You are on page 1of 44

Journal Pre-proof

Device Design Rules and Operation Principles of High-Power Perovskite Solar Cells
for Indoor Applications

Myung Hyun Ann, Jincheol Kim, Moonyong Kim, Ghaida Alosaimi, Dohyung Kim, Na
Young Ha, Jan Seidel, Nochang Park, Jae Sung Yun, Jong H. Kim
PII: S2211-2855(19)31028-6
DOI: https://doi.org/10.1016/j.nanoen.2019.104321
Reference: NANOEN 104321

To appear in: Nano Energy

Received Date: 24 July 2019


Revised Date: 1 November 2019
Accepted Date: 21 November 2019

Please cite this article as: M.H. Ann, J. Kim, M. Kim, G. Alosaimi, D. Kim, N.Y. Ha, J. Seidel, N. Park,
J.S. Yun, J.H. Kim, Device Design Rules and Operation Principles of High-Power Perovskite Solar Cells
for Indoor Applications Nano Energy, https://doi.org/10.1016/j.nanoen.2019.104321.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.


Graphical Abstract:

Device Design Rules and Operation Principles of

High-Power Perovskite Solar Cells for Indoor Applications

In this work, the device design rules for achieving high-power perovskite solar cells under
indoor light are suggested based on the device operation principle under low intensity light
conditions.
Device Design Rules and Operation Principles of

High-Power Perovskite Solar Cells for Indoor Applications

Myung Hyun Anna,1, Jincheol Kimb,1, Moonyong Kimc, Ghaida Alosaimid, Dohyung Kimd,

Na Young Hae, Jan Seideld, Nochang Parkb,***, Jae Sung Yunc,**, Jong H. Kima*

a
Department of Molecular Science and Technology, Ajou University, Suwon 16499,

Republic of Korea.
b
Electronic Convergence Material & Device Research Center, Korea Electronics Technology

Institute, Seong-Nam, Republic of Korea.


c
Australian Centre for Advanced Photovoltaics (ACAP), School of Photovoltaic and

Renewable Energy Engineering, University of New South Wales, Sydney, New South Wales

2052, Australia.
d
School of Materials Science and Engineering, University of New South Wales, Sydney, New

South Wales 2052, Australia.


e
Department of Energy Systems Research, Department of Physics, Ajou University, Suwon

16499, Republic of Korea.


*
Corresponding author.
**
Corresponding author.
***
Corresponding author.

E-mail addresses: jonghkim@ajou.ac.kr (J. H. Kim), j.yun@unsw.edu.au (J. S. Yun),

ncpark@keti.re.kr (N. Park)


1
These authors (Myung Hyun Ann and Jincheol Kim) equally contributed to this work.

Keywords: perovskite solar cells, indoor light, charge transport layer, traps, ideality factor
Abstract:

In this work, we report on the design principles of high-power perovskite solar cells (PSCs)

for low-intensity indoor light applications, with a particular focus on the electron transport

layers (ETLs). It was found that the mechanism of power generation of PSCs under low-

intensity LED and halogen lights is surprisingly different compared to the 1 Sun standard test

condition (STC). Although a higher power conversion efficiency (PCE) was obtained from

the PSC based on mesoporous-TiO2 (m-TiO2) under STC, compared to the compact-TiO2 (c-

TiO2) PSC, c-TiO2 PSCs generated higher power than m-TiO2 PSCs under low-intensity

(200-1600 Lux) conditions. This result indicates that high PCE at STC cannot guarantee a

reliable high-power output of PSCs under low-intensity conditions. Based on the systemic

characterization of the ideality factor, charge recombination, trap density, and charge-

separation, it was revealed that interfacial charge traps or defects at the electron transport

layer/perovskite have a critical impact on the resulting power density of PSC under weak

light conditions. Based on Suns-VOC measurements with local ideality factor analyses, it was

proved that the trap states cause non-ideal behavior of PSCs under low-intensity light

conditions. This is due to the additional trap states that are present at the m-TiO2/perovskite

interface, as confirmed by trap-density measurements. Based on Kelvin probe force

microscopy (KPFM) measurements, it was confirmed that these traps prohibit efficient

charge separation at the perovskite grain boundaries when the light intensity is weak.

According to these observations, it is suggested that for the fabrication of high-power PSCs

under low-intensity indoor light, the interface trap density should be lower than the excess

carrier density to fill the traps at the perovskite’s grain boundaries. Finally, using the

suggested principle, we succeeded in demonstrating high-performance PSCs by employing an

organic ETL, yielding maximum power densities up to 12.36 (56.43), 28.03 (100.97), 63.79
(187.67), and 147.74 (376.85) µW/cm2 under 200, 400, 800, and 1600 Lux LED (halogen)

illumination which are among the highest values for indoor low-intensity-light solar cells.

1. Introduction

Most electronic devices operate with the connected power sources provided by disposable or

rechargeable batteries. With the significant increase in the demand for low-power (average

power consumption of 20—50 µW) Internet of Things (IoT) sensors embedded in buildings,

the development of self-powered and wireless devices is highly desired.[1-3] This is because

batteries have a limited lifespan and huge numbers of corresponding battery replacements

will cause serious operational and environmental issues. In this regard, it has been suggested

that autonomous IoT sensors inside a building can be powered by indoor ambient light-

harvesting systems by exploiting the use of efficient photovoltaic (PV) devices.[4-6] This type

of energy-harvesting system combined with an energy storage unit such as a supercapacitor

facilitates the development of self-powered IoT devices in a “fit-and-forget manner” without

periodic battery replacement. Various types of PV devices based on silicon,[7] III-V

semiconductors,[8,9] organic dyes,[10] and polymers[11-13] have been investigated under low-

intensity-light illumination for indoor applications. Among them, a wide bandgap GaInP PV

has demonstrated the most practical level of power density, 92.7 µW/cm2 under 1000 Lux

lamps.[9]

Generally, indoor ambient light in buildings has a very weak intensity (typically

~5.0 10-2 mW/cm2 within the visible spectrum or 200—1,000 Lux). Therefore, the density

of excess (photogenerated) charge carriers significantly decreases in this range of light

intensity compared to that of outdoor sunlight. For this reason, the density of defects or trap

sites can cause substantial power loss by shunting of the PV devices under low-intensity light,

which was not considered to be a serious issue under 1 Sun standard test condition (STC, 100
mW/cm2 intensity with AM 1.5 G spectrum). This means that an optimized power conversion

efficiency (PCE) of PVs under STC does not guarantee their high performance under low-

intensity light or indoor light conditions. Therefore, to optimize the performance of high-

power indoor PVs, the mitigation of the defects or trap sites in the PVs should be considered

first rather than the enhancement of PCE for 1 Sun condition.

In recent years, organometal halide-type hybrid perovskite has attracted great interest in

the research field of next-generation solar cells because perovskite solar cells (PSCs) have

achieved exceptional PCEs over 24% based on unprecedented photocurrents, photovoltages,

and fill factors (FF) with panchromatic light absorption across the visible spectrum.[14-16]

These advantageous characteristics of perovskite suggest that it is a potentially excellent

candidate for indoor light-harvesting applications.[17,18] However, despite such excellent

performance under 1 Sun condition, there has been paid less attention to the detailed

characterization and fundamental investigation of the underlying mechanism of power

generation of PSCs at low-intensity-illumination conditions. Consequently, device design

principles and fabrication processes of PSCs have been optimized based on their performance

under STC, which is not a reliable rule for indoor energy harvesting applications.

In this article, we explore the device operation mechanism of PSCs under indoor light by

focusing on the electron transport layers (ETLs). The mechanism under low-intensity indoor

condition was revealed to be surprisingly different from that of STC. PSCs exhibited

performance mismatch and nonlinear behavior when 1 Sun and low-intensity indoor light

conditions were compared. When a comparison was made between compact (c)- and

mesoporous (m)-TiO2-based PSCs, higher efficiency was obtained from the latter under STC.

However, the situation is reversed under low-intensity indoor light, indicating that high

efficiencies under STC cannot guarantee a reliable high-power density of PSCs under low-

intensity light conditions. It was found that perovskite/ETLs interfacial charge traps or
defects have a critical impact on the resulting power density of PSCs for weak indoor light,

compared to STC. Suns-VOC measurements along with local ideality factor analyses

confirmed that these trap states cause non-ideal behavior under low-intensity indoor light

conditions. In addition, Kelvin probe force microscopy (KPFM) results indicated that these

traps significantly affect charge separation at the grain boundaries. It was shown that the

interface trap density must be lower than the excess carrier density to fill the traps at the grain

boundaries of a perovskite layer. Our studies suggest that mitigating the interfacial trap sites

by employing a compact or planar charge transport layer is a critical principle in the

optimization of indoor PSCs. Based on this in-depth understanding of the mechanism of the

operation of PSCs under low-intensity light, finally, we employed an organic electron

transport layer that further reduced interface trap states and demonstrated high power

densities up to 23.61 (56.43), 39.13 (100.97), 73.69 (187.67), and 151.27 (376.85) µW/cm2

under 200, 400, 800 and 1600 Lux LED (halogen) illumination, which are among the highest

values for indoor low-intensity light solar cells.

2. Results and discussion

2.1. Photovoltaic performances of PSCs under low-intensity light

We investigated the most common and representative PSC device architecture, the n-i-p

structure. In this architecture, TiO2 is widely used as an ETL or a hole blocking layer (HBL)

in PSCs owing to its high electron affinity, electron mobility, and good optical

transparency.[19-23] The device structure of PSCs adopted in this work is

FTO/TiO2/CH3NH3PbI3 (perovskite)/Spiro-OMeTAD/Au (Figure 1, the detailed device

fabrication method is described in the Experimental Section). In particular, the effect of the

type of TiO2, compact (c-) and compact/mesoporous (m-) type on the photovoltaic

performance was systemically studied under low-intensity (200–1600 Lux) LED and halogen
indoor lighting conditions. The emission spectra of the LED and halogen lights with the

absorption spectrum of the perovskite layer are shown in Figure S1.

Figure 2 presents the current density–voltage (J–V) characteristics of the fabricated PSCs

based on c- and m-TiO2 layers under LED and halogen lighting conditions for various

intensities ranging from 200 to 1600 Lux. Detailed photovoltaic parameters are summarized

in Table 1. Measurements of the PSCs under 1 Sun condition were also conducted for

comparison (Figure S2). Our PSC based on the m-TiO2 layer exhibited a higher PCE

(17.14 %) compared to the c-TiO2-based PSC (15.42 %) owing to the effective infiltration of

perovskite absorber into the mesoporous scaffold, which is beneficial to charge transport or

charge extraction.[24,25]

Surprisingly, the maximum power density (Pmax) of c-TiO2-based PSCs outperformed

those of the m-TiO2-based PSCs under low-intensity LED and halogen lighting conditions.

As shown in Figure 2, c-TiO2-based PSC generated Pmax values’ of 12.47 (50.85), 26.86

(90.37), 65.36 (183.15), and 149.48 (345.24) µW/cm2 under 200, 400, 800, and 1600 Lux

LED (halogen) lighting conditions, whereas those of m-TiO2-based PSCs were 8.07 (37.24),

16.23 (63.38), 38.46 (133.40), and 106.64 (220.87) µW/cm2 under the same intensities,

respectively. These results are surprising in that the photovoltaic performance of PSCs is

reversed under low-intensity indoor lighting conditions compared to the 1 Sun condition due

to the higher values of the open circuit voltage (VOC), short circuit current density (JSC), and

fill factor (FF) of c-TiO2-based PSCs compared to m-TiO2-based devices. The average values

of the photovoltaic parameters are summarized in Table 1. This implies that the higher PCE

values for PSCs under 1 Sun irradiation do not guarantee superior performance under low-

intensity light. As such, in spite of the morphological advantages of m-TiO2, they do not play

a beneficial role in charge transport or charge extraction under low-intensity light conditions.

Based on these observations we suspected that the operational mechanisms under STC and
indoor light conditions are different, possibly because of a low carrier-injection level when

the light intensity is low. We note that the thicknesses of the perovskite layers on c- and m-

TiO2 are comparable (~540 nm), as can be seen from the cross-sectional scanning electron

microscopy (SEM) images (Figure S3). This implies that the higher JSC of c-TiO2-based PSC

is not caused by a difference in light absorption. Therefore, we suspected that the enlarged

surface area provided by the mesoporous scaffold (m-TiO2) may lead to a higher probability

of carrier recombination close to the perovskite/TiO2 interface under low-intensity light

conditions compared to the planar structure (c-TiO2). This suggests that the device design and

operating principles of PSCs for indoor applications should be different from those of PSCs

under standard 1 Sun condition.

2.2. Photovoltage, ideality factor, and trap density measurements

To investigate the relationship between the underlying mechanisms of the nonlinear

photovoltaic behavior of the investigated PSCs under indoor light conditions (or at low

carrier injection levels) and charge carrier recombination, we performed Suns-VOC

measurements. The devices were illuminated with a slowly decaying (~10 ms) flashlight

(Xenon flash lamp) and photovoltages as a function of the illuminated light intensity were

recorded in the absence of series resistance. The light intensity was monitored using a single

crystal silicon reference solar cell.[26]

Figure 3a depicts plots of the measured photovoltages as a function of light intensities

from 1 to 0.0025 Sun. It was observed that m-TiO2 PSC generates higher photovoltages at

close to 1 Sun condition compared to c-TiO2 PSC. However, these values significantly

decrease with a steeper slope as the light intensity is reduced, compared to the photovoltages

of the c-TiO2 device. The photovoltage values of the c- and m-TiO2 PSCs were the same at ~

0.04 Sun, and eventually, the c-TiO2 PSC generated a higher voltage than that of the m-TiO2
PSC at intensities lower than ~ 0.04 Sun. This result correlates well with the higher Voc of

the c-TiO2 PSCs extracted from the above J—V measurements under low-intensity indoor

light conditions, as shown in Table 1.

For the detailed investigation, we measured the carrier diffusion length by performing

time-resolved photoluminescence (TR-PL) measurements for perovskite layers on c- and m-

TiO2 using time-correlated single photon counting (TCSPC) system, as shown in Figure S4

and Table S1. Using the averaged τ value and diffusion coefficient (D) of the perovskite,

diffusion lengths (L=(D*τ)0.5) were calculated for perovskites on c- and m-TiO2. Note that D

value for polycrystalline perovskite (CH3NH3PbI3) films has a large variation depending on

the properties of films such as grain size and thickness. Thus, we introduced D values for

comparable perovskite grain size and thickness to our perovskite for evaluating reasonable L

values as shown in Table S2 (grain sizes of perovskites on c- and m-TiO2, extracted from

AFM topography measurements in Figure 5, are 423±92 nm and 390±75 nm, respectively, as

shown in Figure S5). The obtained L values, ranging from ~1270 to ~2940 nm which are

comparable value to the reported ones,[27-29] are longer than the thickness of the perovskite

layer (~500 nm, Figure S3). Given that the carrier diffusion length is longer than the

thickness of the light absorber (perovskite layer in this study), the VOC is limited by non-

radiative recombination within the structure, and the dominant non-radiative recombination

process occurs via the Shockley-Read-Hall (SRH) process due to the trapping effect of deep

level or shallow level defects.[30-33] We firstly speculate that different trends of the

photovoltage drop under low-intensity light between two devices should originate from the

different degree of recombination processes at the TiO2/perovskite interfaces because the only

differences between the investigated devices (c- and m-TiO2 PSCs) are the type and structure

of the ETL (TiO2).

To explore the recombination effects, we extracted the local ideality factor (n), which is
a widely used and powerful method for examining the recombination process in PV

devices.[34,35] Deviation of the n value from 1 indicates either the existence of unusual

recombination or recombination with varying magnitude. In particular, the n > 1 diode

accounts for SRH recombination in the space-charge region of the junction such as interface

region or deep level defects, whereas the n = 1 diode accounts for an ideal device, i.e. band to

band radiative recombination, and the n < 1 is due to surface recombination or the SRH

recombination from shallow defects at the bulk region.[32,33,36]

Based on this principle, for the detailed characterization of the recombination process

for the investigated PSCs under low-intensity light, we extracted n using the following

relationships:

JSC=J0*exp(VOC/NSVT) (1)[37]

where J0 is the diode saturation current, NS is the carrier density, VT = kT/q, k is the

Boltzmann constant, T is the device temperature, and q is electron charge. Assuming that the

JSC is proportional to the light intensity, JSC=S*JL, where S is the sun concentration factor,

and JL is the photocurrent under 1 Sun condition. Then, the n values can be calculated using

the equation as follows,

n = (VTdlnJSC/dVOC)-1 = (VTdlnS/dVOC)-1 (2)[38,39]

As can be seen from Figure 3b, the n values of c- and m-TiO2 PSCs show completely

different behaviors depending on the intensity of the Sun. The n values (n > 1) of the m-TiO2

PSC has a parabolic shape with a severe light intensity dependence. From 0.007 to 0.1 Suns,

n increases from ~1 to 1.8 which means that the recombination mechanism changes from

band-to-band to deep level bulk SRH or interfacial SRH recombination as the intensity

increases.[33] This behavior is reversed when intensity is changing from 0.1 to 1 Sun.

Meanwhile, the n of the c-TiO2 PSC is <1 for entire intensity range with almost no variation.

It was reported that n < 1 is due to surface or shallow level SRH recombination.[36] Since the
surface recombination is typically more prominent for higher light intensities (> 0.5 Sun),[33]

the shallow level SRH recombination is a dominating factor for n < 1 under low light

intensity ranges (< 0.5 Sun). This indicates the perovskite/m-TiO2 interface contains more

deep level or interface defects which are more recombination active and limit photovoltage

under lower light intensity range (< 0.1 Sun). This result suggests that for indoor application

of high-power PSCs where light intensity below 0.1 Sun is used, it is important to reduce or

passivate those defects in order to maximize the photovoltage.

To further investigate the density of trap states for the c- and m-TiO2 contacts, space

charge-limited current (SCLC) measurements were performed and the results are shown in

Figure 3c, in which the electronic-only device (EOD) structure is FTO/c- or m-

TiO2/Perovskite/PCBM/Ag. The density of the trap states was estimated using equation:

VTFL=(e*nt*d2)/(2*ε*ε0)

where VTFL, e, nt, d, ε, and ε0 are the trap-filled limit voltage, electronic charge, trap density,

thickness of the active layer, dielectric constant of perovskite and permittivity of free space,

respectively.[40,41] The density values of the electron trap states were calculated to be

4.62 1015 and 1.02 1016 cm-3 for c- and m-TiO2 PSCs, respectively.

It is worth to note that the parameters of PSCs such as the photovoltage drop under low

intensity light can be different depending on grain size, roughness, defect states or its density

of perovskite. Comparing various properties between perovskites on c- and m-TiO2, we found

that the largest discrepancy was the trap density values, as summarized in Table S3, (45% less

trap density for perovskite on c-TiO2 compared to that on m-TiO2). Therefore, by combining

the Suns-VOC results with the SCLC trap density measurements, it can be inferred that the

higher density of trap states in the m-TiO2 PSC is the limiting factor that prohibits the full

generation of its potential power density under low-intensity illumination conditions.

It is important to note that m-TiO2 PSC exhibits a higher PCE compared to that of a c-
TiO2 device under standard 1 Sun condition, although they have a higher density of trap

states. The rationale of such behavior can be explained based on the charge carrier dynamics

in PSCs. It is suggested that the excess charge carriers in a lightly doped perovskite absorber

do not lead to immediate recombination that boosts the non-radiative recombination.[42-46]

Rather, a slow carrier trapping process leads to a photodoping effect[42-48] or charge

separation-favored band bending at the interface or grain boundaries.[42-46,49,50] Given that the

trap density at the perovskite/m-TiO2 interface is more than two folds higher than trap density

at the perovskite/c-TiO2 interface, we speculate that this defective interface region could

serve as a buffer layer that facilitates charge separation and transport only when the traps are

“completely” filled with carriers (excess carrier density >> interface trap density). However,

under low-intensity indoor light conditions or a low injection level, most of the relatively low

numbers of excited carriers will be trapped (excess carrier density < interface trap density).

Therefore, only a small portion of the carriers remain for quasi-Fermi level splitting, thus, the

generated VOC should be limited. This mechanism, as schematically illustrated in Figure 4,

strongly supports our findings in that the control of trap states at the perovskite/ETL interface

is a key factor in the optimization of the performance of PSCs for indoor solar cell

applications.

2.3. Kelvin probe force microscopy measurements

To investigate the morphological effect on the charge transport and separation properties at

the perovskite layer, we performed atomic force microscopy (AFM) and Kelvin probe force

microscopy (KPFM) measurements under dark condition and LED illumination with different

light intensities (400, 800, and 1600 Lux) for the perovskite layers on c- and m-TiO2 as

shown in Figure 5. It is apparent from the AFM topography images that the morphology and

grain size of the two perovskite films appear to be very similar. Corresponding contact
potential difference (CPD) images measured under dark condition are also shown in Figure

5c and d.

Previously, we have reported that the grain boundaries in the halide perovskite layer

exhibit higher charge separation and collection properties compared to the grain interior.[51,52]

This is responsible for the formation of a specific band structure established by segregation of

charged ions or PbI2.[53,54] However, this behavior was investigated under irradiation of

relatively high light intensities in excess of 0.1 Sun. Figure 6 shows CPD images measured

under low-intensity LED conditions (400, 800, and 1600 Lux) for perovskite layers on c- (a-c)

and m-TiO2 (d-f). From the result of perovskite on c-TiO2, it is evident that some regions of

the grain boundaries exhibit a higher CPD compared to the grain interiors at 400 Lux (Figure

6a, represented by white lines). This increase in the CPD is attributed to the separated holes

that accumulate on the top surface, which implies a higher charge separation as depicted in

Figure S6.[51,52] As the light intensity increases, those higher CPD grain boundaries become

more distinctive and more grain boundaries have a higher CPD compared to the grain

interiors. In sharp contrast, such CPD variation was not detected in the grain boundaries of

perovskite on m-TiO2 throughout the entire range of LED intensities (Figure 6 d-f).

Normalized line profiles are shown to clearly demonstrate a clear contrast between grain

interiors (GIs) and grain boundaries (GBs) in Figure 6g and h for the perovskite on c- and m-

TiO2, respectively (red area represent the position of the GB). It is evident that CPD at the

GBs is gradually enhanced as the LED intensity increases for the perovskite on c-TiO2.

However, the CPDs of the perovskite on m-TiO2 at the GBs are lower compared to the CPDs

at the GIs, which implies a lower charge separation efficiency at the GBs. This trend was the

same for all intensities. For further investigation, the average CPD difference between the

GBs and GIs (CPDGB-GI) was plotted at different light intensities as shown in Figure 7a for

perovskites on c- and m-TiO2 layers. It is evident that the CPDGB-GI values are positive, which
indicates a higher charge separation efficiency at the GBs compared to GI. These values are

gradually enhanced as the LED intensity is increased for the perovskite on c-TiO2. However,

the CPDGB-GI values for the perovskite on m-TiO2 are negative at all light intensities, which

implies that the charge separation at the GBs is less effective.[51,52] It was also found that

under such low light intensities, the charge separation efficiency at the GBs of perovskite on

m-TiO2 does not improve even with an increase of the number of excess carriers. It is worth

to mention that positively charged iodide vacancies are the main defects at the GBs according

to recent reports.[55,56] These defects can be filled with excess carriers, become neutral, and

stop acting as non-radiative recombination centers.[55] This is only satisfied when there are

enough excess carriers, i.e., the charge carrier density is greater than the trap density as

discussed above. Based on these studies, charge recombination at the GBs and interfaces can

be represented in Figure 7b and c for perovskite on c- and m-TiO2, respectively. Figure 7c

illustrates the situation (m-TiO2) whereby excess carriers only fill the interfacial trap states

and iodide vacancies at the GBs remain unfilled (excess carrier density < trap density) under

low-intensity light conditions. This mechanism is in good agreement with above Suns-VOC

results which exhibited that perovskite on m-TiO2 suffers from the interfacial SRH

recombination under low intensity light condition. Also, recent studies have revealed that the

SRH recombination at the perovskite interfaces limit the generation of VOC.[36,57,58] When

there is a reduction of the trap states at the interface (perovskite on c-TiO2) as shown in

Figure 7b (excess carrier density >> trap density), the charge carriers are less wasted at the

interface, thus, the GB can have a negligible or beneficial effect due to photodoping and band

bending at the GBs.

2.4. The effect of the organic electron transport layer on the PSCs under low-intensity

light
Finally, to investigate the general applicability of our proposed principle on the mitigation of

the interfacial trap density for indoor PSCs, we employed an organic electron transport

material, phenyl-C61-butyric acid methyl ester (PCBM), to the PSCs. The device

configuration is ITO/NiOx/CH3NH3PbI3/PCBM/PEIE/Ag (Figure 8a), and the details of the

fabrication method are described in the Experimental section. For this study, we chose PCBM

as a model organic ETL because it satisfies our suggested requirements of both a planar (or

compact) morphological structure and a low trap density.[59-61] The prepared PSC showed VOC,

JSC, FF, and PCE of 1.0 V, 20.09 mA/cm2, 0.76, and 15.27 %, respectively under STC (Figure

S7).

Similar to the above result, despite the lower PCE under 1 Sun condition, compared to

those of the TiO2-based PSCs, all the photovoltaic parameters (VOC, JSC, and FF) of the

PCBM PSCs under low-intensity LED and halogen illumination were higher than those of the

TiO2 PSCs (the parameters under different light-intensity conditions are summarized in Table

2). These improved VOC, JSC, and FF values facilitated excellent power generation from the

PCBM-based PSCs. As shown in Figure 8b and c, the Pmax values’ of the PCBM-based PSCs

were measured to be 23.61 (56.43), 39.13 (100.97), 73.69 (187.67), and 151.27 (376.85)

µW/cm2 under 200, 400, 800, 1600 Lux LED (halogen) illumination, respectively. To the best

of our knowledge, the Pmax values under halogen illumination are among the highest ones

reported for indoor solar cells so far. Furthermore, as shown in Figure 8d, the stabilized Pmax

of the PCBM PSC under 400 Lux of LED and halogen were monitored at the maximum

power point (MPP) as a function of time. These values were maintained at 38.21 and 103.31

µW/cm2 under LED and halogen illumination, respectively which implies stable power

generation of PCBM PSCs.

For comparison with the results obtained from the TiO2 PSCs, we also performed

measurements of Suns-VOC on the PCMB PSC using the same method. As shown in Figure
S8, the PSC based on the PCBM maintained its photovoltage most effectively among the

three compared devices under low-intensity light conditions. Moreover, it was observed that

the n values close to 1 for low intensities and were maintained under n = 1 throughout the

entire ranges of intensities. As the case of the c-TiO2 PSC, it is possible that low n is

attributed to the shallow-level defects. These results imply that the proposed principle for the

optimization of high-power PSCs under low-intensity light conditions is generally applicable

to the organic electron transport materials.

3. Conclusion

We proposed operating mechanisms for PSCs with different types of electron transport layers

under low-intensity indoor lighting conditions. For this study, we investigated the

photovoltaic performance of the PSCs by employing different types of electron transport

layers, c-TiO2, m-TiO2, and PCBM. At standard 1 Sun condition, the m-TiO2-based PSC

exhibited higher performances, whereas the c-TiO2-based device was superior under 200—

1600 Lux LED and halogen illuminations. The origin of the significant power loss of m-TiO2

PSCs under indoor light condition was revealed to be due to the high density of the interfacial

traps that results in a significant light intensity dependence of the ideality factor, as confirmed

by Suns-VOC measurements. Meanwhile, c-TiO2 PSCs exhibited an almost unchanged ideality

factor over the entire range of intensities and higher VOC values up to 0.4 Sun. Beyond 0.4

Sun, the VOC of the m-TiO2 was higher, resulting in a significantly higher VOC at 1 Sun. These

higher interfacial trap states at the m-TiO2/perovskite are completely filled with excess

carriers at only higher light intensities, and then induce positive effects such as photodoping

and favorable band bending at the interface, which is not anticipated at low illumination

conditions. KPFM characterization revealed that the interface traps play a significant role in

terms of charge separation at the grain boundaries. For the c-TiO2 PSCs, the grain boundaries
exhibited more effective charge separation compared to grain interiors, which is attributed to

the generation of sufficient excess carriers to fill the traps at the interface and iodide

vacancies at the grain boundaries. Meanwhile, there were not sufficient excess carriers to fill

the traps at the grain boundaries for the m-TiO2 PSCs under low-intensity light, owing to the

higher trap states at the interface, thus grain boundaries act as recombination centers. From

these results, we suggest that lowering the trap density at the interface is a key factor that

governs the performance of PSCs for indoor applications. Based on our proposed device

design principle and operating mechanisms for low-intensity indoor light conditions, we

demonstrated maximum power densities up to 23.61 (56.43), 39.13 (100.97), 73.69 (187.67),

and 151.27 (376.85) µW/cm2 under 200, 400, 800, and 1600 Lux LED (halogen) illumination,

by employing an organic electron transport layer (PCBM), which are among the highest

values for indoor low-intensity-light solar cells. The PSC device design principle presented in

this work suggests a new rational platform for PSCs that is compatible with low-intensity

indoor light conditions.

4. Experimental Section

4.1 Device Fabrication and Characterization

For TiO2-based PSCs fabrication, we first cleaned the patterned fluorine-doped tin oxide

(FTO) glass substrates by sonication in detergent and deionized water, acetone, and isopropyl

alcohol in sequence. After drying in a N2 stream, the substrates were further cleaned by UVO

treatment for 5 min. Blocking layers of TiO2 (c-TiO2) were then deposited onto cleaned

substrates by spray pyrolysis deposition using titanium diisopropoxide bis(acetylacetonate)

solution (Aldrich) at 450 oC. Then, mesoporous TiO2 (m-TiO2) films were spin-coated onto

the c-TiO2/FTO substrates and were calcined at 500 oC for 1 hour in air to remove the organic

portion. CH3NH3PbI3 (perovskite) and Spiro-OMeTAD layers were then spin-coated,


following the reported method.[62] Finally, 100 nm-thick gold electrodes were deposited by

thermal evaporation under a vacuum of 5.0ⅹ10-7 Torr at an evaporation rate of 2.0 Å s-1. For

PCBM-based PSC fabrication, after the indium tin oxide (ITO) glass substrates were cleaned

following the same method, nickel oxide layers were formed by spin-coating on the

substrates, following the reported method.[63] After spin-coating of the perovskite layers,

PCBM (10 mg/ml in chloroform), and BCP (5 mg/ml in ethanol) layers were spin-coated in

sequence. Finally, 100 nm-thick silver electrodes were deposited by thermal evaporation

under a vacuum of 5.0ⅹ10-7 Torr at an evaporation rate of 3.0 Å s-1. The photocurrent of the

PSCs was measured under AM 1.5 G illumination at an intensity of 100 mWcm-2 and LED

and halogen light at intensities of 200, 400, 800, and 1600 Lux.

Electron-only devices for trap density measurements were fabricated with a configuration of

FTO/c- or m-TiO2/perovskite/PCBM/Ag. PCBM (30 nm-thick) and Ag (100 nm-thick) layers

were deposited by spin-coating 10 mg/ml PCBM solution in chloroform and by thermal

evaporation under a vacuum of 5.0ⅹ10-7 Torr at an evaporation rate of 3.0 Å s-1, respectively.

All the J–V curves in this study were measured using a Keithley 4200 source meter unit.

4.2 Measurements

Cross-sectional images of the PSCs were measured using field-emission scanning microscopy

(FE-SEM, S-4800, Hitachi) with an acceleration voltage of 15 kV. The KPFM measurements

were performed using a commercial AFM system (AIST-NT SmartSPM 1000) under ambient

conditions at room temperature. Nitrogen gas was used for cleaning the surface of the

samples before measurements. The LED light source was used for illumination. The

measurements were conducted using diamond-coated conductive probes (DCP20, force

constant, k = 48 Nm-1) with a resonance frequency of 420 kHz.


Conflicts of interest

The authors declare no competing financial interests.

Acknowledgements

This work was supported by Global Infrastructure Program through the National Research

Foundation of Korea (NRF) funded by the Ministry of Science and ICT (NRF-

2018K1A3A1A17081404). This study was also supported by a grant from Priority Research

Centers Program (2019R1A6A1A11051471). This work was also financially supported by

the Ministry of Trade, Industry and Energy (MOTIE) and Korea Institute for Advancement of

Technology (KIAT) through the International Cooperative R&D program (Project No.

P0006857).

Appendix A. Supplementary data

Supplementary data to this article can be found at Supplementary Information.

References

[1] J. Gubbia, R. Buyyab, S. Marusic, M. Palaniswami, Future Gener. Comput. Sy

st. 29 (2013) 29 1645–1660.

[2] I. Lee, K. Lee, Business Horizons 58 (2015) 431–440.

[3] P. P. Ray, Comp. Inform. Sci. 30 (2018) 291–319.

[4] J. Matiko, N. Grabham, S. Beeby, M. Tudor, Meas. Sci. Technol. 25 (2014) 0

12002.

[5] M. Gorlatova, P. Kinget, I. Kymissis, D. Rubenstein, X. Wang, G. Zussman,

Wilreless Commun. IEEE 17 (2010) 18–25.


[6] N. Reich, W. Van Sark, W. Turkenburg, Renewable Energy 36 (2011) 642–647.

[7] Y. Li, N. J. Grabham, S. P. Beeby, M. J. Tudor, Sol. Energy 111 (2015) 21–2

9.

[8] A. S. Teran, E. Moon, W. Lim, G. Kim, I. Lee, D. Blaauw, J. D. Phillips, IE

EE Trans. Electron Devices 63 (2016) 2820–2825.

[9] I. Mathews, P. J. King, F. Stafford, R. Frizzell, IEEE Journal of Photovoltaics

6 (2016) 230–235.

[10] F. De Rossi, T. Pontecorvo, T. M. Brown, Appl. Energy 156 (2015) 413–422.

[11] S. -C. Shin, C. W. Koh, P. Vincent, J. S. Goo, J.-H. Bae, J. –J. Lee, C. Shin,

H. Kim, H. Y. Woo, J. W. Shim, Nano Energy 58 (2019) 466–475

[12] S. Y. Park, Y. Li, J. Kim, T. H. Lee, B. Walker, H. Y. Woo, J. Y. Kim, ACS

Appl. Mater. Interfaces 10 (2018) 3885–3894.

[13] R. Steim, T. Ameri, P. Schilinsky, C. Waldauf, G. Dennler, M. Scharber, C. J.

Brabec, Sol. Energy Mater. Sol Cells 95 (2011) 3256–3261.

[14] NREL, Best Research-Cell Efficiency Chart, http://www.nrel.gov.pv/cell-efficienc

y.html (2019).

[15] N. J. Jeon, H. Na, E. H. Jung, T.-Y. Yang, Y. G. Lee, G. Kim, H.-W. Shin,

S. I. Seok, J. Lee, J. Seo, Nat. Energy 3 (2018) 682–689.

[16] W. S. Yang, B.-W. Park, E. H. Jung, N. J. Jeon, Y. C. Kim, D. U. Lee, S. S.

Shin, J. Seo, E. K. Kim, J. H. Noh, S. I. Seok, Science 356 (2017) 356, 13

76–1379.

[17] C-Y. Chen, J. H. Chang, K.-M. Chiang, H.-L. Lin, S.-Y. Hsiao, H.-W. Lin, A

dv. Funct. Mater. 25 (2015) 7064–7070.

[18] J. Dagar, S. Castro-Hermosa, G. Lucarelli, F. Cacialli, T. M. Brown, Nano E

nergy 49 (2018) 290–299.


[19] N. J. Jeon, J. H. Noh, Y. C. Kim, W. S. Yang, S. Ryu, S. I. Seok, Nat. Mat

er. 13 (2014) 897–903.

[20] S. Haque, M. J. Mendes, O. Sanchez-Sobrado, H. Águas, E. Fortunato, R. Ma

rtins, Nano Energy 59 (2019) 91–101.

[21] Y. H. Lee, J. Luo, R. Humphry-Baker, P. Gao, M. Grätzel, M. K. Nazeeruddi

n, Adv. Funct. Mater. 25 (2015) 3925–3933.

[22] X. Liu, Z. Liu, B. Sun, X. Tan, H. Ye, Y. Tu, T. Shi, Z. Tang, G. Liao, Nan

o Eenrgy 50 (2018) 201–211.

[23] X. Li, C.-C. Chen, M. Cai, X. Hua, F. Xie, X. Liu, J. Hua, Y.-T. Long, H. T

ian, L. Han, Adv. Energy Mater. 8 (2018) 1800715.

[24] E. J. W. Crossland, N. Noel, V. Sivaram, T. Leijtens, J. A. Alexander-Webber,

H. J. Snaith, Nature 495 (2013) 215–219.

[25] B. O’Regan, M. Grätzel, Nature 353 (1991) 737–740.

[26] N. -P. Harder, A. B. Sproul, T. Brammer, A. G. Aberle, J. Appl. Phys. 9 (20

03) 2473–2479.

[27] Y. Li, W. Yan, Y. Li, S. Wang, W. Wang, Z. Bian, L. Xiao, Q. Gong, Sci. Rep. 5 (2015)

14485.

[28] P. Scajev, C. Qin, R. Aleksiejunas, P. Baronas, S. Miasojedovas, T. Fujihara, T.

Matsushima, C. Adachi, S. Jursenas, J. Phys. Chem. Lett. 9 (2018) 3167–3172.

[29] C. La-o-vorakiat, T. Salim, J. Kadro, M.-T. Khuc, R. Haselsberger, L. Cheng,

H. Xia, G. G. Gurzadyan, H. Su, Y. M. Lam, R. A. Marcus, M.-E. Michel-B

eyerle, E. E. M. Chia, Nat. Commun., 6 (2015) 7903.

[30] S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. P. Alcocer, T.

Leijtens, L. M. Herz, A. Petrozza, H. J. Snaith, Science 342 (2013) 341–344.

[31] J. Wong, J. L. Huang, O. Kunz, Z. Ouyang, S. He, P. I. Widenborg, A. G. A


berle, M. Keevers, M. A. Green, J. Appl. Phys. 105 (2009) 103705.

[32] W. Tress, Adv. Energy Mater., 7 (2017) 1602358.

[33] W. Tress, M. Yavari, K. Domansski, P. Yadav, B. Niesen, J. P. C. Baena, A.

Hagfeldt, M. Graetzel, Energy Environ. Sci., 11 (2018) 151–165.

[34] O. Kunz, J. Wong, J. Janssens, J. Bauer, O. Breitenstein, A. Aberle, Prog. Ph

otovolt: Res. Appl. 17 (2009) 35–46.

[35] Z. Hameiri, P. Chaturvedi, K. R. McIntosh, Appl. Phys. Lett. 103 (2013) 0235

01.

[36] P. Calado, D. Burkitt, J. Yao, J. Troughton, T. M. Watson, M. J. Carnie, A.

M. Telford, B. C. O’regan, J. Nelson, P. R. F. Barnes, Phys. Rev. Applied, 11

(2019) 04405.

[37] M. A. Green, “Solar cells: operating principles, technology, and system applica

tions”, Englewood Cliffs, NJ, Prentice-Hall, Inc., (1982)

[38] O. Kunz, S. Varlamov, A. Aberle, “Modelling the effects of distributed series

resistance on Suns-VOC, m-VOC and JSC-Suns curves of solar cells”, 2009 34th

IEEE Photovoltaic Specialists Conference (PVSC), IEEE, (2009) 000158.

[39] R. A. Sinton, A. Cuevas, “A quasi-steady-state open-circuit voltage method for

solar cell characterization”, Proceedings of the 16th European Photovoltaic Sol

ar Energy Conference, 1152 (2000) 1–4.

[40] Q. Dong, Y. Fang, Y. Shao, P. Mulligan, J. Qiu, L. Cao, J. Huang, Science 3

47 (2015) 967–970.

[41] B. Li, M. Li, C. Fei, G. Cao, J. Tian, J. Mater. Chem. A 5 (2017) 24168–24

177.

[42] T. Leijtens, G. E. Eperon, A. J. Barker, G. Grancini, W. Zhang, J. M. Ball,

A. R. S. Kandada, H. J. Snaith, A. Petrozza, Energy Environ. Sci. 9 (2016) 3


472–3481.

[43] W.-J. Yin, T. Shi, Y. Yan, Adv. Mater. 26 (2014) 4653–4658

[44] J. M. Ball, A. Petrozza, Nat. Energy 1 (2016) 16419.

[45] A. Walsh, D. O. Scanlon, S. Chen, X. G. Gong, S.-H. Wei, Angew. Chem. Int.

Ed. 54 (2015) 1791–1794.

[46] W.-J. Yin, T. Shi, Y. Yan, Appl. Phys. Lett. 104 (2014) 063903.

[47] T. Leijtens, S. D. Stranks, G. E. Eperon, R. Lindblad, E. M. Johansson, I. J.

McPherson, H. Rensmo, J. M. Ball, M. M. Lee, H. J. Snaith, ACS Nano 8 (2

014) 7147–7155.

[48] E. M. Hutter, G. E. Eperon, S. D. Stranks, T. J. Savenije, J. Phys. Chem. Let

t. 6 (2015) 3082–3090.

[49] J. S. Manser, P. V. Kamat, Nat. Photon. 8 (2014) 737–743.

[50] Y.-F. Chen, Y.-T. Tsai, D. Bassani, R. Clerc, D. Forgács, H. Bolink, M. Wussl

er, W. Jaegermann, G. Wantz, L. Hirsch, J. Mater. Chem A 4 (2016) 17529–1

7536.

[51] J. S. Yun, A. Ho-Baille, S. Huang, S. H. Woo, Y. Heo, J. Seidel, F. Huang,

Y.-B. Cheng, M. A. Green, J. Phys. Chem. Lett. 6 (2015) 875–880.

[52] J. S. Yun, J. Kim, T. Young, R. J. Patterson, D. Kim, J. Seidel, S. Lim, M.

A. Green, S. Huang, A. Ho-Baillie, Adv. Funct. Mater. 28 (2018) 1705363.

[53] Q. Chen, H. Zhou, T.-B. Song, S. Luo, Z. Hong, H.-S. Duan, L. Dou, Y. Liu,

Y. Yang, Nano Lett. 14 (2014) 4158–4163.

[54] J. S. Yun, J. Seidel, J. Kim, A. M. Soufiani, S. Huang, J. Lau, N. J. Jeon, S.

I. Seok, M. A. Green, A. Ho-Baillie, Adv. Energy Mater. 6 (2016) 1600330.

[55] T. S. Sherkar, C. Momblona, L. n. Gil-Escrig, J. Ávila, M. Sessolo, H. J. Bol

ink, L. J. A. Koster, ACS Energy Lett. 2 (2017) 1214–1222.


[56] P. Lopez-Varo, J. A. Jiménez-Tejada, M. Gracía-Rosell, S. Ravishankar, G. Gar

cia-Belmonte, J. Bisquert, O. Almora, Adv. Energy. Mater. 8 (2018) 1702772.

[57] J. Xiang, Y. Li, F. Huang, D. Zhong, Phys. Chem. Chem. Phys., 21 (2019) 1

7836–17845.

[58] C. M. Wolff, P. Caprioglio, M. Stolterfoht, D. Neher, Adv. Mater., Early View

(DOI: 10.1002/adma.201902762).

[59] J. H. Heo, H. J. Han, D. Kim, T. K. Ahn, S. H. Im, Energy Environ. Sci. 8

(2015) 1602–1608.

[60] Y. Bai, H. Yu, Z. Zhu, K. Jiang, T. Zhang, N. Zhao, S. Yang, H. Yan, J. Ma

ter. Chem A 3 (2015) 9098–9102.

[61] V. Trifiletti, V. Roiati, S. Colella, R. Giannuzzi, L. De marco, A. Rizzo, M.

Manca, A. Listorti, G. Gigli, ACS Appl. Mater. Interfaces 7 (2015) 4283–4289.

[62] Y. Cho, A. M. Soufiani, J. S. Yun, J. Kim, D. S. Lee, J. Seidel, X. Deng, M.

A. Green, S. Huang, A. W. Y. Ho-Baillie, Adv. Energy Mater. 8 (2018) 1703

392.

[63] J. H. Kim, P.-W. Liang, S. T. Williams, N. Cho, C.-C. Chueh, M. S. Glaz, D.

S. Ginger, A. K.-Y. Jen, Adv. Mater. 27 (2015) 695–701.


Figure 1. Device structures of perovskite solar cells based on (a) compact- and (b) meso-

TiO2 layers employed in this study.


Figure 2. J—V curves for perovskite solar cells based on (a) compact- and (b) mesoporous-

TiO2 under LED illumination and (c) compact- and (d) mesoporous-TiO2 under halogen

illumination.
Figure 3. (a) Photovoltage and (b) local ideality factor as a function of light intensity for

PSCs based on c- and m-TiO2 ETLs, and (c) J—V curves for electron-only devices based on

c- and m-TiO2.
Figure 4. A comparison of the operational mechanism between (a) 1 Sun condition whereby

the excess carrier density is much higher than that of the interfacial trap density (excess

carrier density >> interface trap density) and (b) indoor condition whereby the excess carrier

density is less than that of the interfacial trap density (excess carrier density < interface trap

density). Under 1 Sun condition, the excited electrons fill all the trap states at the

perovskite/ETL interface, thereby inducing photodoping and establishing charge transport

favorable band bending. For indoor light conditions, most of the photogenerated and

separated electrons are trapped at the perovskite/ETL interface, resulting in dominant SRH

recombination induced by carrier recombination at the trap sites.


Figure 5. Topography images from atomic force microscopy measurements for perovskite

layers on (a) c- and (b) m-TiO2 layers, and (c) and (d) are corresponding CPD images

measured by KPFM under dark condition (2 µmⅹ2 µm).


Figure 6. Contact potential images (4  mⅹ4  m) measured under 400, 800, and 1600 Lux

LED light for perovskites on (a-c) c- and (d-f) m-TiO2. (g) and (h) are normalized line

profiles of grains interior (GI) and grain boundaries (GB) (black dotted line) in (c) and (f),

respectively, at corresponding light intensities.


Figure 7. (a) CPD differences between GBs and GIs at different light intensities for

perovskite on c- and m-TiO2 layers. (b) and (c) are schematic diagrams depicting carrier

trapping at the interfaces and grain boundaries for perovskite on c- and m-TiO2, respectively.

The iodide vacancies are all filled with the excess carriers for the perovskite on c-TiO2

(excess carrier density >> trap density). However, for perovskite on m-TiO2, filling of the

iodide vacancies by excess carriers does not occur due to the interface traps (excess carrier

density < trap density).


Figure 8. (a) Device structure of perovskite solar cell based on PCBM layer and J—V curves

under (b) LED and (c) halogen illumination. (d) Stabilized power density at maximum power

point (0.64 V for LED light and 0.67 V for halogen light) of the PSC measured under LED

and halogen illumination for 60 seconds.


Table 1. Summary of averaged photovoltaic parameters for the PSCs based on c- and m-TiO2

layers under indoor lighting conditions.

Light Source/ Type of VOC JSC FF Power density


Intensity (Lux) TiO2 [V] [µA/cm2] (maximum value)
[µW/cm2]
LED/ c 0.72 33.49 0.47 11.26 (12.47)
200 m 0.62 27.88 0.37 6.44 (8.07)
LED/ c 0.73 63.29 0.55 25.17 (26.86)
400 m 0.66 57.13 0.38 14.33 (16.23)
LED/ c 0.77 126.34 0.62 60.17 (65.36)
800 m 0.70 114.53 0.43 34.03 (38.46)
LED/ c 0.78 261.17 0.64 131.19 (149.48)
1600 m 0.77 240.78 0.51 93.44 (106.64)
Halogen/ c 0.72 104.61 0.60 45.24 (50.85)
200 m 0.70 100.10 0.50 34.52 (37.24)
Halogen/ c 0.76 180.01 0.63 85.74 (96.88)
400 m 0.73 165.40 0.50 59.74 (63.38)
Halogen/ c 0.78 320.70 0.65 162.87 (183.15)
800 m 0.76 305.95 0.51 118.52 (133.40)
Halogen/ c 0.80 583.53 0.65 303.30 (345.24)
1600 m 0.80 497.99 0.50 199.80 (220.87)

Table 2. Summary of average photovoltaic parameters for the PSCs based on PCBM under

indoor lighting conditions.

Power density
Light Source/ VOC JSC
FF (maximum value)
Intensity (Lux) [V] [µA/cm2]
[µW/cm2]
LED/
0.78 37.76 0.76 22.27 (23.61)
200
LED/
0.80 63.64 0.75 38.37 (39.13)
400
LED/
0.83 115.45 0.75 71.80 (73.69)
800
LED/
0.85 223.71 0.75 143.06 (151.27)
1600
Halogen/
0.86 83.44 0.73 52.45 (56.43)
200
Halogen/
0.89 146.41 0.76 98.91 (100.97)
400
Halogen/
0.92 258.89 0.77 182.32 (187.67)
800
Halogen/
0.94 507.61 0.77 365.94 (376.85)
1600
Myung Hyun Ann is a graduate student in the Department of Molecular Science and Technology of

Ajou University. He is under the supervision of Professor Jong H. Kim, and his research field is

fabrication and characterization of perovskite solar cells for indoor applications.

Jincheol Kim is currently a senior researcher in Korea Electronics Technology Institute at Republic of

Korea. He received his B.S. and M.S. in materials science and engineering from Seoul National

University, Korea in 2008 and 2011, respectively. He received his PhD degree on perovskite solar

cells in University of New South Wales in 2018. His research interests focus on the development of

high performance perovskite solar cell and its applications.

Moonyong Kim is currently a Ph.D. candidate and research assistant in the school of photovoltaics

and renewable energy engineering from the University of New South Wales in Australia. He received

his honour’s degree in solar energy and photovoltaics from the University of New South Wales. His

major research interest is understanding the mechanism of light-induced degradation in crystalline

silicon solar cell.


Jae Sung Yun is currently Australian Centre for Advanced Photovoltaics fellow and lecturer at School

of Photovoltaic and Renewable Energy Engineering at University of New South Wales, Australia. He

received his PhD study on fabrication and characterization of polycrystalline Si thin-film solar cells in

SPREE under the supervision of Professor Martin A. Green in 2015. His current research focuses on

functional nanoscale imaging of defects in photovoltaic materials such as perovskite, kesterite, and Si.

Jong H. Kim is an associate professor in the department of materials science and engineering of Ajou

University at Republic of Korea. He received Ph.D. degree in the Department of Materials Science

and Engineering, Republic of Korea, in 2011. After his Ph.D. study, he worked as a postdoctoral

research associate in the department of materials science and engineering of University of Washington.

His research interests are focused on the high-power perovskite solar cells for indoor applications.
Supplementary Information

Device Design Rules and Operation Principles of

High-Power Perovskite Solar Cells for Indoor Applications

Myung Hyun Anna,1, Jincheol Kimb,1, Moonyong Kimc, Ghaida Alosaimid, Dohyung Kimd,

Na Young Hae, Jan Seideld, Nochang Parkb,***, Jae Sung Yunc,**, Jong H. Kima*
a
Department of Molecular Science and Technology, Ajou University, Suwon 16499, Republic of Korea.

b
Electronic Convergence Material & Device Research Center, Korea Electronics Technology Institute, Seong-

Nam, Republic of Korea.


c
Australian Centre for Advanced Photovoltaics (ACAP), School of Photovoltaic and Renewable Energy

Engineering, University of New South Wales, Sydney, New South Wales 2052, Australia.
d
School of Materials Science and Engineering, University of New South Wales, Sydney, New South Wales

2052, Australia.
e
Department of Energy Systems Research, Department of Physics, Ajou University, Suwon 16499, Republic of

Korea.
*
Corresponding author.
**
Corresponding author.
***
Corresponding author.

E-mail addresses: jonghkim@ajou.ac.kr (J. H. Kim), j.yun@unsw.edu.au (J. S. Yun), ncpark@keti.re.kr (N.

Park)
1
These authors (Myung Hyun Ann and Jincheol Kim) equally contributed to this work.

Keywords: perovskite solar cells, indoor light, charge transport layer, traps, ideality factor
Figure S1. Emission spectra of the LED and halogen lights with the absorption spectrum of a

perovskite layer.

Figure S2. J―V curves and photovoltaic parameters for perovskite solar cells based on (a) c-

and (b) m-TiO2 layers under 1 Sun standard condition.


Figure S3. Cross-sectional images of perovskite solar cells based on (a) c- and (b) m-TiO2

layers.

Figure S4. PL decay curves of perovskite layers on c-TiO2 (red line) and m-TiO2 (blue line).
The black lines are the exponential fits to the corresponding curves.
Figure S5. Distribution of grain sizes of perovskite layers on c- and m-TiO2.

Figure S6. KPFM measurement setup.


Figure S7. J―V curves and photovoltaic parameters for perovskite solar cells based on

PCBM layer under standard 1 Sun condition.

Figure S8. (a) Photovoltage and (b) local ideality factor as a function of light intensity for

perovskite solar cells based on c-, m-TiO2, and PCBM electron transport layers.
Table S1. Summary of parameters from biexponential fitting results of TR-PL measurements.
Sample τ1 (ns) fraction (%) τ2 (ns) fraction (%) Averaged τ (ns)
c-TiO2/Perovskite 31.9 38.4 126.4 61.6 90.1
m-TiO2/Perovskite 31.2 20.7 107.2 79.3 91.4

Table S2. Diffusion lengths for perovskite layers on c- and m-TiO2.


Diffusion Length (nm)
Diffusion coefficient (cm2/s)
/Grain size (nm)/Thickness
Perovskite on c-TiO2 Perovskite on m-TiO2
(nm)/
Diffusion length (µm)
0.18/~250/~390/1.7[R1] 1273.5 1282.7
0.95/~303/~400/1.9–3.8[R2] 2925.7 2946.7

Table S3. Summary of properties for perovskite layers on c- and m-TiO2.


Perovskite on c-TiO2 Perovskite on m-TiO2
Thickness (nm) ~500 ~500
Grain size (nm) 423±92 390±75
Roughness (nm) 34.8 40.1
Trap density from Figure 3c (cm-3) 4.62x1015 1.02x1016
Highlights:

The device design principles of high-power perovskite solar cells for indoor light

applications are investigated.

For high-power perovskite solar cells under indoor light, the trap density at the

electron transport layer/perovskite should be lower than excess carrier density.

Based on the suggested principles and mechanism, high-power perovskite solar cells

up to 376.84 µW/cm2 under indoor light are demonstrated.


Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like