Interface-Dependent Radiative and Nonradiative Recombination in

May 2, 2018 - (52,53) We therefore report both the average PCE and stabilized PCE values extracted from maximum power point (MPP) tracking for a relia...
1 downloads 0 Views 1MB Size
Subscriber access provided by Kaohsiung Medical University

C: Energy Conversion and Storage; Energy and Charge Transport

Interface Dependent Radiative and NonRadiative Recombination in Perovskite Solar Cells Ka Kan Wong, Azhar Fakharuddin, Philipp Ehrenreich, Thomas Deckert, Mojtaba Abdi-Jalebi, Richard H. Friend, and Lukas Schmidt-Mende J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b00998 • Publication Date (Web): 02 May 2018 Downloaded from http://pubs.acs.org on May 2, 2018

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 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 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.

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 25 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

The Journal of Physical Chemistry

1

Interface Dependent Radiative and Non-radiative Recombination in

2

Perovskite Solar Cells

3

Ka Kan Wong,1 Azhar Fakharuddin*,1 Philipp Ehrenreich,1 Thomas Deckert,1 Mojtaba Abdi-Jalebi,2

4

Richard H. Friend,2 Lukas Schmidt-Mende1

5 6

1

7 8

2

Department of Physics, University of Konstanz, 78457 Konstanz, Germany

Cavendish Laboratory, Department of Physics, University of Cambridge, JJ Thomson Avenue, Cambridge CB3 0HE, UK

9 10

Correspondence: [email protected]

11

Abstract

12

Interfacial engineering has shown to play an essential role to optimizing recombination losses in

13

perovskite solar cells, however an in-depth understanding the various loss mechanisms are still

14

underway. Herein, we study the charge transfer process and reveal the primary recombination

15

mechanism at inorganic electron-transporting contact such as TiO2 and its modified organic rivals.

16

The modifiers are chemically ([6,6]-Phenyl C61 butyric acid, PC60BA) or physically ([6,6]-Phenyl

17

C61 butyric acid methyl ester, PC60BM and C60) attached fullerene to the TiO2 surface in order to

18

passivate the density of surface states. We do not observe any change in morphology, crystallinity

19

and bulk defect density of halide perovskite (CH3NH3PbI3 in this case) upon interface modification.

20

However, we observe compelling evidences via photoluminescence and electroluminescence

21

studies that the recombination dynamics at both time scales (slow and fast) are largely influenced

22

by the choice of the selective contact. We note a strong correlation between the hysteresis and the

23

so-called slow charge dynamics, both significantly influenced by the characteristics of the selective

24

contact, e.g., a the presence of surface traps at the selective contact not only shows a larger

25

hysteresis but also leads to higher charge accumulation at the interface and distinguishable slow

26

dynamics (a slower stabilization of recombination dynamics at time scale of several minutes).

27 28 29

Keywords: Interfacial charge transfer, coupled charge dynamics, radiative versus non-radiative recombination in perovskite solar cells, slow and fast charge dynamics, interfacial traps. 1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

Introduction

2

Hybrid lead halide perovskites have emerged as a new class of absorber materials for photovoltaic

3

devices. They show a strong absorption in the visible spectrum1-2 and a tunable band-gap via

4

compositional engineering3-4, a high charge carrier mobility5-6, a low exciton binding energy7-8 and

5

a low trap density9. Although the power conversion efficiency (PCE) in perovskite solar cell (PSC)

6

has reached > 22%10, research is now focused to improve device stability11-15, control the

7

microstructure morphology and crystallinity16-18, and to understand the various charge

8

recombination phenomena19-22. In general, it is shown that the high performance is always coupled

9

to an appropriate choice of electron and hole transport layers (ETL and HTL, respectively) that

10

enable not only efficient charge transfer but also reduce non-radiative recombination losses of the

11

photo generated charges23-24.

12

Since the inception of PSCs, research has been focused on improving device performance and

13

only recently, research activities are growing to understand the recombination mechanisms in

14

PSCs22,

15

recombination14, 28-31 and also degradation32-34 in the PSCs. This is often linked to the presence of

16

surface defects and catalytic sites at the TiO2 surface,35 respectively, which is overcome by

17

employing surface modifiers36-37. Recently, Olthof et al.38 showed that volatile bi-products exist on

18

metal oxide surfaces due to a chemical decomposition at the metal oxide/perovskite interface which

19

may change the energetic landscape and consequently induce an energetic barrier against effective

20

charge injection. Extensive research reports have demonstrated the use of organic or inorganic

25-27

. It is reported that metal oxide selective contacts, e.g. TiO2, induce interfacial

2

ACS Paragon Plus Environment

Page 3 of 25 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

The Journal of Physical Chemistry

1

modifiers to alter interfacial properties in order to improve the PCE and also the operational

2

stability as recently reviewed comprehensively by Fakharuddin et al23.

3

Two notable features associated with the PSCs are (i) anomalous hysteresis39-42 i.e., a change

4

in current-voltage characteristics depending on the measurement parameters and (ii) ion migration43.

5

Whereas, there is a consensus that hysteresis is largely determined by the selective contacts23, 40-41,

6

44-45

7

recently being discussed46-48. What makes PSCs more interesting is the slow (from seconds to

8

minutes) and fast (milliseconds to sub-microseconds) charge dynamics, which are conceived to be

9

due to ionic and electronic properties of halide perovskites49. However, recently, it has been shown

10

that both the ionic and electronic processes are coupled48. Whereas various reports stated that the

11

ionic environment dictates the non-radiative charge recombination in PSCs and also the energetics20,

12

35, 49-51

13

contacts and charge accumulation/transfer at the interface are relatively fresh46, 48. Until now, the

14

impact of selective contacts has been characterized individually for organic and inorganic selective

15

contacts, or to demonstrate the charge extraction capabilities of the two, often shown to influence

16

hysteresis and stability23. The conceptual difference between the two material classes, however,

17

offers a large potential to investigate interfacial charge accumulation/transfer processes in order to

18

find a link between interfacial recombination dynamics to the bulk properties such as ion migration.

19

In this study, we choose the most common inorganic selective contact, TiO2 and its modified

20

analogues with fullerene derivatives (C60, PCBA and PCBM) to investigate various charge transport

the origin of ion migration and to what extent it influences charge recombination kinetics is

, the understanding that the ionic movement is also influenced by the properties of selective

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

mechanisms. These modifiers compare the effect of chemically anchored fullerene to TiO2 via a

2

carboxyl group (PCBA), and physically attached fullerene (PCBM and C60). To mimic the influence

3

of the carboxyl group, a fullerene-free organic material, benzoic acid (which also anchors to the

4

TiO2 via a carboxyl-linker) is employed as an interface modifier. Our experiments note no change in

5

the morphology, crystallinity, and bulk defect density of the perovskite deposited on all these

6

selective contacts, however, a clear difference is apparent in the recombination kinetics, notably

7

between the slow and fast charge carrier dynamics. We show compelling evidences via

8

photoluminescence and electroluminescence studies that the recombination dynamics at both time

9

scales are largely influenced by the choice of the selective contact. The fullerene modified

10

interfaces showed significantly lower non-radiative (or higher radiative) recombination, and

11

distinguishable slow dynamics (a faster stabilization of recombination dynamics at time scale of

12

several minutes) compared to a pristine TiO2 or that modified with fullerene-free benzoic acid only.

13 14

Results and discussion

15

To evaluate the impact of TiO2/MAPbI3 interface modification on the photovoltaic performance, we

16

first examined J-V characteristics of the various PSCs made using pristine and surface modified

17

TiO2 (see supporting information for experimental details). Photovoltaic parameters of all samples

18

are shown in Fig. 1 and summarized in Table 1. It must be noted that J-V measurements may not

19

truly represent solar cells performance because the PCE is largely influenced by the measurement

20

protocol52-53. We therefore report both the average PCE and stabilized PCE values extracted from 4

ACS Paragon Plus Environment

Page 5 of 25 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

The Journal of Physical Chemistry

1

MPP tracking for a reliable comparison of our devices53. In general, the devices with

2

fullerene-based modifiers showed a significant improvement in the PV parameters and higher

3

stabilized (and average) PCE. In order to mimic the effect of carboxylic groups only and to

4

investigate whether it has an effect on the TiO2 surface passivation, we compare the J-V curves of

5

PSCs employing benzoic acid as a modifier with a pristine TiO2 analogue. The benzoic acid based

6

PSCs showed a drop in the PCE indicating its deleterious effect on the TiO2 surface. An indication

7

of the magnitude of the hysteretic effect (the so called hysteresis index, HI)54-55 by scanning in the

8

forward and backward direction shows a larger hysteresis for benzoic acid modified TiO2 than a

9

pristine TiO2 rival (Fig SI 2). In contrary to the benzoic modifier, we observe a strong improvement

10

and a smaller HI for devices with fullerene-based modifiers, which is consistent with recent

11

reports31, 56-59.

12

Surprisingly, the PSCs employing the PCBA monolayer surpass in performance those employing a

13

spin-coated PCBM thin film and also the thermally evaporated C60, suggesting a mono passivation

14

layer sufficiently inhibits recombination related to TiO2 surface defects or photocatalytic process at

15

the TiO2-perovsktie interface31-32, 60. A champion PCBA device showed a stabilized PCE greater

16

than 17% (Fig. 1d), probably due to an almost ideal surface coverage of the fullerene derivative on

17

the TiO2 (PCBA is not highly soluble in chloroform and care has to be taken to get nicely dissolved

18

solutions that will lead to complete surface coverage). The EQE measurements in ambient condition

19

at 1 sun systematically affirmed improved charge collection in the PSCs employing fullerene

20

derivatives and a lowest charge collection when benzoic acid is employed as TiO2/MAPbI3 interface 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

modifier. In the case of benzoic acid, the slightly lower EQE compared to pristine TiO2 suggests the

2

formation of a resistive passivation layer at the TiO2-perovskite interface which eventually hinders

3

charge injection. We also note the measured JSC values in JV curves are significantly lesser than

4

those calculated from EQE. This is due to the different measurement conditions of the two: The

5

EQE is measured in ambient at 1 sun using a monochromatic source with background light, whereas

6

the JV curves are measured at 82 mW/cm2 (in inert). With an increased light intensity, the JSC would

7

increase as also reported in literature.61-63

8 9 10

Table 1: TiO2/modifier and MAPbI3 film thickness, Urbach energy (calculated by fitting the Urbach tail in PDS spectra),

11

The perovskite layer thickness is 300±20 nm in all the devices. Average J-V values measured at irradiation of ~82

12

mW.cm-2 with voltage stabilization for 5 s in N2 filled glove-box. The average is calculated for 20 samples each for

13

pristine TiO2, PCBA and PCBM, 10 samples for benzoic acid and 8 samples for 8 nm C60. The value in brackets

14

corresponds to the PV parameters of the champion cell. Stabilized PCE values are obtained from MPP protocol of a

15

typical PSC for 180 s in N2 filled glove-box.

and average (champion) and stabilized photovoltaic parameters of PSCs made using pristine and surface modified TiO2.

16 17 Selective

TiO2/ modifier

Urbach

JSC

contact

thickness (nm)

energy

(mA.cm-2)

VOC (V)

FF

(meV) TiO2

50±10

16.9 ±

monolayer

13.6±2.0

8 nm

16.5 ±

11.6±1.9

--

16.4 ± 0.5

56.2±8.1

(best) PCE

PCE (%)

0.86±0.0

14.8±0.9

0.99±0.0

59.2±5.1

0.96±0.0

4.5

7.2±1.4

3.4

(8.9)

65.8±3.5

2

16.0±1.5

8.6±1.3 (11.4)

6

0.3

TiO2/PCBM

0.94±0.0 6

0.5

Acid

TiO2/C60

16.8 ±

Stabilized

(%)

0.4

TiO2/Ben.

Average

12.0±1.0

8.2

(13.2)

60.9±4.3

4

11.5±1.2

10.4

(13.4)

6

ACS Paragon Plus Environment

Page 7 of 25 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

The Journal of Physical Chemistry

TiO2/PCBA

monolayer

16.2 ± 0.3

17.0±1.0

1.00±0.0

66.6±3.8

2

13.9±1.2

12.0

(17.5)

1

2 3 4

Fig. 1: (a) J-V curves (not stabilized) of typical representative PSCs made using pristine and modified TiO2 measured at

5

external quantum efficiency of the PSCs, and (d) stabilized PCE of a champion device employing TiO2/PCBA as an

6

ETL. Inset of (d) shows the J-V curve in forward and reverse scans.

irradiation of 82 mW.cm-2, (b) stabilized PCE (or MPP tracking) of the PSCs (in a) for a reliable comparison, (c)

7 8

To verify that the observed improvement is mainly based on the interface modification and is

9

not influenced from, for instance, perovskite crystal growth after surface treatment as reported in

10

literature15,

59, 64-65

11

absorbance and crystallinity of the perovskite grown on the various selective layers. We found

, we further investigated any observable changes in the film morphology,

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

similar surface coverage and grain distribution of the perovskite (Fig. S3a). All the MAPbI3 films

2

show an absorption onset at 780 nm and there is no notable difference in the absorbance of

3

perovskite films deposited on the various interfacial layers (Fig. S3b). X-ray diffraction patterns

4

affirmed complete cubic phase transformation, and similar peak intensities further confirmed no

5

differences in the crystal growth on TiO2 layers with and without modification (Fig. 2a). No traces

6

of PbI2 are found in all the perovskite layers showing a complete transformation of initial precursors

7

to perovskite. These findings prompt our assumption that the improved performance may not be

8

attributed to a non-stoichiometry or a change of grain distribution and film quality due to the

9

surface modification, but are dominantly based on the modified TiO2/MAPbI3 interface.

8

ACS Paragon Plus Environment

Page 9 of 25 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

The Journal of Physical Chemistry

1 2 3

Fig. 2: (a) XRD spectra of the MAPbI3 films deposited on the various charge selective contacts on glass, (b) PDS

4

absorption spectra of TiO2 only and its modified analogues.

absorption spectra of MAPbI3 films on TiO2 with different modifiers deposited on quartz glass. The inset shows PDS

5 6

Another plausible explanation which has been proposed66-67 is that the fullerene interfacial

7

modifiers may diffuse into the perovskite absorber layer and suppress recombination centres at the

8

grain boundaries. However, we note a similar sub-bandgap level and crystallinity of the bulk

9

CH3NH3PbI3 layers as observed from PDS and XRD spectra, respectively (Fig. 2 a & b). A similar

10

peak to peak ratio of the various CH3NH3PbI3 phases in the XRD spectra which corresponds to the 9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

bulk crystallinity of the perovskite and the identical steepness of the PDS absorption spectra as well

2

as similar sub-bandgap levels depict no change in the bulk properties of the CH3NH3PbI3. The

3

similar Urbach energy, Eu, (Table 1, calculated from the slope of absorption onset; A ∝ exp( E /

4

Eu ), where A is the absorbance and E is the excitation energy in electron-volts) that corresponds to

5

the energetic disorder of a material further affirms a very similar defects density of the perovskite

6

films68. Contrarily, the PDS spectra of modified TiO2 interfaces compared to the pristine analogue

7

(Fig. 2b inset) shows a significant drop in the sub-bandgap states of TiO2 confirming passivation of

8

surface traps. One may argue that the PDS spectra of the modified TiO2 contains footprints of C60 as

9

reported in literature69. However, whereas the PCBM and C60 modified TiO2 may contain signature

10

from fullerene absorbance, a more than two order of magnitude drop in the tail absorbance of TiO2

11

for PCBA (which only forms a monolayer), or in the case of benzoic acid (an order of magnitude

12

drop with no fullerene) clearly suggests passivation of surface states. Our findings suggest that

13

interfaces play an important role to achieve high performance devices which is in good agreement

14

with the literature31, 56-59.

15

In order to probe the interfacial charge extraction and the radiative/non-radiative

16

recombination, we measured photoluminescence (PL) and electroluminescence (EL) spectra of the

17

films and devices, respectively. As shown in Fig 3a & b, the steady-state PL spectra of all the

18

devices showed a reduced emission intensity due to charge quenching at the TiO2/perovskite

19

interface. In general, fullerene modified TiO2 interfaces showed slightly higher PL quenching than a

20

pristine TiO2 rival. The EL spectra (Fig. 3d) showed a highest intensity for C60 modified interface 10

ACS Paragon Plus Environment

Page 11 of 25 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

The Journal of Physical Chemistry

1

followed by PCBM and PCBA, whereas the benzoic acid modified perovskite device showed least

2

EL intensity suggesting a poor charge injection in this device. A quantitative analysis of the

3

time-resolved photoluminescence (TRPL) spectra of the perovskite films on the various interfaces

4

also affirmed a more efficient charge extraction for the fullerene derivatives compared to pristine or

5

benzoic acid modified TiO2 interface (Fig. 3c). As reported in the literature70-72, the observed PL

6

decay of a perovskite film changes with illumination time. To ensure reproducible results, we have

7

measured the TRPL after stabilizing the PL intensity by ~20 minutes of light exposure. Figure SI

8

4a shows the PL intensity where the TRPL spectra of a reference perovskite films (on glass) before

9

and after light soaking are compared. We observe significantly different lifetimes for the charge

10

carriers. Interestingly, as the total emission intensity remained unchanged (as indicated from the

11

charge carrier density ‘n’ in the absolute PL spectra, Fig. SI 4a), it suggests a variation in the bulk

12

properties of perovskite under light soaking. Given the time frame of PL saturation of minutes (Fig.

13

SI 4b), it cannot be ascribed to trap filling by photoexcited electrons and holes (a process which

14

takes place in fractions of second)70 and rather needs to be assigned to halide redistribution or ion

15

migration which may change the bulk properties of the perovskite.

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1 2 3

Figure 3: (a) Steady-state PL spectra of the MAPbI3 deposited on the various selective contacts and a reference

4

time-resolved PL decay of the MAPbI3 films in measured in vacuum EL and (d) EL spectra of all devices employing the

5

various selective contacts measured in vacuum at forward bias (1.5 V).

(MAPbI3 film on glass), (b) the same as (a) without a reference on glass for clarity and to compare charge quenching, (c)

6 7

A comparison of EL emission intensity versus applied bias (Fig. 4 a) further revealed reduced

8

non-radiative recombination in PSCs employing fullerene based selective contacts. The TiO2/C60

9

modified interface yielded nearly an order of magnitude higher EL than a pristine TiO2, followed by

10

the PCBA and PCBM modified interfaces. For all the samples, a correlation between the EL intensity

11

and applied bias is observed which is attributed to higher injection currents (Jinj) in the device, i.e.

12

higher charge carrier density ‘n’, which leads to enhanced bimolecular recombination (Fig. SI 5a). 12

ACS Paragon Plus Environment

Page 13 of 25 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

The Journal of Physical Chemistry

1

The TiO2 modified with benzoic acid resulted in the lowest EL intensity and also the lowest Jinj.

2

Noteworthy is the comparison of Jinj and the EL intensity of the different samples. Firstly, the Jinj of

3

TiO2 and TiO2/Ben. acid samples cross at higher applied voltage (V > 1.4 V). Although initially

4

Jinj,TiO2 was the lowest, it becomes higher than Jinj,TiO2/Ben. Acid at V ~ 1.4V. Second notable finding is

5

that the Jinj does not always follow the EL intensity. Whereas the TiO2/C60 modified interface showed

6

the highest EL intensity and also the highest Jinj, the trend is not consistent for PCBA and PCBM

7

modified samples. Despite significantly higher EL intensity in PCBA, i.e., a higher bi-molecular

8

(radiative) recombination, a significantly lower Jinj (~8 mA at 1.5 V) is measured compared to PCBM

9

(Jinj ~12.5 mA at 1.5 V).

10

The Continuity Equation and boundary condition help correlate total current to interfacial defect

11

density at perovskite/TiO2 interface as

12





 /



= −  ∇  +  −  (1)

13

where n is the charge carrier density, q is elementary charge,    is total current density of drift

14

and diffusion processes,  denotes the charge generation rate, and  represents

15

recombination rates in the bulk and at the interface. At steady state and non open-circuit conditions,

16

∇  at TiO2 interface equals  −  ≠ 0 where  and recombination in bulk are

17

considered identical for all samples, i.e. similar absorption profile as well as bulk defects as

18

revealed by PDS spectra. Equation 1 suggests ‘n’ dependency on the current flowing through a

19

device as well as the charge generation and recombination rates. Assuming charge generation (similar 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 14 of 25

1

absorbance for all samples and similar bulk defect density as revealed by the PDS spectra), and

2



3

the cell, see Fig, SI 5a), decrease (increase) in current extracted out of the cell suggests a change in

4

interfacial trap density.



= 0, i.e., a constant amount of charge carriers in the device (for a constant forward bias applied to

5

14

ACS Paragon Plus Environment

Page 15 of 25 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

The Journal of Physical Chemistry

1 2

measured at a continuous forward applied bias (+1.5 V). At ~700 s, the applied bias is reversed (-1.5 V) and the devices

3

were measured again to track changes in EL.

4

A similar Jinj trend for the TiO2 and Benzoic acid is suggestive of similar recombination rates.

5

However, the lower EL signal for benzoic acid samples suggests that non-radiative (mostly

6

monomolecular) recombination is more dominant than bi-molecular compared to pristine TiO2

7

resulting in decreased radiative recombination. Regarding fullerene derivatives, the higher Jinj with

8

respect to TiO2 suggests an improved interfacial charge extraction owing to smaller interfacial defect

9

density and thus a lower recombination. The higher EL intensity suggests that bi-molecular (or

10

radiative) recombination dominates over non-radiative to still show higher EL signal than the pristine

11

TiO2. Given the bulk trap density of all the perovskite is similar, the increased Jinj and a higher EL

12

signal is attributed to inhibition of interfacial traps, whereas the trap density seems to be even

13

increased with benzoic acid as interfacial modifier. Regarding the interfacial traps, particularly, in the

14

case of TiO2, a possible chemical reaction could take place leading to decomposition of perovskite

15

crystals35,

16

formation of an insulating monolayer is expected which leads to charge accumulation at the interface

17

and thereby a higher non-radiative recombination and a lower Jinj. For a detailed understanding of EL

18

(transient) and the role of recombination we refer to works by Tress. et al.74-75

19

In order to investigate the temporal changes in interfacial charge accumulation and transfer, we

20

recorded EL of all the samples for a time scale much longer than that required for trap filling (even

21

deeper traps are reported to be filled in few seconds)70. As can be seen in Fig. 4b, a sudden drop in the

Figure 4: (a) Bias dependent EL of all the PSCs made using pristine and modified TiO2, and (b) EL of the same devices

73

. This suppresses the EL signal significantly. For benzoic acid modified interface,

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

EL intensity in the first 100s is evident for TiO2 and TiO2/Benzoic acid interfaces and a slower drop

2

for PCBM and PCBA modified TiO2. The C60 modified interface did not show such a significant drop

3

and seems to stabilize much earlier than the other samples. One must note that the C60 (8 nm) is

4

deposited via thermal evaporation, a method that leads to a compact fully covered layer deposition

5

(ensuring that the crack or pin-holes in the TiO2 film are patched) unlike solution procced PCBA and

6

PCBM which may contain uncovered TiO2 films areas. As the time scale of EL saturation is several

7

minutes, it cannot be attributed to trap filling. More likely it seems that the ion migration and halide

8

distribution, processes which are assumed to take place at such longer time scale dictates the EL

9

response70, 76.

10

To probe whether the monotonic decrease in the EL intensity is due to degradation, as reported in

11

literature77, we applied a negative bias (-1.5 V). We note that the drop of the EL intensity for TiO2 and

12

benzoic acid is reversible, whereas for fullerene derivative is not. One could argue that charge

13

accumulation at the former two interfaces hinder charge injection (which employs the Jinj to be also

14

consistent)78, the measured temporal current through the device (Fig. SI 6) remains stabilized and

15

unchanged before and after the applied bias. Referring back to the Continuity Equation (Equation 1),

16

this inconsistency could be attributed to a change in interfacial trap density over time and thus the

17

recombination ‘R’. Once again, as the time scale of this saturation (before and after applying negative

18

bias) is rather much longer than that expected for electronic trap filling, it suggests that the ionic

19

processes within the bulk are coupled with interfacial processes: The more defective the interface is,

20

the longer is the time required to reach a steady-state within the device. This could explain the faster 16

ACS Paragon Plus Environment

Page 17 of 25 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

The Journal of Physical Chemistry

1

stabilization of fullerene derivatives (C60 being the fastest) and a slower stabilization for TiO2 and

2

benzoic acid. Very recently, a clear link between ionic movement and interfacial recombination is

3

also reported by Pockett et al.46 using transient measurement. We also note that, although the negative

4

bias could eliminate the charges accumulated at the interfaces (TiO2 and benzoic acid), the ionic

5

processes remains less affected by the applied bias (the EL intensity stabilized much faster around the

6

negative bias region).

7

Conclusions

8

In this study, we investigate the charge transfer process and the primary recombination mechanism

9

at inorganic electron-transporting contact such as TiO2 and its modified organic rivals (with and

10

without a fullerene molecule). The modifiers are chemically ([6,6]-Phenyl C61 butyric acid,

11

PC60BA) or physically ([6,6]-Phenyl C61 butyric acid methyl ester , PC60BM and C60) attached

12

fullerene to the TiO2 surface in order to passivate the density of surface states. We do not observe

13

notable changes in morphology, composition, and sub-bandgap states in MAPbI3, suggesting no

14

change took place for the bulk properties due to interface modification. The devices with

15

fullerene-based modifiers exhibited great improvement in all photovoltaic parameters, and a lesser

16

hysteresis effect, compared to devices with pristine TiO2 and those modified with benzoic acid (a

17

carboxylic group with a benzene ring). An impressive stabilized PCE >17 % is achieved using only

18

a monolayer of PCBA despite the fact that the surface coverage by the PCBA molecules are not

19

perfect, indicating that such thin interfacial layers are essential to positively modulate surface

20

condition as well as to passivate surface traps on TiO2. We further show compelling evidences that 17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

the recombination dynamics at slow and fast time scales are largely influenced by the choice of the

2

selective contact (pristine or modified TiO2). We note a strong correlation between the hysteresis

3

and the so-called slow charge dynamics, both significantly influenced by the characteristics of the

4

selective contact, e.g., the presence of surface traps leading to a higher energetic disorder at the

5

selective contact not only shows a larger hysteresis but also leads to higher charge accumulation at

6

the interface and distinguishable slow dynamics (a slower stabilization of recombination dynamics

7

at time scale of several minutes). It appears that the metal oxide such as TiO2 in this case does not

8

form an ideal interface with MAPbI3 and in order to avoid the interfacial recombination and

9

stabilize the hysteresis due to interfaces, a fully covered thin mono-layer passivation is required.

10 11

Supporting information

12

The supporting information includes experimental methods, chemical structure of the interfacial modifiers, details of

13

PDS measurements, current-voltage curves of all the devices, SEM images and absorption spectra of perovskites,

14

additional details on photoluminescence and electroluminescence and currents of the films/devices.

15 16

Acknowledgements

17

We acknowledge funding by the BMBF in the frame of The ENARET-MED-ENERG-11-132 project “HYDROSOL”.

18

K.K.W would like to acknowledge DAAD for doctoral scholarship and thank Prof. Thomas Bein (LMU) and his group

19

members for perovskite synthesis demonstration. A.F. acknowledge Alexander von Humboldt Foundation for the

20

postdoctoral fellowship award. M.A.J. gratefully acknowledges Nava Technology Limited and Nyak Technology

21

Limited for a Ph.D. scholarship. M.A.J. and R.H.F. thank the EPSRC.

22 23

References

24

(1)

Brittman, S.; Adhyaksa, G. W. P.; Garnett, E. C. The Expanding World of Hybrid Perovskites: Materials 18

ACS Paragon Plus Environment

Page 19 of 25 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

The Journal of Physical Chemistry

1 2

Properties and Emerging Applications. MRS Commun. 2015, 5, 7-26.

3 4

2014, 8, 506-514.

5

A., Sakai, N., Korte, L., Rech, B., et al. A Mixed-Cation Lead Mixed-Halide Perovskite Absorber for Tandem

6 7

Solar Cells. Science 2016, 351, 151-155.

8 9

of Perovskite Materials for High-Performance Solar Cells. Nature 2015, 517, 476-480.

(2) (3)

(4) (5)

Green, M. A.; Ho-Baillie, A.; Snaith, H. J. The Emergence of Perovskite Solar Cells. Nat. Photonics McMeekin, D. P., Sadoughi, G., Rehman, W., Eperon, G.E., Saliba, M., Hörantner, M.T., Haghighirad,

Jeon, N. J.; Noh, J. H.; Yang, W. S.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I. Compositional Engineering Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G. Semiconducting Tin and Lead Iodide Perovskites

10

with Organic Cations: Phase Transitions, High Mobilities, and near-Infrared Photoluminescent Properties.

11 12

Inorg. Chem. 2013, 52, 9019-9038.

13

H.; Ball, J. M.; Lee, M. M.; Snaith, H. J. Electronic Properties of Meso-Superstructured and Planar

14 15

Organometal Halide Perovskite Films: Charge Trapping, Photodoping, and Carrier Mobility. ACS Nano 2014,

16

(7)

17

Lanzani, G.; Snaith, H. J.; Petrozza, A. Excitons Versus Free Charges in Organo-Lead Tri-Halide Perovskites.

18 19

Nat. Commun. 2014, 5, 3586.

20 21

the High Efficiency of Hybrid Perovskite Solar Cells. APL Mater. 2016, 4, 091505.

22

Rothenberger, A., Katsiev. et. al. K., Low Trap-State Density and Long Carrier Diffusion in Organolead

23 24

Trihalide Perovskite Single Crystals. Science 2015, 347, 519-522.

25

Noh, J. H. et al., Iodide Management in Formamidinium-Lead-Halide–Based Perovskite Layers for Efficient

26 27

Solar Cells. Science 2017, 356, 1376-1379.

28

S. M.; Tress, W.; Abate, A.; Hagfeldt, A.; et al. Cesium-Containing Triple Cation Perovskite Solar Cells:

29 30

Improved Stability, Reproducibility and High Efficiency. Energ. Environ. Sci.2016, 9, 1989-1997.

31

M.; et al. Rubidium Multication Perovskite with Optimized Bandgap for Perovskite‐Silicon Tandem with

32 33

over 26% Efficiency. Adv. Energ. Mater. 2017, 7, 1700228-n/a.

34

J.-P.; Tress, W. R.; Abate, A.; Hagfeldt, A. Incorporation of Rubidium Cations into Perovskite Solar Cells

35 36

Improves Photovoltaic Performance. Science 2016, 354, 206-209.

37

M.; Zhang, B.; Zhao, Y.; et al. Efficient and Stable Solution-Processed Planar Perovskite Solar Cells Via

38 39

Contact Passivation. Science 2017.

(6)

Leijtens, T.; Stranks, S. D.; Eperon, G. E.; Lindblad, R.; Johansson, E. M. J.; McPherson, I. J.; Rensmo,

8, 7147-7155.

(8) (9)

D’Innocenzo, V.; Grancini, G.; Alcocer, M. J. P.; Kandada, A. R. S.; Stranks, S. D.; Lee, M. M.;

Fakharuddin, A.; De Rossi, F.; Watson, T. M.; Schmidt-Mende, L.; Jose, R. Research Update: Behind Shi, D., Adinolfi, V., Comin, R., Yuan, M., Alarousu, E., Buin, A., Chen, Y., Hoogland, S.,

(10) Yang, W. S., Park, B-W., Jung, E.H., Jeon, N. J., Kim, Y. C., Lee, D. U., Shin, S. S., Seo, J., Kim, E. K.,

(11) Saliba, M.; Matsui, T.; Seo, J.-Y.; Domanski, K.; Correa-Baena, J.-P.; Nazeeruddin, M. K.; Zakeeruddin,

(12) Duong, T.; Wu, Y.; Shen, H.; Peng, J.; Fu, X.; Jacobs, D.; Wang, E. C.; Kho, T. C.; Fong, K. C.; Stocks,

(13) Saliba, M.; Matsui, T.; Domanski, K.; Seo, J.-Y.; Ummadisingu, A.; Zakeeruddin, S. M.; Correa-Baena,

(14) Tan, H.; Jain, A.; Voznyy, O.; Lan, X.; García de Arquer, F. P.; Fan, J. Z.; Quintero-Bermudez, R.; Yuan,

(15) Li, Y.; Zhao, Y.; Chen, Q.; Yang, Y.; Liu, Y.; Hong, Z.; Liu, Z.; Hsieh, Y. T.; Meng, L.; Li, Y.; et al. 19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

Multifunctional Fullerene Derivative for Interface Engineering in Perovskite Solar Cells. J. Am. Chem. Soc.

2 3

2015, 137, 15540-15547.

4

Organic-Inorganic Hybrid Perovskite Films for High Efficiency Inverted Planar Heterojunction Solar Cells.

5 6

Electrochim. Acta 2016, 191, 750-757.

7

Semiconducting Oxide Nanostructures as Efficient Charge Extraction Thin Films for Perovskite Solar Cells

8 9

with Efficiency Exceeding 16%. Adv. Energ. Mater. 2016, https://doi.org/10.1002/aenm.201502027.

(16) Yin, X.; Que, M.; Xing, Y.; Liu, X.; Que, W.; Niu, C. Solution-Induced Morphology Change of

(17) Wu, W. Q.; Huang, F.; Chen, D.; Cheng, Y. B.; Caruso, R. A., Solvent-Mediated Dimension Tuning of

(18) Xue, Q.; Chen, G.; Liu, M.; Xiao, J.; Chen, Z.; Hu, Z.; Jiang, X. F.; Zhang, B.; Huang, F.; Yang, W. et al.

10

Improving Film Formation and Photovoltage of Highly Efficient Inverted-Type Perovskite Solar Cells

11 12

through the Incorporation of New Polymeric Hole Selective Layers. Adv. Energ. Mater. 2016, 6.

13 14

Trap-Assisted Non-Radiative Recombination in Organic–Inorganic Perovskite Solar Cells. Adv. Mater. 2015,

15

(20) Azarhoosh, P.; McKechnie, S.; Frost, J. M.; Walsh, A.; van Schilfgaarde, M., Research Update:

16

Relativistic Origin of Slow Electron-Hole Recombination in Hybrid Halide Perovskite Solar Cells. APL

17 18

Mater. 2016, 4, 091501.

19

Garcia-Belmonte, G., Surface Recombination and Collection Efficiency in Perovskite Solar Cells from

20 21

Impedance Analysis. J. Phys. Chem. Lett. 2016, 7, 5105-5113.

22

Recombination in Perovskite Solar Cells: Significance of Grain Boundaries, Interface Traps, and Defect Ions.

23 24

ACS Energ. Lett. 2017, 2, 1214-1222.

25 26

Perovskite Solar Cells. Adv. Energ. Mater. 2017, 7, https://doi.org/10.1002/aenm.201700623.

27 28

Transport Materials Engineering for Stable and Efficient Perovskite Solar Cells. Nano Energy 2017.

29 30

Boundaries Dominate Non-Radiative Recombination in Ch3nh3pbi3 Perovskite Thin Films? PCCP. 2017, 19,

31

(26) Poindexter, J. R.; Hoye, R. L. Z.; Nienhaus, L.; Kurchin, R. C.; Morishige, A. E.; Looney, E. E.;

32

Osherov, A.; Correa-Baena, J.-P.; Lai, B.; Bulović, V.; et al., High Tolerance to Iron Contamination in Lead

33 34

Halide Perovskite Solar Cells. ACS Nano 2017, 11, 7101-7109.

35

Gratzel, M.; Saliba, M.; Abate, A.; et al., Identifying and Suppressing Interfacial Recombination to Achieve

36 37

High Open-Circuit Voltage in Perovskite Solar Cells. Energ. Environ. Sci. 2017, 10, 1207-1212.

38

Nazeeruddin, M. K.; Hagfeldt, A.; Graetzel, M. Enhanced Electronic Properties in Mesoporous Tio2 Via

39

Lithium Doping for High-Efficiency Perovskite Solar Cells. Nat. Commun. 2016, 7, 10379.

(19) Wetzelaer, G.-J. A. H.; Scheepers, M.; Sempere, A. M.; Momblona, C.; Ávila, J.; Bolink, H. J., 27, 1837-1841.

(21) Zarazua, I.; Han, G.; Boix, P. P.; Mhaisalkar, S.; Fabregat-Santiago, F.; Mora-Seró, I.; Bisquert, J.;

(22) Sherkar, T. S.; Momblona, C.; Gil-Escrig, L.; Ávila, J.; Sessolo, M.; Bolink, H. J.; Koster, L. J. A.,

(23) Fakharuddin, A.; Schmidt-Mende, L.; Garcia-Belmonte, G.; Jose, R.; Mora-Sero, I. Interfaces in (24) Bakr, Z. H.; Wali, Q.; Fakharuddin, A.; Schmidt-Mende, L.; Brown, T. M.; Jose, R. Advances in Hole (25) Yang, M.; Zeng, Y.; Li, Z.; Kim, D. H.; Jiang, C.-S.; van de Lagemaat, J.; Zhu, K., Do Grain 5043-5050.

(27) Correa-Baena, J.-P.; Tress, W.; Domanski, K.; Anaraki, E. H.; Turren-Cruz, S.-H.; Roose, B.; Boix, P. P.;

(28) Giordano, F.; Abate, A.; Correa Baena, J. P.; Saliba, M.; Matsui, T.; Im, S. H.; Zakeeruddin, S. M.;

20

ACS Paragon Plus Environment

Page 21 of 25 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

The Journal of Physical Chemistry

1

(29) Heo, J. H.; You, M. S.; Chang, M. H.; Yin, W.; Ahn, T. K.; Lee, S.-J.; Sung, S.-J.; Kim, D. H.; Im, S. H.

2

Hysteresis-Less Mesoscopic Ch3nh3pbi3 Perovskite Hybrid Solar Cells by Introduction of Li-Treated TiO2

3 4

Electrode. Nano Energy 2015, 15, 530-539.

5

S. I.; Hollman, D. J.; Noel, N. et al., Performance and Stability Enhancement of Dye-Sensitized and

6 7

Perovskite Solar Cells by Al Doping of TiO2. Adv. Function. Mater. 2014, 24, 6046-6055.

8

Li, C. Z.; Friend, R. H.; Jen, A. K. Y.; et al. Heterojunction modification for highly efficient

(30) Pathak, S. K.; Abate, A.; Ruckdeschel, P.; Roose, B.; Gödel, K. C.; Vaynzof, Y.; Santhala, A.; Watanabe,

(31) Wojciechowski, K.; Stranks, S. D.; Abate, A.; Sadoughi, G.; Sadhanala, A.; Kopidakis, N.; Rumbles, G.;

9 10

organic-inorganic perovskite solar cells. ACS Nano 2014, 8, 12701-12709.

11

Light Instability of Sensitized TiO2 with Meso-Superstructured Organometal Tri-Halide Perovskite Solar

12 13

Cells. Nat. Commun. 2013, 4, 2885.

14

Crystallinity of Nanorod and Planar Electron Transport Layers on the Performance and Long Term

15 16

Durability of Perovskite Solar Cells. J. Power Sources 2015, 283, 61-67.

17

Licoccia, S.; Ismail, J.; Di Carlo, A.; Brown, T. M. et al., Vertical TiO2 Nanorods as a Medium for Stable and

18 19

High-Efficiency Perovskite Solar Modules. ACS Nano 2015, 9, 8420-8429.

20 21

Hysteresis. J. Phys. Chem. Lett. 2017, 8, 2298-2303.

22 23

Nanocrystalline TiO2 for Ch3nh3pbi3 Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 16995-17000.

24 25

Surface by First-Principles Molecular Dynamics. Surface Science 1998, 402, 219-222.

26 27

Perovskites. Sci. Rep. 2017, 7, 40267.

28 29

Hybrid Perovskite Solar Cells. APL Mater. 2014, 2.

30 31

Effects of Perovskite Crystal Size and Mesoporous TiO2 Layer. J. Phys. Chem. Lett. 2014, 5, 2927-2934.

32

W.; Wojciechowski, K.; Zhang, W., Anomalous Hysteresis in Perovskite Solar Cells. J. Phys. Chem. Lett.

33 34

2014, 5, 1511-1515.

35

H. J.; Herz, L. M.; Johnston, M. B., Influence of Interface Morphology on Hysteresis in Vapor-Deposited

36 37

Perovskite Solar Cells. Adv. Electron. Mater. 2017, 3, 10.1002/aelm.201600470.

38

Grancini, G.; Binda, M.; Prato, M.; Ball, J. M.; et al., Ion Migration and the Role of Preconditioning Cycles

39

in the Stabilization of the J-V Characteristics of Inverted Hybrid Perovskite Solar Cells. Adv. Energ. Mater.

(32) Leijtens, T.; Eperon, G. E.; Pathak, S.; Abate, A.; Lee, M. M.; Snaith, H. J., Overcoming Ultraviolet

(33) Fakharuddin, A.; Di Giacomo, F.; Ahmed, I.; Wali, Q.; Brown, T. M.; Jose, R., Role of Morphology and

(34) Fakharuddin, A.; Di Giacomo, F.; Palma, A. L.; Matteocci, F.; Ahmed, I.; Razza, S.; D'Epifanio, A.;

(35) Kerner, R. A.; Rand, B. P., Linking Chemistry at the Tio2/Ch3nh3pbi3 Interface to Current–Voltage (36) Ito, S.; Tanaka, S.; Manabe, K.; Nishino, H., Effects of Surface Blocking Layer of Sb2s3 on (37) Selloni, A.; Vittadini, A.; Grätzel, M., The Adsorption of Small Molecules on the Tio2 Anatase (101) (38) Olthof, S.; Meerholz, K., Substrate-Dependent Electronic Structure and Film Formation of Mapbi3 (39) Frost, J. M.; Butler, K. T.; Walsh, A., Molecular Ferroelectric Contributions to Anomalous Hysteresis in (40) Kim, H. S.; Park, N. G., Parameters Affecting I-V Hysteresis of Ch3nh3pbi3 Perovskite Solar Cells: (41) Snaith, H. J.; Abate, A.; Ball, J. M.; Eperon, G. E.; Leijtens, T.; Noel, N. K.; Stranks, S. D.; Wang, J. T.

(42) Patel, J. B.; Wong-Leung, J.; Van Reenen, S.; Sakai, N.; Wang, J. T. W.; Parrott, E. S.; Liu, M.; Snaith,

(43) De Bastiani, M.; Dell'Erba, G.; Gandini, M.; D'Innocenzo, V.; Neutzner, S.; Kandada, A. R. S.;

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

1 2

2016, 6.

3 4

Hysteresis in CH3NH3PbI3 Perovskite Solar Cell. J. Phys. Chem. Lett. 2015, 6, 4633-4639.

5

in Mixed Perovskite Solar Cells: Polarization, Energy Barriers, and Defect Recombination. Adv. Energ.

6 7

Mater. 2016, 6, https://doi.org/10.1002/aenm.201600396.

8 9

Letters 2017, 2, 1683-1689.

Page 22 of 25

(44) Kim, H.-S.; Jang, I.-H.; Ahn, N.; Choi, M.; Guerrero, A.; Bisquert, J.; Park, N.-G., Control of I–V (45) Tress, W.; Correa Baena, J. P.; Saliba, M.; Abate, A.; Graetzel, M., Inverted Current–Voltage Hysteresis

(46) Pockett, A.; Carnie, M. J., Ionic Influences on Recombination in Perovskite Solar Cells. ACS Energy (47) Pockett, A.; Eperon, G. E.; Sakai, N.; Snaith, H. J.; Peter, L. M.; Cameron, P. J., Microseconds,

10

Milliseconds and Seconds: Deconvoluting the Dynamic Behaviour of Planar Perovskite Solar Cells. PCCP

11 12

2017, 19, 5959-5970.

13 14

Cesium Lead Bromide Perovskite. ACS Energy Letters 2017, 2, 488-496.

15 16

Perovskites. J. Phys. Condens. Matter 2017, 29, 193001.

17

Interactions as the Cause of Slow Dynamics and Hysteresis in Dye and Perovskite Solar Cells: A

18 19

Small-Perturbation Study. PCCP 2016, 18, 31033-31042.

20 21

S., Iodine Migration and Its Effect on Hysteresis in Perovskite Solar Cells. Adv. Mater. 2016, 28, 2446-2454.

22 23

Erroneous Efficiency Reports Harm Organic Solar Cell Research. Nat. Photon. 2014, 8, 669-672.

24

S., Mathiazhagan G., Hinsch A, et al. Characterization of Perovskite Solar Cells: Towards a Reliable

25 26

Measurement Protocol. APL Mater. 2016, 4, 091901.

27 28

Effects of Perovskite Crystal Size and Mesoporous TiO2 Layer. J. Phys. Chem. Lett. 2014, 5, 2927-2934.

29 30

Cells. Chem. Mater. 2016, 28, 802-812.

31

Jen, A. K. Y.; Lee, T.-L.; Snaith, H. J., C60 as an Efficient N-Type Compact Layer in Perovskite Solar Cells. J.

32 33

Phys. Chem. Lett. 2015, 6, 2399-2405.

34 35

Chemical States on the Performance of Perovskite Solar Cells. J. Mater. Chem. A2016, 4, 4392-4397.

36

Hysteresis and Stabilized Power Output over 20% in Planar Heterojunction Perovskite Solar Cells by

37

Compositional and Surface Modifications to the Low-Temperature-Processed Tio2 Layer. J. Mater. Chem.

38 39

A2017, 5, 9402-9411.

(48) Tirmzi, A. M.; Dwyer, R. P.; Hanrath, T.; Marohn, J. A., Coupled Slow and Fast Charge Dynamics in (49) Cheng, L.; Antonio, G.; Yu, Z.; Sven, H., Origins and Mechanisms of Hysteresis in Organometal Halide (50) Contreras, L.; Idigoras, J.; Todinova, A.; Salado, M.; Kazim, S.; Ahmad, S.; Anta, J. A., Specific Cation

(51) Li, C.; Tscheuschner, S.; Paulus, F.; Hopkinson, P. l. E.; Kießling, J.; Köhler, A.; Vaynzof, Y.; Huettner, (52) Zimmermann, E.; Ehrenreich, P.; Pfadler, T.; Dorman, J. A.; Weickert, J.; Schmidt-Mende, L., (53) Zimmermann E., Wong K. K., Müller M., Hu H., Ehrenreich P., Kohlstädt M., Würfel U., Mastroianni

(54) Kim, H.-S.; Park, N.-G., Parameters Affecting I–V Hysteresis of Ch3nh3pbi3 Perovskite Solar Cells: (55) Zhang, F., et al., Interfacial Oxygen Vacancies as a Potential Cause of Hysteresis in Perovskite Solar (56) Wojciechowski, K.; Leijtens, T.; Siprova, S.; Schlueter, C.; Hörantner, M. T.; Wang, J. T.-W.; Li, C.-Z.;

(57) Ma, T.; Tadaki, D.; Sakuraba, M.; Sato, S.; Hirano-Iwata, A.; Niwano, M., Effects of Interfacial (58) Cai, F.; Yang, L.; Yan, Y.; Zhang, J.; Qin, F.; Liu, D.; Cheng, Y.-B.; Zhou, Y.; Wang, T., Eliminated

(59) Li, W.; Zhang, W.; Van Reenen, S.; Sutton, R. J.; Fan, J.; Haghighirad, A. A.; Johnston, M. B.; Wang, L.; 22

ACS Paragon Plus Environment

Page 23 of 25 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

The Journal of Physical Chemistry

1

Snaith, H. J. Enhanced Performance and Light Soaking Stability of Planar Perovskite Solar Cells Using an

2 3

Amine-Based Fullerene Interfacial Modifier. J. Mater. Chem. A2016, 4, 18509-18515.

4

M.; Friend, R. H. Impact of a Mesoporous Titania-Perovskite Interface on the Performance of Hybrid

5 6

Organic-Inorganic Perovskite Solar Cells. J. Phys. Chem. Lett. 2016, 7, 3264-3269.

7

Yang, Y. A Dopant-Free Organic Hole Transport Material for Efficient Planar Heterojunction Perovskite

8 9

Solar Cells. J. Mater. Chem. A 2015, 3, 11940-11947.

(60) Abdi-Jalebi, M.; Dar, M. I.; Sadhanala, A.; Senanayak, S. P.; Giordano, F.; Zakeeruddin, S. M.; Grätzel,

(61) Liu, Y.; Chen, Q.; Duan, H. S.; Zhou, H.; Yang, Y.; Chen, H.; Luo, S.; Song, T. B.; Dou, L.; Hong, Z.;

(62) Zhao, D.; Sexton, M.; Park, H. Y.; Baure, G.; Nino, J. C.; So, F., High‐Efficiency Solution‐Processed

10 11

Planar Perovskite Solar Cells with a Polymer Hole Transport Layer. Adv. Energ. Mater. 2015, 5, 1401855.

12 13

Performance of Lead Halide Perovskite Solar Cells. J. Photopolym. Sci. Tech. 2017, 30, 577-582.

14

CH3NH3PbI3 Perovskite Solar Cells through Interfacial Engineering Using Self-Assembling Monolayer. J.

15 16

Am. Chem. Soc. 2015, 137, 2674-2679.

17

Engineering in Planar Perovskite Solar Cells: Energy Level Alignment, Perovskite Morphology Control and

18 19

High Performance Achievement. J. Mater. Chem. A 2017, 5, 1658-1666.

20 21

Route to Planar Perovskite Cells Exhibiting Reduced Hysteresis. Appl. Phys. Lett. 2015, 106, 143902.

22 23

Fullerene Passivation in CH3NH3PbI3 Planar Heterojunction Solar Cells. Nat. Commun. 2014, 5, 5784.

24 25

Optoelectronic Properties of CH3NH3PbI3 Perovskite. Adv. Energ. Mater. 2016, 6, 10.1002/aenm.201502472.

26 27

Lett. 1991, 182, 486-490.

28

Huettner, S., Emission Enhancement and Intermittency in Polycrystalline Organolead Halide Perovskite

29 30

Films. Molecules 2016, 21, 1081.

31 32

Solar Cells Using a Fullerene Derivative. Energ. Environ. Sci. 2016, 9, 2444-2452.

33 34

for the Band Gap Tunability of Metal Halide Perovskites. J. Mater. Chem. A 2017, 5, 11401-11409.

35 36

Nanocrystalline Tio2 for Ch3nh3pbi3 Perovskite Solar Cells. J. Phys. Chem. C 2014, 118, 16995-17000.

37

Understanding the Rate-Dependent J-V Hysteresis, Slow Time Component, and Aging in CH3NH3PbI3

38 39

Perovskite Solar Cells: The Role of a Compensated Electric Field. Energ. Environ. Sci.2015, 8, 995-1004.

(63) Liu, M.; Endo, M.; Shimazaki, A.; Wakamiya, A.; Tachibana, Y., Light Intensity Dependence of (64) Zuo, L.; Gu, Z.; Ye, T.; Fu, W.; Wu, G.; Li, H.; Chen, H., Enhanced Photovoltaic Performance of

(65) Yang, G.; Wang, C.; Lei, H.; Zheng, X.; Qin, P.; Xiong, L.; Zhao, X.; Yan, Y.; Fang, G., Interface

(66) Ip, A. H.; Quan, L. N.; Adachi, M. M.; McDowell, J. J.; Xu, J.; Kim, D. H.; Sargent, E. H., A Two-Step (67) Shao, Y.; Xiao, Z.; Bi, C.; Yuan, Y.; Huang, J., Origin and Elimination of Photocurrent Hysteresis by (68) Abdi-Jalebi, M., et al., Impact of Monovalent Cation Halide Additives on the Structural and (69) Skumanich, A., Optical Absorption Spectra of Carbon 60 Thin Films from 0.4 to 6.2 Ev. Chem. Phys. (70) Li, C.; Zhong, Y.; Luna, C.; Unger, T.; Deichsel, K.; Gräser, A.; Köhler, J.; Köhler, A.; Hildner, R.;

(71) Shao, S., et al., Elimination of the Light Soaking Effect and Performance Enhancement in Perovskite (72) Unger, E. L.; Kegelmann, L.; Suchan, K.; Sorell, D.; Korte, L.; Albrecht, S., Roadmap and Roadblocks (73) Ito, S.; Tanaka, S.; Manabe, K.; Nishino, H., Effects of Surface Blocking Layer of Sb2s3 on (74) Tress, W.; Marinova, N.; Moehl, T.; Zakeeruddin, S. M.; Nazeeruddin, M. K.; Gratzel, M.,

(75) Tress, W.; Yavari, M.; Domanski, K.; Yadav, P.; Niesen, B.; Correa Baena, J. P.; Hagfeldt, A.; Graetzel, 23

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 25

1

M., Interpretation and Evolution of Open-Circuit Voltage, Recombination, Ideality Factor and Subgap Defect

2

States During Reversible Light-Soaking and Irreversible Degradation of Perovskite Solar Cells. Energ.

3 4

Environ. Sci.2018, 11, 151-165.

5

Slow Dynamic Processes in Lead Halide Perovskite Solar Cells. Characteristic Times and Hysteresis. J. Phys.

6 7

Chem. Lett. 2014, 5, 2357-2363.

8

Mechanism of Perovskite CH3NH3PbI3 Diode Devices Studied by Electroluminescence and

(76) Sanchez, R. S.; Gonzalez-Pedro, V.; Lee, J. W.; Park, N. G.; Kang, Y. S.; Mora-Sero, I.; Bisquert, J.,

(77) Makoto, O.; Masaru, E.; Atsushi, W.; Masahiro, Y.; Hidefumi, A.; Yoshihiko, K., Degradation

9 10

Photoluminescence Imaging Spectroscopy. Applied Physics Express 2015, 8, 102302.

11

Processed Low Turn-on Voltage near Infrared Leds Based on Core-Shell Pbs-Cds Quantum Dots with

12

Inverted Device Structure. Nanoscale 2014, 6, 8551-8555.

(78) Sanchez, R. S.; Binetti, E.; Torre, J. A.; Garcia-Belmonte, G.; Striccoli, M.; Mora-Sero, I., All Solution

13 14 15

24

ACS Paragon Plus Environment

Page 25 of 25 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

1

The Journal of Physical Chemistry

TOC Graphic

2 3 4

25

ACS Paragon Plus Environment