, C. J.; Ginley, D. S.; Berry, J. J.; Olson, D. C. Enhanced Efficiency in Plastic Solar Cells via Energy Matched Solution Processed NiOx Interlayers. Adv. Energy Mater. 2011, 1 (5), 813–820.

(15)

Widjonarko, N. E.; Ratcliff, E. L.; Perkins, C. L.; Sigdel, A. K.; Zakutayev, A.; Ndione, P. F.; Gillaspie, D. T.; Ginley, D. S.; Olson, D. C.; Berry, J. J. Sputtered Nickel Oxide Thin Film for Efficient Hole Transport Layer in Polymer–fullerene Bulk-Heterojunction Organic Solar Cell. Thin Solid Films 2012, 520 (10), 3813–3818.

(16)

Garcia, A.; Welch, G. C.; Ratcliff, E. L.; Ginley, D. S.; Bazan, G. C.; Olson, D. C. Improvement of Interfacial Contacts for New Small-Molecule Bulk-Heterojunction Organic Photovoltaics. Adv. Mater. 2012, 24 (39), 5368–5373.

(17)

Irwin, M. D.; Buchholz, D. B.; Hains, A. W.; Chang, R. P. H.; Marks, T. J. P-Type Semiconducting Nickel Oxide as an Efficiency-Enhancing Anode Interfacial Layer in Polymer Bulk-Heterojunction Solar Cells. Proc. Natl. Acad. Sci. 2008, 105 (8), 2783– 2787.

(18)

Zhai, Z.; Huang, X.; Xu, M.; Yuan, J.; Peng, J.; Ma, W. Greatly Reduced Processing Temperature for a Solution-Processed NiOx Buffer Layer in Polymer Solar Cells. Adv. Energy Mater. 2013, 3 (12), 1614–1622.

(19)

Ratcliff, E. L.; Garcia, A.; Paniagua, S. A.; Cowan, S. R.; Giordano, A. J.; Ginley, D. S.; Marder, S. R.; Berry, J. J.; Olson, D. C. Investigating the Influence of Interfacial Contact Properties on Open Circuit Voltages in Organic Photovoltaic Performance: Work Function versus Selectivity. Adv. Energy Mater. 2013, 3 (5), 647–656.

(20)

Liu, S.; Liu, R.; Chen, Y.; Ho, S.; Kim, J. H.; So, F. Nickel Oxide Hole Injection/Transport Layers for Efficient Solution-Processed Organic Light-Emitting Diodes. Chem. Mater. 2014, 26 (15), 4528–4534.

(21)

Singh, A.; Gupta, S. K.; Garg, A. Inkjet Printing of NiO Films and Integration as Hole Transporting Layers in Polymer Solar Cells. Sci. Rep. 2017, 7 (1), 1775.

(22)

Schulz, P.; Cowan, S. R.; Guan, Z. L.; Garcia, A.; Olson, D. C.; Kahn, A. NiOx/MoO3 BiLayers as Efficient Hole Extraction Contacts in Organic Solar Cells. Adv. Funct. Mater. 2014, 24 (5), 701–706.

(23)

Steirer, K. X.; Chesin, J. P.; Widjonarko, N. E.; Berry, J. J.; Miedaner, A.; Ginley, D. S.; Olson, D. C. Solution Deposited NiO Thin-Films as Hole Transport Layers in Organic Photovoltaics. Org. Electron. physics, Mater. Appl. 2010, 11 (8), 1414–1418.

(24)

Knesting, K. M.; Hotchkiss, P. J.; MacLeod, B. A.; Marder, S. R.; Ginger, D. S. Spatially Modulating Interfacial Properties of Transparent Conductive Oxides: Patterning Work

ACS Paragon Plus Environment

21

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 27

Function with Phosphonic Acid Self-Assembled Monolayers. Adv. Mater. 2012, 24 (5), 642–646. (25)

Sharma, A.; Hotchkiss, P. J.; Marder, S. R.; Kippelen, B. Tailoring the Work Function of Indium Tin Oxide Electrodes in Electrophosphorescent Organic Light-Emitting Diodes. J. Appl. Phys. 2009, 105 (8), 84507.

(26)

Hotchkiss, P. J.; Li, H.; Paramonov, P. B.; Paniagua, S. A.; Jones, S. C.; Armstrong, N. R.; Brédas, J. L.; Marder, S. R. Modification of the Surface Properties of Indium Tin Oxide with Benzylphosphonic Acids: A Joint Experimental and Theoretical Study. Adv. Mater. 2009, 21 (44), 4496–4501.

(27)

Kedem, N.; Blumstengel, S.; Henneberger, F.; Cohen, H.; Hodes, G.; Cahen, D. Morphology-, Synthesis- and Doping-Independent Tuning of ZnO Work Function Using Phenylphosphonates. Phys. Chem. Chem. Phys. 2014, 16 (18), 8310.

(28)

Paniagua, S. A.; Hotchkiss, P. J.; Jones, S. C.; Marder, S. R.; Mudalige, A.; Marrikar, F. S.; Pemberton, J. E.; Armstrong, N. R. Phosphonic Acid Modification of Indium−Tin Oxide Electrodes: Combined XPS/UPS/Contact Angle Studies †. J. Phys. Chem. C 2008, 112 (21), 7809–7817.

(29)

Appleyard, S. F. J.; Day, S. R.; Pickford, R. D.; Willis, M. R. Organic Electroluminescent Devices: Enhanced Carrier Injection Using SAM Derivatized ITO Electrodes. J. Mater. Chem. 2000, 10 (1), 169–173.

(30)

Stolz, S.; Scherer, M.; Mankel, E.; Lovrinčić, R.; Schinke, J.; Kowalsky, W.; Jaegermann, W.; Lemmer, U.; Mechau, N.; Hernandez-Sosa, G. Investigation of Solution-Processed Ultrathin Electron Injection Layers for Organic Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2014, 6 (9), 6616–6622.

(31)

Koh, S. E.; McDonald, K. D.; Holt, D. H.; Dulcey, C. S.; Chaney, J. a; Pehrsson, P. E. Phenylphosphonic Acid Functionalization of Indium Tin Oxide: Surface Chemistry and Work Functions. Langmuir 2006, 22 (14), 6249–6255.

(32)

Gliboff, M.; Li, H.; Knesting, K. M.; Giordano, A. J.; Nordlund, D.; Seidler, G. T.; Brédas, J. L.; Marder, S. R.; Ginger, D. S. Competing Effects of Fluorination on the Orientation of Aromatic and Aliphatic Phosphonic Acid Monolayers on Indium Tin Oxide. J. Phys. Chem. C 2013, 117 (29), 15139–15147.

(33)

Alt, M.; Jesper, M.; Schinke, J.; Hillebrandt, S.; Reiser, P.; Rödlmeier, T.; Angelova, I.; Deing, K.; Glaser, T.; Mankel, E.; Jaegermann, W.; Pucci, A.; Lemmer, U.; Bunz, U. H. F.; Kowalsky, W.; Hernandez-Sosa, G.; Lovrincic, R.; Hamburger, M. The Swiss-ArmyKnife Self-Assembled Monolayer: Improving Electron Injection, Stability, and Wettability of Metal Electrodes with a One-Minute Process. Adv. Funct. Mater. 2016, 26 (18), 3172– 3178.

(34)

Ostapenko, A.; Klöffel, T.; Meyer, B.; Witte, G. Formation and Stability of Phenylphosphonic Acid Monolayers on ZnO: Comparison of In Situ and Ex Situ SAM Preparation. Langmuir 2016, 32 (20), 5029–5037.

ACS Paragon Plus Environment

22

Page 23 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(35)

Paramonov, P. B.; Paniagua, S. A.; Hotchkiss, P. J.; Jones, S. C.; Armstrong, N. R.; Marder, S. R.; Bredas, J. L. Theoretical Characterization of the Indium Tin Oxide Surface and of Its Binding Sites for Adsorption of Phosphonic Acid Monolayers. Chem. Mater. 2008, 20 (16), 5131–5133.

(36)

Quiñones, R.; Raman, A.; Gawalt, E. S. Functionalization of Nickel Oxide Using Alkylphosphonic Acid Self-Assembled Monolayers. Thin Solid Films 2008, 516 (23), 8774–8781.

(37)

Zimmerman, J. D.; Song, B.; Griffith, O.; Forrest, S. R. Exciton-Blocking Phosphonic Acid-Treated Anode Buffer Layers for Organic Photovoltaics. Appl. Phys. Lett. 2013, 103 (24), 243905.

(38)

Ndione, P. F.; Garcia, A.; Widjonarko, N. E.; Sigdel, A. K.; Steirer, K. X.; Olson, D. C.; Parilla, P. A.; Ginley, D. S.; Armstong, N. R.; Richards, R. E.; Ratcliff, E. L.; Berry, J. J. Highly-Tunable Nickel Cobalt Oxide as a Low-Temperature P-Type Contact in Organic Photovoltaic Devices. Adv. Energy Mater. 2013, 3 (4), 524–531.

(39)

Bulusu, A.; Paniagua, S. A.; MacLeod, B. A.; Sigdel, A. K.; Berry, J. J.; Olson, D. C.; Marder, S. R.; Graham, S. Efficient Modification of Metal Oxide Surfaces with Phosphonic Acids by Spray Coating. Langmuir 2013, 29 (12), 3935–3942.

(40)

Stolz, S.; Petzoldt, M.; Dück, S.; Sendner, M.; Bunz, U. H. F.; Lemmer, U.; Hamburger, M.; Hernandez-Sosa, G. High-Performance Electron Injection Layers with a Wide Processing Window from an Amidoamine-Functionalized Polyfluorene. ACS Appl. Mater. Interfaces 2016, 8 (20), 12959–12967.

(41)

Greiner, M. T.; Helander, M. G.; Tang, W.-M.; Wang, Z.-B.; Qiu, J.; Lu, Z.-H. Universal Energy-Level Alignment of Molecules on Metal Oxides. Nat. Mater. 2012, 11 (1), 76–81.

(42)

Manders, J. R.; Tsang, S. W.; Hartel, M. J.; Lai, T. H.; Chen, S.; Amb, C. M.; Reynolds, J. R.; So, F. Solution-Processed Nickel Oxide Hole Transport Layers in High Efficiency Polymer Photovoltaic Cells. Adv. Funct. Mater. 2013, 23 (23), 2993–3001.

(43)

De Jesus, J. C.; González, I.; Quevedo, A.; Puerta, T. Thermal Decomposition of Nickel Acetate Tetrahydrate: An Integrated Study by TGA, QMS and XPS Techniques. J. Mol. Catal. A Chem. 2005, 228 (1–2), 283–291.

(44)

Hall, D. S.; Lockwood, D. J.; Bock, C.; MacDougall, B. R. Nickel Hydroxides and Related Materials: A Review of Their Structures, Synthesis and Properties. Proceedings. Math. Phys. Eng. Sci. / R. Soc. Publ. 2015, 471 (2174), 20140792.

(45)

Kamal, H.; Elmaghraby, E. K.; Ali, S. A.; Abdel-Hady, K. Characterization of Nickel Oxide Films Deposited at Different Substrate Temperatures Using Spray Pyrolysis. J. Cryst. Growth 2004, 262 (1), 424–434.

(46)

Mustafa, B.; Griffin, J.; Alsulami, A. S.; Lidzey, D. G.; Buckley, A. R. Solution Processed Nickel Oxide Anodes for Organic Photovoltaic Devices. Appl. Phys. Lett. 2014, 104 (6), 63302.

ACS Paragon Plus Environment

23

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 27

(47)

Ratcliff, E. L.; Meyer, J.; Steirer, K. X.; Garcia, A.; Berry, J. J.; Ginley, D. S.; Olson, D. C.; Kahn, A.; Armstrong, N. R. Evidence for near-Surface NiOOH Species in SolutionProcessed NiOx Selective Interlayer Materials: Impact on Energetics and the Performance of Polymer Bulk Heterojunction Photovoltaics. Chem. Mater. 2011, 23 (22), 4988–5000.

(48)

Biesinger, M. C.; Payne, B. P.; Lau, L. W. M.; Gerson, A.; Smart, R. S. C. X-Ray Photoelectron Spectroscopic Chemical State Quantification of Mixed Nickel Metal, Oxide and Hydroxide Systems. Surf. Interface Anal. 2009, 41 (4), 324–332.

(49)

Kim, K. S.; Winograd, N. X-Ray Photoelectron Spectroscopic Studies of Nickel-Oxygen Surfaces Using Oxygen and Argon Ion-Bombardment. Surf. Sci. 1974, 43 (2), 625–643.

(50)

Cappus, D.; Xu, C.; Ehrlich, D.; Dillmann, B.; Ventrice, C. A.; Al Shamery, K.; Kuhlenbeck, H.; Freund, H. J. Hydroxyl Groups on Oxide Surfaces: NiO(100), NiO(111) and Cr2O3(111). Chem. Phys. 1993, 177 (2), 533–546.

(51)

Biesinger, M. C.; Payne, B. P.; Hart, B. R.; Grosvenor, A. P.; McIntryre, N. S.; Lau, L. W.; Smart, R. S. Quantitative Chemical State XPS Analysis of First Row Transition Metals, Oxides and Hydroxides. J. Phys. Conf. Ser. 2008, 100 (1), 12025.

(52)

Barber, M.; Connor, J. A.; Guest, M. F.; Hillier, I. H.; Schwarz, M.; Stacey, M. Bonding in Some Donor–acceptor Complexes Involving Boron Trifluoride. Study by Means of ESCA and Molecular Orbital Calculations. J. Chem. Soc., Faraday Trans. 2 1973, 69 (0), 551–558.

(53)

Timpel, M.; Nardi, M. V.; Ligorio, G.; Wegner, B.; Pätzel, M.; Kobin, B.; Hecht, S.; Koch, N. Energy-Level Engineering at ZnO/Oligophenylene Interfaces with PhosphonateBased Self-Assembled Monolayers. ACS Appl. Mater. Interfaces 2015, 7 (22), 11900– 11907.

(54)

Tanuma, S.; Powell, C. J.; Penn, D. R. Calculations of Electron Inelastic Mean Free Paths. V. Data for 14 Organic Compounds over the 50-2000 eV Range. Surf. Interface Anal. 1994, 21 (3), 165–176.

(55)

Greiner, M. T.; Helander, M. G.; Wang, Z. Bin; Tang, W. M.; Lu, Z. H. Effects of Processing Conditions on the Work Function and Energy-Level Alignment of NiO Thin Films. J. Phys. Chem. C 2010, 114 (46), 19777–19781.

(56)

Khazaei, M.; Arai, M.; Sasaki, T.; Ranjbar, A.; Liang, Y.; Yunoki, S. OH-Terminated Two-Dimensional Transition Metal Carbides and Nitrides as Ultralow Work Function Materials. Phys. Rev. B 2015, 92 (7), 75411.

(57)

Leung, T. C.; Kao, C. L.; Su, W. S.; Feng, Y. J.; Chan, C. T. Relationship between Surface Dipole, Work Function and Charge Transfer: Some Exceptions to an Established Rule. Phys. Rev. B 2003, 68 (19), 195408.

(58)

Wheeler, S.; Deledalle, F.; Tokmoldin, N.; Kirchartz, T.; Nelson, J.; Durrant, J. R. Influence of Surface Recombination on Charge-Carrier Kinetics in Organic Bulk Heterojunction Solar Cells with Nickel Oxide Interlayers. Phys. Rev. Appl. 2015, 4 (2),

ACS Paragon Plus Environment

24

Page 25 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

24020. (59)

Tress, W.; Leo, K.; Riede, M. Influence of Hole-Transport Layers and Donor Materials on Open-Circuit Voltage and Shape of I-V Curves of Organic Solar Cells. Adv. Funct. Mater. 2011, 21 (11), 2140–2149.

(60)

Nanova, D.; Scherer, M.; Schell, F.; Zimmermann, J.; Glaser, T.; Kast, A. K.; Krekeler, C.; Pucci, A.; Kowalsky, W.; Schröder, R. R.; Lovrinčić, R. Why Inverted Small Molecule Solar Cells Outperform Their Noninverted Counterparts. Adv. Funct. Mater. 2015, 25 (41), 6511–6518.

(61)

Tress, W.; Inganäs, O. Simple Experimental Test to Distinguish Extraction and Injection Barriers at the Electrodes of (Organic) Solar Cells with S-Shaped Current-Voltage Characteristics. Sol. Energy Mater. Sol. Cells 2013, 117, 599–603.

(62)

Ojala, A.; Petersen, A.; Fuchs, A.; Lovrincic, R.; Pölking, C.; Trollmann, J.; Hwang, J.; Lennartz, C.; Reichelt, H.; Höffken, H. W.; Pucci, A.; Erk, P.; Kirchartz, T.; Würthner, F. Merocyanine/C60 Planar Heterojunction Solar Cells: Effect of Dye Orientation on Exciton Dissociation and Solar Cell Performance. Adv. Funct. Mater. 2012, 22 (1), 86–96.

(63)

Chen, S. C.; Kuo, T. Y.; Lin, Y. C.; Lin, H. C. Preparation and Properties of P-Type Transparent Conductive Cu-Doped NiO Films. Thin Solid Films 2011, 519 (15), 4944– 4947.

(64)

Kim, J. H.; Liang, P. W.; Williams, S. T.; Cho, N.; Chueh, C. C.; Glaz, M. S.; Ginger, D. S.; Jen, A. K. Y. High-Performance and Environmentally Stable Planar Heterojunction Perovskite Solar Cells Based on a Solution-Processed Copper-Doped Nickel Oxide HoleTransporting Layer. Adv. Mater. 2015, 27 (4), 695–701.

(65)

Yu, Z. K.; Fu, W. F.; Liu, W. Q.; Zhang, Z. Q.; Liu, Y. J.; Yan, J. L.; Ye, T.; Yang, W. T.; Li, H. Y.; Chen, H. Z. Solution-Processed CuOx as an Efficient Hole-Extraction Layer for Inverted Planar Heterojunction Perovskite Solar Cells. Chinese Chem. Lett. 2017, 28 (1), 13–18.

(66)

Kim, J.; Lee, H. R.; Kim, H. P.; Lin, T.; Kanwat, A.; Mohd Yusoff, A. R. bin; Jang, J. Effects of UV-Ozone Irradiation on Copper Doped Nickel Acetate and Its Applicability to Perovskite Solar Cells. Nanoscale 2016, 8 (17), 9284–9292.

(67)

Kim, K. H.; Takahashi, C.; Abe, Y.; Kawamura, M. Effects of Cu Doping on Nickel Oxide Thin Film Prepared by Sol–gel Solution Process. Opt. - Int. J. Light Electron Opt. 2014, 125 (12), 2899–2901.

(68)

Cowan, S. R.; Schulz, P.; Giordano, A. J.; Garcia, A.; Macleod, B. A.; Marder, S. R.; Kahn, A.; Ginley, D. S.; Ratcliff, E. L.; Olson, D. C. Chemically Controlled Reversible and Irreversible Extraction Barriers Via Stable Interface Modification of Zinc Oxide Electron Collection Layer in Polycarbazole-Based Organic Solar Cells. Adv. Funct. Mater. 2014, 24 (29), 4671–4680.

(69)

Wagenpfahl, A.; Rauh, D.; Binder, M.; Deibel, C.; Dyakonov, V. S-Shaped Current-

ACS Paragon Plus Environment

25

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 27

Voltage Characteristics of Organic Solar Devices. Phys. Rev. B - Condens. Matter Mater. Phys. 2010, 82 (11), 115306. (70)

Yaffe, O.; Scheres, L.; Segev, L.; Biller, A.; Ron, I.; Salomon, E.; Giesbers, M.; Kahn, A.; Kronik, L.; Zuilhof, H.; Vilan, A.; Cahen, D. Hg/Molecular Monolayer−Si Junctions: Electrical Interplay between Monolayer Properties and Semiconductor Doping Density. J. Phys. Chem. C 2010, 114 (22), 10270–10279.

ACS Paragon Plus Environment

26

Page 27 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

TOC

ACS Paragon Plus Environment

27