You are on page 1of 61

Review

Cite This: Chem. Rev. 2018, 118, 3447−3507 pubs.acs.org/CR

Nonfullerene Acceptor Molecules for Bulk Heterojunction Organic


Solar Cells
Guangye Zhang,†,‡,∥ Jingbo Zhao,†,∥ Philip C. Y. Chow,†,‡,∥ Kui Jiang,†,‡,∥ Jianquan Zhang,†,‡
Zonglong Zhu,† Jie Zhang,§ Fei Huang,§ and He Yan*,†,‡,§

Department of Chemistry and Hong Kong Branch of Chinese National Engineering Research Center for Tissue Restoration &
Reconstruction, Hong Kong University of Science and Technology (HKUST), Clear Water Bay, Kowloon, Hong Kong, China

HKUST-Shenzhen Research Institute, No. 9 Yuexing first RD, Hi-tech Park, Nanshan, Shenzhen 518057, China
§
Institute of Polymer Optoelectronic Materials and Devices, State Key Laboratory of Luminescent Materials and Devices, South
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

China University of Technology, Guangzhou 510640, P. R. China


Downloaded via NATL INST OF TECH KOZHIKODE on August 22, 2022 at 09:10:56 (UTC).

ABSTRACT: The bulk-heterojunction blend of an electron donor and an electron


acceptor material is the key component in a solution-processed organic photovoltaic
device. In the past decades, a p-type conjugated polymer and an n-type fullerene
derivative have been the most commonly used electron donor and electron acceptor,
respectively. While most advances of the device performance come from the design of
new polymer donors, fullerene derivatives have almost been exclusively used as electron
acceptors in organic photovoltaics. Recently, nonfullerene acceptor materials,
particularly small molecules and oligomers, have emerged as a promising alternative
to replace fullerene derivatives. Compared to fullerenes, these new acceptors are
generally synthesized from diversified, low-cost routes based on building block materials
with extraordinary chemical, thermal, and photostability. The facile functionalization of
these molecules affords excellent tunability to their optoelectronic and electrochemical
properties. Within the past five years, there have been over 100 nonfullerene acceptor
molecules synthesized, and the power conversion efficiency of nonfullerene organic solar
cells has increased dramatically, from ∼2% in 2012 to >13% in 2017. This review summarizes this progress, aiming to describe
the molecular design strategy, to provide insight into the structure−property relationship, and to highlight the challenges the field
is facing, with emphasis placed on most recent nonfullerene acceptors that demonstrated top-of-the-line photovoltaic
performances. We also provide perspectives from a device point of view, wherein topics including ternary blend device,
multijunction device, device stability, active layer morphology, and device physics are discussed.

CONTENTS 5. A-D-A Type Acceptors 3464


5.1. Early Reports of A-D-A Type Acceptors 3465
1. Introduction 3448 5.2. A-D-A Type Acceptors with a Core Unit of
2. Brief Introduction to OPV Working Principle 3449 IDT or Its Derivatives 3468
2.1. Light Absorption and Exciton Generation 3449 5.2.1. Design of the Core Units 3468
2.2. Charge-Transfer State and Free Ccarrier 5.2.2. Side Chain Effects 3469
Generation 3450 5.2.3. Effect of the Spacer Unit 3470
2.3. Charge Transport 3452 5.2.4. Design of the “A” Unit 3470
3. Molecular Design Principle 3452 5.3. A-D-A Type Acceptors with Other Core Units 3471
4. Rylene Diimide-Based NFAS 3453 5.4. Summary 3472
4.1. PDI-Based NFAs 3453 6. Other NFAS 3472
4.1.1. PDI Monomers: Requirement for Twist- 6.1. DPP-Based NFAs 3472
ing 3454 6.2. Others 3475
4.1.2. Twisting Strategy-PDI Dimers 3454 6.2.1. Fused-Ring Aromatics 3475
4.1.3. Twisting Strategy-PDI Trimers and Tet- 6.2.2. Linear Oligomers 3475
ramers 3457 6.2.3. Twisted Dimers 3475
4.1.4. Discussion: How Much Twisting Is 6.2.4. Three Dimensional Structures 3476
Needed? 3459 7. Comparison among Different Core Units 3476
4.1.5. Fused-Ring PDI 3459 8. Nonfullerene All Small Molecule OPV 3477
4.1.6. α-Substituted PDI 3462
4.2. Other Rylene-Based NFAs 3463
4.3. Summary and Challenges 3463 Received: September 5, 2017
Published: March 20, 2018

© 2018 American Chemical Society 3447 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

9. Ternary Blend Nonfullerene OPV 3478 fabricated using high-throughput processes at low temperatures,
9.1. Motivations to Use Ternary Blend 3478 making them considered as a promising candidate for the next
9.2. Benefits of NFAs in Ternary Blend OPVs 3478 generation PV technologies.
9.3. Utilizing NFAs in Ternary Blend OPVs 3480 Most OPV devices are fabricated using the blend of a
9.3.1. Pair of Miscible NFAs in Ternary Blend 3480 conjugated polymer and a fullerene derivative as the active
9.3.2. Blending NFA with a Pair of Polymer components, 7−9 with the maximum power conversion
Donors 3481 efficiency (PCE) over 11%.10,11 The majority of these
9.3.3. Combining NFA and Fullerene in a achievements are enabled by the design of new polymer
Ternary Blend 3481 donors. In contrast, the archetypal fullerene derivative,
10. Enhanced Device Stability of Nonfullerene OPV 3482 PC61BM,12 and its C70 analog, PC71BM,13 have been
10.1. Enhanced Stability of BHJ Solar Cells with dominantly used as the electron accepting materials ever
IDT Based A-D-A Type NFAs 3482 since the application of the former in 1995.7 Electronically,
10.2. Enhanced Stability of BHJ Solar Cells with fullerenes have a proficient electron withdrawing ability14 and a
PDI or DPP Based NFAs 3483 high electron mobility.15 The physical shape of fullerenes has
10.3. Enhanced Stability of Ternary Blend Non- also offered them three-dimensional electron transport property
fullerene Solar Cells 3484 and the capability of forming favorable blend morphology that
10.4. Semitransparent Nonfullerene Solar Cell balances charge generation and transport.16
with Enhanced Stability 3484 Despite these beneficial properties, fullerenes possess
10.5. Summary 3485 considerable limitations. First, fullerenes do not absorb strongly
11. Reducing Voltage Loss 3485 in the visible region of the solar spectrum. Meanwhile, chemical
12. Challenges and Perspectives 3486 modification of their backbone is not straightforward, rendering
12.1. Perspectives on Molecular Design Strategy a low structural flexibility and an elevated difficulty in tuning
and Synthetic Complexity 3486 the electronic/optical properties. This not only increases the
12.1.1. Molecular Design to Enhance Light synthetic complexity but also makes fullerenes less likely to gain
Absorption 3486 complementary light harvesting to the polymers. When made
12.1.2. Molecular Design to Balance Morphol- into devices, fullerenes have been shown to have poor
ogy and Charge Transport 3487 photostability in air in both pristine and blend films.17
12.1.3. Synthetic Complexity 3487 The demand for replacing fullerenes in OPVs has led to the
12.2. Studying Charge Transport 3488 rapid development of nonfullerene acceptors (NFAs), including
12.3. Understanding Morphology of Nonfuller- both polymeric and small molecular organic acceptors. In the
ene Active Layers 3488 past few years, dramatic progress has been made in the field of
12.3.1. Quantitative Morphology-Performance both branches. If not specified, NFAs throughout this review
Correlations 3488 refer to the small molecular organic nonfullerene electron
12.3.2. Molecular Miscibility 3489 accepting materials. Unlike fullerenes,18 NFAs have excellent
12.3.3. Solvent Additive Effect 3489 synthetic flexibility with more readily available source materials,
12.3.4. Vertical Phase Segregation 3490 which affords easily tunable optical/electronic properties and
12.4. Multijunction NFA OPVs 3491 improved solubility. In conjunction with their high absorption
12.5. Device Structure Engineering and New coefficients, NFAs can be easily tailored to work with novel
Interfacial Layers 3492 polymers in terms of both optical complementarity and
12.6. Thick Active Layer, Large-Area Devices, energetic compatibility, allowing the solar cell to attain broad
Environmentally Friendly Processing, and solar spectrum coverage. In addition, the low-cost cores, facile
Roll-to-Roll Printing 3493 synthesis and simplified purification can significantly reduce the
13. Concluding Remarks 3494 production cost of NFAs.
Author Information 3494 As a result of these advantageous properties, huge efforts
Corresponding Author 3494 have been devoted to designing novel NFAs and applying them
ORCID 3494 in organic solar cells. Particularly, dramatic progress has been
Author Contributions 3494 made in the NFA-based OPV field with over 400 publications
Notes 3494 in the past four years (Figure 1). Over 10% PCEs have been
Biographies 3494 reported for organic solar cells based on different combinations
Acknowledgments 3495 of polymer and NFA, with the best PCE up to 13.1% for single-
References 3495 junction device19 and 13.8% for tandem structure.20
Besides the high and fast-growing PCE (Figure 1), OPV
devices based on NFAs manifest other distinct features that
1. INTRODUCTION have not been widely recognized in fullerene based OPVs. For
The growing demand for renewable energy sources has caused example, the fraction of publications on NFA-based OPVs that
the rapid development in photovoltaic (PV) technologies. Solar mention a reduction in voltage loss is much higher than that on
energy is the largest carbon-neutral energy source available fullerene-based OPVs. 21−23 Evidence suggests that the
today1 and is the fastest-growing form of renewable energy.2 commonly believed energetic offset for generating free charge
Among various PV technologies, organic photovoltaics (OPVs), carriers in OPVs may not be a constraint in NFA-based
whose building blocks are based on earth abundant nontoxic systems.23,24 Such low voltage loss could significantly improve
materials, have demonstrated a short energy payback time the overall performance of OPVs. Besides low voltage loss,
(EPBT)3,4 and a great potential to reduce the levelized cost of enhanced device stability has also been reported for NFA solar
energy (LCOE).5,6 In addition, OPVs are flexible and can be cells.19,25−37
3448 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

goal of sections 4−6 is to compare the effects of different


molecular design strategies on device performance, to generate
insights from previous work, and to outline directions for
further improvement, which are captured by a discussion
section (section 7). With sections 4−7 focusing on the
polymer:NFA based devices, section 8 summarizes the limited
but promising result on all small-molecule nonfullerene solar
cells, i.e., BHJ solar cells based on a small molecule donor and a
nonfullerene small molecule acceptor. In section 9, we extend
our discussion beyond binary single-junction BHJ and
summarize the role of NFA in ternary blend OPVs. As the
maximum reported PCE reached above 13%, OPV device’s
operational stability becomes increasingly vital for the objective
of commercialization. Therefore, section 10 summarizes the
results on the improved stability of NFA-based binary or
ternary BHJ solar cells with the topic semitransparent OPV also
Figure 1. Nonfullerene organic solar cell publications each year (via covered. Section 11 delineates the voltage loss in OPV and
Web of Science) and the maximum power conversion efficiency highlights the potential of using NFA to achieve a low voltage
reported each year up to Aug 2017. loss system. Section 12 provides perspectives on molecular
design and synthetic complexity of NFAs and highlights the
Despite these exciting results, problems and challenges problems and challenges facing the field, mainly from the
remain for this young field. For instance, molecular twisting is morphological and device physics points of view, where topics
introduced to prevent the molecules from overaggregation, but such as energy loss, active layer morphology, multijunction
it presumably hinders charge transport.23,35,38−43 This paradox device, interface engineering, and roll-to-roll printing are
is prominent in NFAs as they are not as diffusive as fullerenes discussed, with perspectives and outlook provided. Concluding
and they do not have the three-dimensional charge transport remarks are given in section 13.
capability in general. A thorough understanding of both
intramolecular and intermolecular interactions for NFAs in 2. BRIEF INTRODUCTION TO OPV WORKING
the pristine form as well as in the blend film is needed. Besides, PRINCIPLE
there lack systematic morphological studies on the general 2.1. Light Absorption and Exciton Generation
morphology forming process and device physics studies on
In order to harvest sunlight for the generation of electrical
exciton dissociation at the polymer/NFA interface, which are
energy, a solar cell has to convert photons into free charge
key issues to be addressed for further advancement of
carriers and be able to transport these carriers to the electrodes
photovoltaic performance.
to give photovoltage and photocurrent. Molecular electronic
By the means of summarizing material property and
materials such as π-conjugated polymers are ideal solar light-
photovoltaic performance, this review aims to provide a
harvesting systems due to their (1) large extinction coefficients
balanced assessment of the state-of-the-art molecular design
as a result of the large wavefunction overlap between the
strategies, to facilitate discussions on how to better understand
electronic ground state and the lowest excited state54,55 and (2)
the structure−property relationship for further optimizing
broad absorption bands as a result of the significant geometry
active layer morphology, to highlight the fundamental materi-
relaxations that take place in the excited state.55−57 The intense
al/device properties that lack understanding, and to delineate
absorption bands over a broad wavelength range enable a good
major challenges facing the NFA-based OPV field. The
matching with a sizable portion of the solar spectrum for
emphasis of this review is the solution-processed BHJ organic
efficient light harvesting in relatively thin layers (∼100−200
solar cell application of nonfullerene molecular electron
nm).
accepting materials. Consequently, polymeric electron accept-
In a neutral organic molecule, the highest occupied molecular
ors44−53 and the design of novel polymer donors to be
orbital (HOMO) in the ground state configuration contains
compatible with NFAs are beyond the scope of this review.
two electrons with opposite spins. The absorption of an
Similarly, the application of NFAs in other organic electronics,
incident photon promotes an electron in the HOMO into the
such as organic field effect transistors (OFETs), organic light
lowest unoccupied molecular orbital (LUMO), leaving behind a
emitting diodes (OLEDs), and organic rectifiers, are not
hole to form an electron−hole pair (Figure 2). In an ideal
covered in this review.
photovoltaic system, the electron−hole pair can easily separate
This review begins with a brief introduction to OPV working
to form long-lived, free charge carriers with high quantum
principle (section 2), where light absorption, exciton
yields, such that a large photocurrent is created when they are
dissociation, charge generation, and charge transport are
collected by the electrodes. However, in practice such efficient
covered and the concept of bulk-heterojunction is discussed.
free carrier generation requires the electron−hole pair to
As a consequence of the limited exciton diffusion length, BHJ
overcome their mutual Coulomb attraction, V, which is given
solar cells with specific morphology requirement are necessary by
for realizing the full potential of the materials. In section 3, an
overview of the molecular design strategy is provided. Then we e2
categorize NFAs based upon structural similarities and provide V=
4πε0εrr (1)
detailed material/device property summaries for each category
(sections 4−6), with a particular focus on recent materials where e is the electronic charge, ε0 is the permittivity of
representing top-of-the-line performance in OPV devices. The vacuum, εr is the dielectric constant of the surrounding
3449 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

the material by random diffusion. This incoherent hopping


process is typically driven by both Förster resonance energy
transfer (FRET) and Dexter energy transfer and is unaffected
by electric fields due to the electrical neutrality of excitons. This
energy transfer process occurs either intramolecularly or
intermolecularly following a downhill energy gradient and
may lead to trapping of the exciton at trap or defect sites that
leads to inhomogeneous broadening of the density of states.
Further migration of the exciton energy relies on thermal
fluctuation. Organic films typically have short singlet exciton
lifetimes before they decay to the ground state, and thus they
have relatively short diffusion lengths (in the range of ∼5−20
nm). Therefore, a compromise with regards to the film
thickness is required to optimize both exciton dissociation
(favoring thin layers) and light absorption (favoring thick
layers).
Figure 2. Electronic state diagram describing the photoinduced
charge-carrier formation mechanism in an organic solar cell. Adapted
A key breakthrough that overcomes the limitation of short
with permission from ref 55. Copyright (2009) American Chemical exciton diffusion lengths of organic materials was the
Society. introduction of the bulk heterojunction.7,64 Bulk heterojunc-
tions comprise a bicontinuous interpenetrating network of the
donor and acceptor material, which can be formed by solution
medium, and r is the distance between the electron−hole pair. processing or by coevaporation of the two materials. Due to the
In solar cells based on inorganic semiconductors, such as silicon large interfacial area between donor and acceptor, the distance
p−n junctions, electron−hole pairs can easily overcome their that the excitons have to travel before reaching an interface is
Coulomb attraction to generate free charges due to the high significantly decreased, and efficient harvesting of exciton
dielectric constant (εr ≈ 12 in silicon) and the highly energies can therefore be achieved when the nanoscale phase
delocalized nature of the photoexcited states (corresponds to segregation between the two materials matches with the
a large r in eq 1).56,58,59 However, overcoming this Coulomb corresponding exciton diffusion lengths. For planar hetero-
attraction between electron−hole pairs is much more junctions, exciton harvesting can also be improved by means of
challenging in organic materials due to the low dielectric enhancing exciton diffusion lengths and/or using Förster
constant (εr ≈ 2−4) and the more localized nature of the resonance energy transfer (FRET) in multilayer structures.65,66
photoexcited states. Nevertheless, we will focus on bulk heterojunction device
As a result of the strong Coulomb interaction between structures in this review.
electrons and holes, as well as strong electron−lattice and 2.2. Charge-Transfer State and Free Ccarrier Generation
electron−electron interactions, photoexcitation of organic
conjugated materials generates a tightly bound electron−hole Although the charge generation process following exciton
pair known as an exciton. These primary photoexcitations have dissociation is fundamental to the operation of organic solar
singlet (spin-zero) character due to the conservation of spin. In cells, its mechanism has remained a disputed topic in the
order to facilitate efficient charge generation, it is necessary to literature. In this section, we briefly summarize the key
separate these singlet excitons by overcoming its binding experimental findings and theoretical models that help shed
energy. Typical binding energies of singlet excitons in organic light on this topic.
materials have been reported to be ∼0.5 eV.55 This binding Upon reaching a donor−acceptor interface, excitons can
energy is significantly greater than the thermal energy kBT at dissociate following the transfer of an electron (hole) from the
room temperature (25 meV), and thus an additional driving donor (acceptor) material to the acceptor (donor) material.
energy is required to create free charges. Historically, this charge transfer process of OPV is described
The ubiquitous pathway to effectively separate excitons into within a modified Marcus framework, which considers the
free carriers is by using a heterojunction between an electron- tunnelling of point like charges.56,67 This initial electron (or
donating (donor) and electron-accepting (acceptor) material. hole) transfer process generates an electron−hole pair across
The differing electron affinity (and/or ionization potentials) the interface, which is commonly known as a charge-transfer
between the two materials creates an energetic offset at their state (CTS). Despite being on separate material domains, the
interface and thereby provides a driving energy for exciton electron and hole are still 0.5−1 nm apart such that they remain
dissociation. In 1986, Tang reported the first example of an bound by Coulomb binding energy. Once formed, the CTS can
OPV device based on a vacuum-deposited CuPc/perylene either overcome its binding energy and separate into free
derivative planar bilayer heterojunction.60 However, despite charge carriers or undergo relaxation and recombine geminately
achieving relatively efficient exciton dissociation, the overall to the ground state resulting in energy loss. The dissociation of
efficiency of such a planar heterojunction device is limited by CTS has often been described using the Onsager−Braun
the requirement of exciton migration to the donor−acceptor framework, which considers hopping of charges within a
interface. This is because the rate of photoinduced electron disordered density of states by thermal activation.
transfer decreases exponentially with the separation between However, recent studies have challenged the application of
the donor and acceptor molecule, and therefore excitons must such a theoretical framework to describe charge generation
migrate to the interface in order to dissociate into charges.61−63 processes in high performance OPV devices. These include
Due to the localized nature of excited states in organic experimental studies showing that charge generation yield is
materials, photogenerated excitons typically migrate through independent of electric field and temperature, and dissociation
3450 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

may happens on ultrafast generation time scales (within 100 While charge generation in a wide range of OPV systems
fs).68 The difficulty in developing a general framework for based on fullerene acceptors has been studied over the years,
charge generation in OPV is in part due to the sheer variation careful examination of those based on nonfullerene acceptors
in material systems and blend combinations that have different has remained limited. This research topic is an area of great
energetics, structural and morphological properties. In the interest driven by the rapid improvement of device efficiencies
absence of a comprehensive theoretical framework for CTS and reported with nonfullerene based systems in recent years. It is
its dissociation, material design of OPV has been primarily possible that the charge generation process in fullerene systems
driven by empirical design rules.69,70 It is generally believed that may not directly apply to nonfullerene systems. For instance,
a significant driving energy (>0.2 eV) provided by the recent studies have shown that very high charge dissociation
interfacial energy offset is required to facilitate efficient charge yields can be achieved in nonfullerene systems even in the
generation. This rule was developed based on the observation absence of a significant driving energy.87,88 These findings
that many OPV material systems (particular those based on represent a major advancement for OPV technology because
fullerene acceptors) suffer from low charge generation yields in the requirement of a driving energy to generate free carriers in
the absence of this energetic offset.71 turn limits the open-circuit voltage. A detailed discussion about
This general observation has led to the debate over whether voltage loss in OPVs is discussed in section 11. We note that
charge generation happens primarily through vibronically “hot” the driving energy should not be simply determined by taking
or “cold” CTS. Given the driving (or excess) energy involved in the LUMO energy difference between the donor and acceptor
the initial charge transfer process, it is often thought that a hot material because the energy of the frontier orbital is the
CTS may facilitate charge separation more easily compared to a property of isolated material, which does not account for
cold (relaxed) CTS by thermalizing into more loosely bound exciton binding energy and other interfacial effects69 and are
electron−hole pairs. These hot states should intuitively be less typically estimated by electrochemical measurements. Instead,
prone to geminate recombination compared to more tightly to evaluate the energy loss, a meaningful driving energy should
bound, cold CTS, provided that they can separate before they be evaluated spectroscopically (e.g., PDS)89 or electrically (e.g.,
relax vibronically. Jailaubekov et al. have shown that the FTPS-EQE)69 using the blend film or better an actual OPV
relaxation time of hot CTS in a planar heterojunction of copper device. (PDS and FTPS-EQE refer to photothermal deflection
phthalocyanine and fullerene acceptor happens within ∼1 ps.72 spectroscopy and Fourier-transform photocurrent spectroscopy
This concept implies that the dissociation yield should occur on external quantum efficiency, respectively.)
an ultrafast time scale and correlate with the amount of excess Due to the relatively weak and narrow absorption of fullerene
energy involved in the charge transfer reaction, and this notion derivatives (the archetypal acceptors used in a BHJ),
is supported by a number of studies.73−76 However, this historically, most discussions on charge generation focus on
concept has also been challenged by studies showing that the the photoinduced electron transfer from the “donor” to the
charge generation yield is independent of the energy of the “acceptor”, where the donor refers to the species with lower
absorbed photon, whether in the donor or acceptor, and thus ionization potential and the acceptor refers to that with higher
the excess energy does not play a significant role in promoting electron affinity. This charge generation process is termed
charge generation.77,78 In particular, Vandewal et al.77 have channel I. However, the mirror process, namely the channel II
shown that the same charge generation yield can be achieved charge generation, also occurs in a BHJ system, where the
even when the low-energy cold CTS is directly excited, which photoexcitation of the acceptor is followed by a hole transfer
strongly implies that free carriers are generated from the relaxed from the acceptor to the donor. With the rapid development of
intermolecular states without requiring any excess energies. nonfullerene acceptors, numbers of new small molecular or
Apart from energetics, the importance of structural polymeric acceptors have been realized with strong and
morphology at the donor−acceptor interface on charge complementary absorption to the donor materials. Conse-
generation is also highlighted in several studies. For instance, quently, equal attention needs to be placed on channel II
for OPV blends based on fullerene acceptors it has been charge generation. Experimental methods for probing channel
proposed that charge separation efficiency is improved in the I/II in a BHJ OPV system include device internal quantum
presence of fullerene aggregates near the interface.79−82 These efficiency measurements, photoluminescence quenching meas-
aggregates promote the delocalization of electronic wave urements, transient absorption spectroscopy measurements,
functions, which may have a significant role on the charge etc. For a detailed discussion on the charge generation
separation mechanism of OPV blends. In particular, Gelinas et pathways, readers can refer to the review article by Stoltzfus
al.83 concluded that the control of the availability of states in et al.90
the fullerene phase by wavefunction delocalization can lead to It is important to note that CTS can also travel along the
an ultrafast coherent electron motion in the fullerene via donor−acceptor interface. Recent work by Deotare et al. shows
ballistic transport, resulting in long-range electron−hole that the CTS can move over distances of 5−10 nm in a model
separation (∼4 nm within 50 fs). Such a long-range charge donor−acceptor system, driven by energetic disorder and
separation mechanism would also be less sensitive to interfacial diffusion to lower energy sites.91 Such distance is comparable to
energetic offset than that predicted by the Marcus/Onsager the exciton diffusion lengths in many organic materials and
framework.84 However, while CTS delocalization favors charge phase segregation of BHJ blends, and therefore it is possible
generation, excessive delocalization may also lead to reduced that CTS diffusion also plays an important role in the charge
device open-circuit voltage by providing more pathways generation process in OPV. For instance, CTS formed in
through which recombination can occur.85,86 This implies blends with large density of low energy sites at the interface
that there may exist a trade-off between charge generation and may have a greater chance of being trapped in those sites,
voltage loss and optimizing the delocalization of CTS is followed by recombination. We will review the various charge
necessary to balance photocurrent and photovoltage. recombination pathways in OPV in the next section.
3451 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

2.3. Charge Transport state. Although triplet excitons have long lifetimes (up to μs)
Free charge carriers are generated upon the dissociation of and may be able to thermalize back into charges if given enough
excitons and/or CTS at the donor−acceptor interface, and they time, it was found that, once formed, these triplet excitons are
move toward their respective electrodes driven by the applied rapidly quenched by surrounding charges (triplet-charge
electric field. The efficiency at which the charges are annihilation) which subsequently lead to energy loss. It is
transported depends on their mobilities. While the ordered most likely that triplet recombination occurs in blends where
nature of inorganic semiconductors lead to high charge low-energy, relaxed CTS are formed.97−99 In the absence of
mobilities (typically ∼102 cm2 V−1 s−1), organic semi- these relaxed CTS, the separation of CTS into free charge
conductors tend to exhibit relatively low charge mobilities as carriers is kinetically more favorable compared to triplet exciton
a result of disorder effects, weak electronic couplings, and large formation, and thus triplet recombination is suppressed. On the
electron-vibration couplings (leading to polaron formation).55 other hand, in systems where triplets are allowed to form,
Charge transport thus relies on polarons hopping from site to bimolecular triplet recombination on subnanosecond time
site, with carrier mobilities strongly depending upon morphol- scales represents a major loss pathway that limits device
ogy that can range over several orders of magnitude from efficiency.
10−6−10−3 cm2 V−1 s−1 (highly disordered amorphous films) to It is thus clear that recombination events in OPV are
above 1 cm2 V−1 s−1 (highly ordered materials). Photocurrent is governed by microscopic details of the interface between donor
formed when the charge reaches the electrode, and efficient and acceptor molecules, and improved device performance can
charge extraction at the electrode is required to suppress be achieved by understanding the origins of these processes and
recombination.92,93 developing strategies to limit them.
During the charge transport process, free electrons and holes
may encounter near an interface and form a CTS, which then 3. MOLECULAR DESIGN PRINCIPLE
either recombines (through vibronic coupling to the ground- Historically, the main light absorber in a polymer/fullerene BHJ
state) or dissociates back into free carriers. This recombination blend is the p-type polymer because of the low absorptivity of
process is detrimental to OPV operation by limiting both the fullerenes, which makes the use of the words “donor” and
photocurrent and photovoltage. In the limit of Langevin “acceptor” specific to electrons. We follow this nomenclature as
recombination, as in the case of organic light-emitting diodes it was widely recognized by the field. However, many NFAs
(OLED), every encounter leads to recombination to the contain strong dye-based chromophores that could make them
ground state with no probability of separating once again into an even stronger light absorber than polymers and achieve
free carriers. However, it is well-known that the rate of significant photocurrent generation through hole transfer to the
recombination in OPV is several orders of magnitude slower electron donor material after photoexcitation.90 This is one of
than predicted by Langevin theory, leading to the need of the original motivations of utilizing NFAs, which could
introducing a “reduced Langevin factor” to correctly model significantly enhance the photocurrent as both channels
charge kinetics in OPV.92 This raises the question of whether (donor and acceptor) can contribute to free carrier generation.
Langevin theory mispredicts the recombination rate in OPV To maximize this effect, the BHJ should be designed using
because it overestimates the frequency of the electron−hole donors and acceptors that have broad and complementary
encounter or because only a small fraction of encounters spectral coverage. Thanks to the development of the polymeric
eventually lead to recombination. As explained by Burke et al.,86 and small molecular donor materials in the past decades, there
understanding the origin of this reduction factor has important exists a large library of donor materials to choose from or
consequences for OPV design. For instance, will increasing modify to this end.
carrier mobility in order to improve fill-factor in turn reduce The design rules of conjugated polymers are also valuable for
open circuit voltage (VOC) by making carriers recombine the development of novel NFAs. Many approaches can be
quicker? While it remains difficult to precisely control mobility, employed to tailor the optical and electrochemical properties
higher OPV efficiencies are typically achieved with higher such as adjusting the conjugation lengths to tune the spectral
carrier mobilities because the fill-factor improves without a response, using fluorination to alter the frontier energy levels,
corresponding loss in VOC.94 This implies that the limiting and tuning the extent of the HOMO−LUMO overlap to
factor for recombination is not the frequency of electron−hole modify the extinction coefficients. As described in previous
encounters but instead the rate at which the CTSs recombine sections, there has been evidence suggesting the traditionally
to the ground state and how quickly they can dissociate back believed >0.2 eV excess energy might not be a strict restriction
into free carriers.77,86,95 for efficient free carrier generation. Therefore, BHJ design
Recombination can also occur through spin-triplet states. should not be limited to a pair of materials with a specific
Similar to excitons, CTS may have spin-singlet or spin-triplet LUMO−LUMO (or HOMO−HOMO) offsets. Here, we
properties, with the difference being that the energy difference should note that we use the term LUMO level to indicate
(or exchange energy) between singlet and triplet states of CTS the electron affinity energy, which follows the standard usage in
is much smaller than that typically found for excitons. This is discussing organic electronics, but it needs to bear in mind that
due to the weak electronic coupling of CTS.75,96 The triplet the occupation of this level will cause the energy to change due
CTS states may be populated either geminately (following to a combination of Coulomb forces, exchange, and vibronic
intersystem crossing on nanosecond time scales) or bimolec- interactions.100
ularly (following encounter of free carriers with opposite In contrast to energetics, designing a BHJ with favorable
charges). The detrimental role of triplet CTS on charge nanoscale phase separation is far more complicated. A planar
dissociation yield of OPV is well documented.56 Previous geometry with extended conjugation is typically beneficial for
studies have shown that, in blends where there exists a lower- light absorption of the molecule, but it could induce a strong
lying triplet exciton state (say in the donor polymer domain), aggregation tendency that leads to excessively large domains
the energy of the CTS can be transferred to the triplet exciton with an inadequate interface between the donor and acceptor
3452 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

for efficient exciton dissociation. Therefore, a molecular design OPVs, OFETs,112,113 light harvesting arrays,114 and even
strategy, e.g., reducing the coplanarity, must be considered to sodium-ion batteries as organic electrodes.115
introduce enough miscibility and solubility to the molecule. Compared to PDI, its higher homologues such as terrylene
However, reducing the coplanarity of the molecule is typically and quaterrylene tetracarboxdiimides (TDIs and QDIs), have
accompanied by a reduction of the charge carrier mobility. relatively less flexible synthesis and a much higher propensity to
Therefore, in addition to energetics and absorption, the essence crystallize (Figure 3b−d) into a large domain with sizes of
of molecular design is to judiciously control the molecular hundreds of micrometers,103 which make them candidates for
geometry to balance exciton dissociation and charge carrier organic transistor applications116 but less suitable for BHJ-type
transport in a BHJ blend. OPV device applications.
Furthermore, chemical modification needs to retain the In terms of OPV application, naphthalene diimides (NDIs)
chemical, thermal, and photostability of the molecule. Mean- and PDIs are the most widely studied rylene derivatives. The
while, enabling solution processability, preferentially in environ- earliest OPV application of PDIs can be traced back to the
mentally friendly solvents, is another design consideration for pioneering work by Tang.60 Compared to PDIs, NDIs are less
the eventual commercialization of this technology. crystalline in general and have relatively weaker absorption in
the visible, which make them more applied as building blocks in
4. RYLENE DIIMIDE-BASED NFAS polymeric acceptors.44−47,51−53 This section summarizes the
development of PDIs, NDIs, and other rylene derivatives as
Rylene diimides are a robust, versatile category of polycyclic
electron accepting materials in BHJ type OPVs, with a
aromatic n-type (or ambipolar) materials with remarkable
particular focus on the most recent publications.
electron mobility, high electron affinity, high absorption
coefficients, and outstanding oxidative/thermal stability,101,102 4.1. PDI-Based NFAs
which make them ideal candidates for application in a wide For BHJ OPV applications, shown in Figure 3b, the PDI
range of organic electronics. monomer still has too strong an aggregation for achieving
Among various rylene dyes, perylene dyes have been known domain sizes small enough for efficient exciton dissociation.
for over a century. Perylene diimides (PDIs) have generally Therefore, functionalization of PDI on various positions is
been used as pigments and dyes. Owing to their rigid aromatic introduced to make the molecule twisted, which is essential for
core with multiple grafting sites, PDIs show strong propensity reducing aggregation. Depending on the position of function-
to self-assemble into ordered bars, spheres, or ordered alization, the PDI based NFAs can be categorized into bay, α,
structures in other shapes.104−109 Thanks to their exceptional and imide functionalized derivatives (Figure 4). Among them,
electronic and optical properties (Figure 3a) along with their
outstanding thermal, oxidative, and photostability, PDI-based
semiconductors have been widely applied to OLEDs,110,111

Figure 4. Functionalization positions on a PDI.

the bay position is the most explored grafting site due to the
feasible synthesis and effectiveness to reduce intermolecular
aggregation. Connecting different numbers of PDIs through a
central core has also been proved an effective approach to
introduce twisting to the molecule. These approaches have
caused differences in intramolecular twisting and/or intermo-
lecular interaction between neighboring PDIs, whereas an
expected outcome of the twisting approach is the hindered
charge transport within the less crystalline network. Natural
questions raised are how much twisting is needed? Were any of
the molecules overtwisted? Would there be alternative method
to maintain good intermolecular interaction while reducing the
Figure 3. (a) Absorption spectra of PDI and its higher homologues.
Adapted with permission from ref 102. Copyright (2010) American
domain size? These are concerns to bear in mind at this stage of
Chemical Society. 2D-WAXS patterns of (b) PDI 1, (c) TDI 2, and PDI research. On the other hand, α position functionalization
(d) QDI 3. The numbers after PDI, TDI, and QDI refer to the original and fused-ring strategies emerge as promising approaches to
names of the molecules given by the authors of ref 103. (b−d) address the twisting-coplanarity paradox. Subsection 4.1 begins
Adapted with permission from ref 103. Copyright (2006) American with recapping the early application of PDI-based NFAs in
Chemical Society. OPVs such as the monomeric PDIs, followed by summarizing
3453 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

various twisting strategies, focusing on the bay-position A simple PDI dimer linked at the bay region through a single
functionalized PDIs. Fused-ring strategies and α position bond was reported later by Jiang et al.132 Among the PDI
functionalization are then discussed. Chemical structures of dimers connected through one, two or three singles bonds, the
selected PDI-based materials are shown in each subsection with one with a single bond between two PDIs (s-diPBI, Figure 7,
their material properties and OPV device parameters molecule 4.7) resulted in a more twisted molecular structure
summarized in Table 1. than those with two or three single bonds. Consequently, an
4.1.1. PDI Monomers: Requirement for Twisting. OPV device based on s-diPBI exhibited a high PCE of 3.63%
Pioneering work employed unsubstituted PDI dyes in OPVs. with PBDTTT-C-T as the donor polymer and a mixture of 1,8-
To enable solution processability, only imide-functionalized diiodooctane (DIO)/1-chloronaphthalene (CN) as solvent
PDIs were used in the beginning (Figure 5).117,118 Schmidt- additive whereas PDI dimers linked by two or three single
Mende et al. demonstrated using discotic liquid crystal hexa- bonds displayed significantly lower device performance. Further
perihexabenzocoronene in combination with a monomeric PDI improvement of the device performance was achieved by Zang
dye to produce a thin film with vertically segregated donor and et al. through combining s-diPBI with PTB7-Th. A high JSC of
acceptor materials.119 An EQE >34% was demonstrated for a 11.98 mA cm−2 and a high fill factor (FF) of 59% contributed
thickness of almost 490 nm. The low PCE, particularly the low to an excellent PCE of 5.9%.133 Later, Sun et al.134 and Meng et
short circuit current (JSC) indicates the strong aggregation al.135 reported two derivatives of s-diPBI with bay regions at
problem. In 2013, Sharenko et al. reported the use of p- both ends of the dimer end-capped by either sulfur (SdiPBI-S,
DTS(FBTTh2)2 and PDI to fabricate all-small-molecule solar Figure 7, molecule 4.8) or selenium (SdiPBI-Se, Figure 7,
cells with a PCE up to 3.0%.120 In a later report, mixing PDI molecule 4.9), rendering the PDI unit a planar structure.
with PBDTTT-C-T (Figure 6) resulted in a PCE of 3.7%, Compared with s-diPBI, SdiPBI-S and SdiPBI-Se exhibited
which was the best performance of PDI monomers function- larger dihedral angles between the two PDI units, as well as
alized only at the imide position.121 larger optical bandgaps. When combined with a medium
To alleviate the overaggregation, substitutions at other bandgap donor polymer PDBT-T1 (Figure 6), SdiPBI-Se
positions were introduced.122−126 One representative report achieved a high PCE of 8.42% while SdiPBI-S showed a PCE of
was a slip-stacked PDI by introducing substituents at the α 7.16%.
positions.127 The best-performing acceptor phenyl-PDI (Figure In terms of synthetic cost, Hendsbee et al. demonstrated the
5, molecule 4.3) was obtained by adding a phenyl group at each gram-scale synthesis of three N-annulated PDI compounds and
α position, and the steric hindrance of the substituents caused a a PDI dimer linked via a single bond at the bay position (Figure
slipped stack of the PDI planes. As a result, a PCE up to 3.67% 7, molecule 4.10), without the need for purification using
was obtained by combining with a large bandgap donor column chromatography.136 One of the PDI dimers exhibited a
polymer PBTI3T (Figure 6). In another report, introducing PCE of 7.55% in BHJ solar cell.
phenyl groups at the bay region was also reported.128 The In addition to connecting two PDIs through a single bond, to
acceptor TP-PDI (Figure 5, molecule 4.4) achieved a PCE of further tune the morphology of the active layer, Yan et al.
4.1% when blended with PTB7-Th (Figure 6). studied the effect of spacers on aggregation and thus phase
Despite increased PCE, the overaggregation is still one of the separation between PDI and polymer. A series of spacers
key issues limiting the device performance in monomeric PDI- including thiophene, benzene, bithiophene and spirobifluorene
based BHJ solar cells. To reduce the aggregation tendency, were inserted between the two PDI units.137 It was found that
various molecular design strategies have been developed that the planar dimer with a phenyl ring as the spacer showed the
led to dramatic improved photovoltaic performance ever since. largest domain size and the worst solar cell performance due to
4.1.2. Twisting Strategy-PDI Dimers. Other than strong intermolecular interactions. In contrast, the best PDI
functionalizing a PDI monomer, one straightforward approach dimer in this report was the one with spirobifluorene as the
to reduce the planarity is to connect two PDIs together through spacer (SF-PDI2), which led to a small domain size of the blend
a single bond or the aid of a spacer unit (Figure 7). film and a PCE of 2.35% with P3HT as the donor polymer.
Pioneering research in this area was reported by Rajaram et In 2016, Liu et al. designed a new conjugated polymer,
al., who linked two PDIs with a single bond between the two P3TEA (Figure 6), and combined it with SF-PDI2 (Figure 7,
nitrogen atoms at each PDI’s imide region,129 forming a twisted molecule 4.11) in a BHJ solar cell.87 The nonfullerene solar cell
PDI dimer (Per 1). Compared to the PDI monomer (Per 2, see demonstrated not only a 9.5% efficiency with nearly 90%
Figure 5, molecule 4.2), the reduced planarity of Per 1 (Figure internal quantum efficiency but also an extremely high VOC
7, molecule 4.5) suppressed the excessive aggregation and (1.11 V) and a decent FF (0.64). The authors quantified the
boosted the photocurrent dramatically (Figure 8). A high PCE energy loss of the device by taking the difference between the
of 2.77% was obtained for Per 1 with PBDTTT-C-T as a low- bandgap of the donor (obtained using the crossing point
bandgap donor polymer. between the absorption and emission spectra) and the VOC of
The similar strategy was also used by Zhang et al., who the device, and a low voltage loss (0.61 V) was determined.
presented a PDI dimer linked at the bay region with a Noticeably, subgap-EQE and electroluminescence measure-
thiophene used as the spacer (Bis-PDI-T-EG, Figure 7, ments exhibited almost overlapped spectra between the pure
molecule 4.6).130 Such dimerization, together with the side polymer and the polymer:SF-PDI2 blend (Figure 9a,b), which
chains at the bay region, suppressed the intermolecular indicated a nearly zero driving force for charge separation in
interaction and aggregation of PDI. An outstanding PCE of this highly efficient OPV system. In addition, the authors
4.03% was achieved using PBDTTT-C-T as the donor polymer. employed another polymer:NFA system with greater driving
In a follow-up study, a new acceptor with shortened side chains force (PffBT4T-2DT:SF-PDI2) as a control group, and
at the bay region enabled a higher PCE of 4.34% in demonstrated that P3TEA:SF-PDI2 showed roughly identical
combination with the same donor polymer through tuning exciton dynamics, indicating an ultrafast charge separation in
the fraction of additive in the solvent mixture.131 spite of its minimal driving force. Moreover, through analyzing
3454 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Table 1. Summary of Representative Rylene Diimide-Based Acceptors
mobility (cm2 V‑1 s‑1)
internal LUMO HOMO Eg,opt D:A PCE VOC JSC device
acceptor ref (eV) (eV) (eV) donor ratio processing solvent (%) (V) (mA cm‑2) FF (%) E (neat) E (blend) H (blend) structure ref
phenyl-PDI 4.3 −4.01b −6.02c 2.01 PBTI3T 1:1 CF+0.5%DIO 3.67 1.024 6.56 54.59 2.8 × 10−3e NR NR inverted 127
TP-PDI 4.4 −3.82a −5.69a 1.89 PTB7-Th 1:1 DCB+1%CN 4.1 0.87 10.1 46.4 NR NR NR inverted 128
Chemical Reviews

Per 1 4.5 −4.06b −6.12d 2.068 PBDTTT-C-T 1:1 CB 2.78 0.76 7.9 46 8 × 10−3e NR NR inverted 129
Bis-PDI-T- 4.6a −3.84b −5.65b 1.69 PBDTTT-C-T 1:1 DCB+5%DIO 4.03 0.85 8.86 54.1 3.9 × 10−3e 1.0 × 10−3 3.0 × 10−3 conventional 130
EG
s-diPBI 4.7 −3.87a −5.95a 2.08 PBDTTT-C-T 1:1 DCB+1.5%DIO 3.63 0.73 10.58 46.80 3.21 × 10−5e NR NR conventional 132
+1.5%CN
a c
SdiPBI-S 4.8 −3.85 −6.05 2.20 PDBT-T1 1:1 CB+0.75%DIO 7.16 0.90 11.98 66.1 3.2 × 10−3 2.8 × 10−3 1.2 × 10−3 conventional 134
SdiPBI-Se 4.9 −3.87d −6.09d 2.22 PDBT-T1 1:1 CB+0.25%DIO 8.42 0.96 12.49 70.2 6.4 × 10−3 4.8 × 10−3 3.6 × 10−3 conventional 135
7b 4.10 −3.8b −6.0b 2.2 P3TEA 1:1.5 TMB+2.5%ODT 7.55 1.13 11.03 61 NR >10−7 NR inverted 136
SF-PDI2 4.11 −3.71b −5.71c 2.03f P3HT 1:1 DCB 2.35 0.61 5.92 65 NR 7.1 × 10−5 NR inverted 137
IDT-2PDI 4.12 −3.83a −5.53a 1.54 BDT-2DPP 1:1 DCB 3.12 0.95 7.75 42.4 NR 2.3 × 10−6 2.0 × 10−5 conventional 138
P1TP 4.13 −3.72a −5.66a 1.8 PBDTTT-C-T 1:1 DCB+3%DIO 3.61 0.89 7.78 52.1 NR 2.4 × 10−4 2.0 × 10−5 conventional 139
i-Me2T2-PDI2 4.14 −3.57b −5.55c 1.98 PffBT4T-2DT 1:1.4 CB 4.1 0.91 8.0 56 NR 8.3 × 10−5 1.8 × 10−3 inverted 140
CP-V 4.15 −3.98a −6.16c 2.18 PPDT2FBT 1:2 CF+DPE 5.28 0.87 10.04 60.16 NR 1.39 × 10−5 1.07 × 10−4 inverted 141
F2B-T2PDI 4.16 −3.83a −5.94a 1.77 PTB7-Th 1:1.2 NR 5.05 0.84 10.60 57 NR 3.4 × 10−5 9.8 × 10−5 inverted 142
S(TPA-PDI) 4.17 −3.70a −5.40a 1.76 PBDTTT-C-T 1:1 DCB+5%DIO 3.32 0.88 11.92 33.6 3.0 × 10−5 2.32 × 10−5 7.17 × 10−4 conventional 143
H-tri-PDI 4.18 −3.93b −6.01b 2.09 PBDT-TS1 1:1 CB+7%DPE 7.25 0.732 16.52 60.03 NR 1.4 × 10−5 1.2 × 10−4 inverted 144
B(PDI)3 4.19 −3.86b −6.0b 2.03f PTB7-Th 1:1.5 CB+3%CN 5.65 0.83 13.12 52 NR 4.20 × 10−5 1.748 × 10−4 inverted 145

3455
Ta-PDI 4.20 −3.81a −6.03a 2.05 PTB7-Th 1:1 CB+0.5%CN 9.15 0.78 17.10 68.5 NR 2.7 × 10−4 3.6 × 10−4 inverted 38
TPE-PDI4 4.21 −3.72b −5.77c 2.05 PTB7-Th 1:1.4 CB 5.53 0.91 11.7 52 1 × 10−3 NR NR inverted 146
Tetra-PDI 4.22 −4.00b −5.97c 1.97 PTB7-Th 1:1.2 DCB+3%CN 3.54 0.86 8.39 49 1.4 × 10−3e 8.66 × 10−5 1.88 × 10−3 inverted 147
TPC-PDI4 4.23 −3.75b −6.00c 2.25 PffBT4T-2DT 1:1.5 CB 4.3 0.96 9.2 49 2.8 × 10−4 NR NR inverted 148
TPPz-PDI4 4.24 −3.76b −5.86c 2.10 PffBT-T3(1,2)-2 1:1.5 CB+DCB(7:3) 7.1 0.987 12.5 56 2.3 × 10−3 NR NR inverted 35
Me-PDI4 4.25 −3.82b −5.96c 2.14 PBDTTT-C-T 1:1 DCB+3%DIO 2.73 0.77 7.83 45.0 NR 1.78 × 10−6 5.55 × 10−5 inverted 149
SF-PDI4 4.26 −3.78a −5.97a 2.05 PV4T2FBT 1:0.8 CB+2%DIO 5.98 0.90 12.02 54.2 NR 1.93 × 10−5 2.46 × 10−4 inverted 150
P4N4 4.27 −3.69a −5.97a 2.11 PDBT-T1 1:1 DCB+0.5%DIO 5.71 0.958 9.40 63.4 2.69 × 10−3 1.90 × 10−3 1.21 × 10−3 conventional 151
PBI-Por 4.28 −3.68a −5.46a 1.48 PBDB-T 1:1 CB+1%DIO 7.4 0.78 14.5 66 1.0 × 10−2e NR NR inverted 152
helical PDI 1 4.29 −3.77d −6.04d 2.14f PTB7-Th 3:7 CB+1%DIO+1% 6.05 0.803 13.3 56.6 NR 3.4 × 10−4 2.9 × 10−4 inverted 155
CN
FPDI-T 4.30 −3.77d −5.98d 2.22 PTB7-Th 1:2 CB+2%CN 6.72 0.94 12.48 58 NR 1.63 × 10−4 5.92 × 10−2 inverted 156
Ph2a 4.32 −4.03b −6.36b 2.10f PTB7-Th 1:2.25 CF+1%DIO 3.89 0.93 7.68 54.3 NR 4.6 × 10−5 NR inverted 157
FITP 4.34 −3.75b −5.48b 1.77f PTB7-Th 1.2:1 CB+2%CN 7.33 0.99 13.24 56 NR 3.66 × 10−4 5.60 × 10−4 inverted 158
hPDI4 4.35 −3.91d −6.26d 2.00f PTB7-Th 1:1 CB+1%DIO 8.27 0.802 15.1 68.2 NR 1.5 × 10−5 1.2 × 10−4 inverted 160
hPDI3-Pyr- 4.36 −3.76b −5.77c 2.01 PTB7-Th 1:1 CB 7.6 0.80 15.1 62.9 NR 7.1 × 10−4 3.9 × 10−4 inverted 161
hPDI3
TPH-Se 4.37 −3.80b −5.97c 2.17 PDBT-T1 1:1 DCB+0.75%DIO 9.28 1.0 12.99 71.5 3.2 × 10−2e 2.2 × 10−3 1.7 × 10−3 conventional 162
βTPB6-C 4.38 −3.75d −6.21d 2.14f PTB7-Th 1:1.5 CB+2.5%DIO 7.69 0.92 14.9 56 NR 4.67 × 10−5 2.67 × 10−4 inverted 163
+2.5%DPE
Fused-TriPDI 4.39 −3.73b −5.95b 2.20 PTB7-Th 1:1.5 CB 6.19 0.91 12.39 55 1.26 × 10−3e 2.83 × 10−4 4.17 × 10−4 inverted 164
FTTB-PDI4 4.40 −3.58a −5.74a 1.88 P3TEA 1:1.5 TMB+2.5%ODT 10.58 1.13 13.89 65.9 2.2 × 10−4 1.1 × 10−4 1.5 × 10−4 inverted 165
αPBDT 4.41 −3.78b −5.60b 2.18f PTB7-Th 1:1.5 CB+3%CN 4.92 0.81 12.74 46 NR 8.00 × 10−4 1.79 × 10−5 inverted 166
TPB 4.42 −3.89d −5.71d 2.14f PTB7-Th 1:1 CB+8%DPE 8.47 0.79 18.40 58 NR 6.10 × 10−6 1.08 × 10−5 inverted 167
Review

DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Measured by CV using NFA films. bMeasured by CV in solutions. cCalculated using measured LUMO/HOMO and the optical bandgap. dDetailed method was not reported. eMeasured using OFET
devices. fDetermined from the film absorption onset by the authors of this review; NR Not reported. CF: chloroform; DIO:1,8-diiodooctane; DCB: 1,2-dichlorobenzene; CN: 1-chloronaphthalene; CB:
chlorobenzene; TMB: 1,2,4-trimethylbenzene; ODT: 1,8-octanedithiol; DPE: diphenyl ether; CV: cyclic voltammetry; E (neat): electron mobility in NFA neat films; E (blend): electron mobility in
168
173
174
ref

conventional
structure
device

inverted

inverted
3.40 × 10−5
H (blend)
NR
NR
mobility (cm2 V‑1 s‑1)

1.73 × 10−5
1.59 × 10−7
E (blend)
2.2 × 10−5

Figure 5. Chemical structures of monomeric PDIs.

the electroluminescence external quantum efficiency (EL-EQE)


of four combinations of polymer:NFA with different driving
forces (Figure 9c), the authors showed that a decreased driving
E (neat)

force corresponded to an increased EL-EQE. The increased EL-


0.365e

EQE can be further translated into a small nonradiative


NR

NR

recombination loss, which also contributed to the overall low


energy loss. Further details of the detailed balance theory that
FF (%)
52.36
59.1
46

the authors employed to delineate energy loss are discussed in


section 11). These results were the first example of efficient
(mA cm‑2)

charge separation on a small driving force in a nonfullerene


12.57

10.77
5.47

organic solar cell, which manifested the great potential of using


JSC

NFAs to achieve a low energy loss system.


The spacer approach stimulated the design of many other
0.78
0.75
0.73
VOC
(V)

PDI dimers with different spacers. For instance, when a bulky


IDT unit was utilized as the spacer, the PDI dimer named IDT-
2PDI138 (Figure 7, molecule 4.12) demonstrated different
PCE

5.43
2.41
3.64
(%)

performance in combination with different donors: IDT-2PDI


exhibited a higher electron mobility in the P3HT blend, which
processing solvent

resulted in an excellent FF of 67%. When a molecular donor,


DCB+0.5%DIO

BDT-2DPP (Figure 6), was employed, a higher PCE (3.12%)


CB+2%DIO

was achieved due to a more compatible electronic structure


between the donor and acceptor. In another report by Wang et
al., the authors studied the structure−property relationship
CB

among PDI dimers containing oligothiophene spacers with


different lengths.139 A longer spacer provided the PDI dimers
ratio
D:A

1:1.5

donor:NFA blend films; H (blend): hole mobility in donor:NFA blend films.


1:1
1:1

with lower bandgaps, stronger absorption, and higher energy


levels. Using the same donor polymer PBDTTT-C-T, the
dimer with one thiophene (P1TP, Figure 7, molecule 4.13)
exhibited the best OPV performance (3.61% PCE). It was
PBDT-TS1
donor
PTB7-Th

hypothesized that the dimer with no spacer had a highly twisted


PTB7

structure and a large domain size. On the other hand, too long a
spacer led to an overmixed morphology that was detrimental to
charge transport. In another report, Zhao et al. synthesized PDI
2.18f
Eg,opt
(eV)

2.39
1.68

dimers containing methyl substituted bithiophene spaces


(Figure 7, molecule 4.14) and compared the effect of
substitution position on film morphology.140 The dimer
HOMO

−5.71a
(eV)

−6.6b
−5.9c

whose substitution offered a head-to-head geometry exhibited


a better blend morphology with a smaller domain size and thus
a higher PCE (4.1%) in the BHJ device with PffBT4T-2DT (c)
LUMO

−3.88a
−3.7a
−4.0b

as the donor polymer. Park et al. reported a V-shape PDI dimer


(eV)

(CP-V, Figure 7, molecule 4.15) where the two PDIs were


connected through the imide position.141 Compared to an M-
Table 1. continued

internal

shape dimer connected through the bay position, CP-V


4.43
4.44
4.45
ref

exhibited a larger bandgap, a higher electron mobility, and a


better solar cell device performance (PCE up to 5.28%).
Hadmojo et al. inserted a 2,5-difluorobenzene unit into a
2,2′-Bi(PDI)
acceptor

bithiophene-bridged PDI dimer and investigated the effect of


BTDI3
BiNDI

this unit on material property.142 The resultant acceptor F2B-


T2PDI (Figure 7, molecule 4.16) exhibited a larger bandgap, a
a

3456 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 6. Chemical structure of selected p-type polymers used in combination with NFAs.

stronger twisting between the two PDI units, and a better 4.1.3. Twisting Strategy-PDI Trimers and Tetramers.
device performance than the original dimer. To further explore the twisting strategy and design NFAs that
3457 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 7. Chemical structures of selected PDI dimers.

could form three-dimension (3D) or quasi-3D molecular which leads to a higher electron mobility in the blend. With
structures, other routes such as combining three or more PTB7-Th as the donor polymer, the Ta-PDI based OPV device
PDIs in a molecule were explored (Figure 10). A pioneering showed an outstanding PCE of 9.18% with a high JSC of 17.1
study was reported by Lin et al., who synthesized a PDI trimer mA cm−2 and a high FF of 68.5%.
named S(TPA-PDI) (Figure 10, molecule 4.17), in which three Besides trimeric PDIs, tetrameric PDIs were developed by
PDI units were connected by a triphenylamine core.143 Due to connecting four PDI arms to a central core. Liu et al. designed a
the existence of a sp3 hybridized nitrogen atom at the center, tetrameric PDI molecule with a tetraphenylethylene core (TPE-
the whole molecule possesses a quasi-3D nonplanar structure. PDI4, Figure 10, molecule 4.21).146 As a result of the highly
Due to the suppressed intermolecular interaction and molecular twisted nature of the TPE core, the molecule exhibited a weak
aggregation, an efficiency of 3.32% was achieved when the PDI aggregation, which enabled a suitable phase separation and thus
tetramer was combined with a low bandgap polymer PBDTTT- a high PCE of 5.53% with PffBT4T-2DT as the donor polymer.
C-T. However, the electron mobility of S(TPA-PDI) in the Akin to this approach, a PDI tetramer based on a
blend film was only 2.32 × 10−5 cm2 V−1 s−1, which may explain tetraphenylsilane core was developed (Figure 10, molecule
the low FF (33.6%) of the device. In another study, a trimer by 4.22).147 The compact core made the PDIs interlocked, and a
elongation of Per 1 (H-tri-PDI, Figure 10, molecule 4.18) proof-of-concept device showed a PCE of 3.54% when blended
achieved a PCE of 7.25% and an excellent JSC of 16.52 mA cm−2 with PTB7-Th. Another study investigated the effect of the
with a high-performance donor polymer PBDT-TS1 (c).144 central atom of the tetrahedron core on film morphology and
Later, a trimeric PDI (B(PDI)3, Figure 10, molecule 4.19) with device performance, including carbon (TPC-PDI4, Figure 10,
the three PDI arms connected to a central benzene core was molecule 4.23), silicon (TPSi-PDI4), and germanium (TPGe-
designed by Li et al.145 The nonplanar molecule, when PDI4).148 It was shown that, using the same donor polymer
combined with the prototypical polymer PTB7-Th, exhibited PffBT4T-2DT, both TPC-PDI4 and TPSi-PDI4 could achieve a
a PCE of 5.65%. In a more recent report by Duan et al., a PDI high efficiency of >4%, while the germanium-based one showed
trimer with triazine as the core unit (Ta-PDI, Figure 10, an inferior performance (1.6%). Motivated by the PDI tetramer
molecule 4.20) was reported and compared with a benzene- TPE-PDI4,146 Lin et al. made a further modification to the
core analogue.38 It was found that the triazine-based trimer had molecule by using a tetraphenylpyrazine core, which reduced
a less twisted structure and thus higher aggregation propensity, the extent of twisting compared with TPE-PDI4, enhanced the
3458 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

was used as the core by Zhang et al. to construct a PDI


tetramer PBI-Por (Figure 10, molecule 4.28), which had a
significantly lower bandgap and weaker crystallinity than most
other PDI acceptors.152 A polymer donor PBDTBDD with
complementary absorption was used and a PCE of 7.4% was
achieved.
4.1.4. Discussion: How Much Twisting Is Needed? The
structure−property relationship discussed in refs 38 and 35 has
stimulated the demand to reexamine the twisting strategy.
Specifically, through examining the effect of the twisting angle
on device performance for the series of structurally similar PDI
tetramers35 (Figure 11), it is not difficult to conclude that the
enlarged nonplanarity does not necessarily correspond to an
enhanced device performance. These results indicate that when
the nanoscale phase separation is within a reasonable range, i.e.,
not too large for efficient exciton dissociation, the essential
intermolecular π−π stacking that could dramatically change the
local or long-range charge transport property starts to play a
dominant role in determining the overall device performance.
This indication bears substantial importance to the PDI-based
NFA research, where the majority of effort has been focused on
introducing nonplanarity in the past.
Similarly, regarding the comparison between Ta-PDI and Ph-
PDI in ref 38, the Ph-PDI molecule bears resemblance to a
Figure 8. (a, b) Optical images illustrating the overaggregation of
monomeric PDI (Per 2, molecule 4.2) and the reduced stacking of a
“three-wing propeller” with large twisting angles (76°, 54°, and
dimerized PDI (Per 1). (c) J−V characteristics of BHJ solar cells made 42°) between the PDI units; in contrast, two of the three PDI
with PBDTTT-C-T:Per 1 (Figure 7, molecule 4.5) and PBDTTT-C- units in Ta-PDI displayed a more coplanar geometry (19°)
T:Per 2. Adapted with permission from ref 129. Copyright (2012) while the third PDI unit showed a large twisting angle with
American Chemical Society. respect to the other two (83° and 74°; Figure 12). Nonetheless,
the blend films did not exhibit significant variation in terms of
electron mobility of the acceptor TPPz-PDI4 (Figure 10, phase separation, but the Ta-PDI demonstrated higher electron
molecule 4.24), and improved the efficiency to 7.1% with mobility and considerably higher PCE than Ph-PDI when
PffBT-T3(1,2)-2 (Figure 6) as the donor polymer.35 In combined with the same polymer.
addition to connecting the PDI units at the bay region, four 4.1.5. Fused-Ring PDI. Ring fusion is an effective route to
PDIs connected to a tetraphenylcarbon core unit via the imide enhance coplanarity of a molecule,153,154 which could lead to
region was also reported (Me-PDI4), which showed a PCE of enhanced molecular order that is beneficial for charge carrier
2.35% combined with the donor polymer PBDTTT-C-T.149 transport. As the previous work has demonstrated that the
Another PDI tetramer with spirobifluorene as the core unit chemists were able to introduce sufficient (in some cases even
(Figure 10, molecule 4.26) was also developed, exhibiting a too much) twisting,35,38 a proper extent of ring-fusion within
PCE of 5.98% when PV4T2FBT (Figure 6) was used as the the molecule could potentially improve charge transport while
donor polymer.150 Liu et al. reported another PDI tetramer maintaining adequate donor/acceptor interface for charge
with 2,3,7,8-tetraphenylpyrazino[2,3-g]quinoxaline as the core generation. Chemical structures of selected fused-ring PDIs
unit (Figure 10, molecule 4.27),151 which enabled a PCE up to are shown in Figure 13. An example was given by Zhong et al.,
5.71% with PDBT-T1 as the donor polymer. Later, a porphyrin who designed a helical PDI dimer with an ethylene spacer

Figure 9. (a) Normalized FTPS-EQE spectra of pure P3TEA and blend P3TEA:SF-PDI2 blend devices. (b) Normalized electroluminescence spectra
pure P3TEA and P3TEA:SF-PDI2 blend devices. (c) EL-EQE of P3TEA:SF-PDI2 (blend A), PffBT4T-2DT:SF-PDI2 (blend B), P3TEA:diPDI
(blend C), and PffBT4T-2DT:diPDI (blend D)-based solar cells at different voltages. The value of the driving force for blend A is negligible; for
blend B it is 160 meV; for blend C it is 200 meV; and for blend D it is 370 meV. Adapted with permission from ref 87. Copyright 2016 Nature
Publishing Group.

3459 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 10. Chemical structures of trimerized and tetramerized PDIs.

(Figure 13, molecule 4.29). The rigid yet nonplanar structure ring-fusion between the spacer and PDI increased the
enabled an appropriate phase separation when blended with intramolecular electronic coupling, which, in general, resulted
PTB7-Th.155 An ultrafast electron transfer from the donor to in an upshifted LUMO and therefore a blue-shifted absorption
the acceptor and hole transfer from the acceptor to the donor with sharp peaks (Figure 15). Along with increased electron
(∼0.2 ps) were revealed (Figure 14). As a result, an excellent mobilities, the molecule with a fused benzene spacer achieved
efficiency of 6.05% was achieved. In another report, three fused the highest PCE of 3.89%, with PTB7-Th as the donor
analogue PDI dimers were developed using furan, thiophene, or polymer. In a later report, two PDI units were fused with an
selenophene as the spacer.156 The dimer with thiophene as the IDTT bridge, which resulted in a fully planar dimer (Figure 13,
spacer was found to have the strongest intermolecular packing, molecule 4.34).158 Nevertheless, the PCE when combined with
the highest electron mobility, and the best performance PTB7-Th reached 7.33%, which was likely the result of the
(6.72%) with PTB7-Th. suppressed aggregation caused by the four bulky alkylphenyl
Hartnett et al. studied the effect of ring-fusing between the side chains.
PDI unit and the spacer for three series of PDI dimers with In addition to PDI dimers, linear PDI oligomers were
different spacers, namely thiophene (Figure 13, molecule 4.30), reported and have been used to further improve the
benzene (Figure 13, molecule 4.31−4.33), and thienothio- photovoltaic device performance. In 2014, Zhong et al.
phene.157 They found that, compared to nonfused PDI dimers, synthesized a series of PDI helical structures by fusing the
3460 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 11. Structures of TPC-PDI4, TPE-PDI4, and TPPz-PDI4 with a decreasing extent of intramolecular twisting. Reprinted with permission from
ref 35. Copyright 2016 John Wiley and Sons.

but also introduced more absorption bands with a signature


intense absorption peak in the higher-energy region, e.g., peak
at ∼385 nm for the dimer 2. Through DFT calculations and
femtosecond transient absorption spectroscopical studies, the
high-energy absorptions were attributed (at least partially) to
the electron being promoted from the olefinic bridges to the
LUMO situated on the PDI framework and/or the electron
being promoted from the HOMO situated on the PDI subunits
to the olefinic bridges.
The elongation of helical PDI 1 to fused PDI trimer and
tetramer increased the efficiency to over 8%.160 Similar to the
dimer, the trimer and tetramer were also nonplanar and rigid,
and their bandgaps decreased slightly with increasing number
of PDI units due to an extended conjugation. The strong
absorption and high mobility of h-PDI4 (Figure 13, molecule
4.35) enabled a PCE up to 8.3% with PTB7-Th as the donor.
Recently, a linear PDI oligomer with a pyrene connecting two
PDI trimers was developed (hPDI3-Pyr-hPDI3, Figure 13,
molecule 4.36).161 The ribbon had a molecular length of ∼5
nm and showed broad and strong absorption in the range of
300−650 nm, and the initial test with PTB7-Th achieved a PCE
of 7.6%.
Recently, a PDI trimer that combined design strategies
including (i) spacer fusion, (ii) functionalization of the end bay
regions, and (iii) construction of a 3D structure was reported
by Meng et al. (Figure 13, molecule 4.37).162 The three PDI
units, with selenium-annulated bay regions, were fused to a
benzene core. Due to a strong steric hindrance, the acceptor
TPH-Se exhibited a highly twisted three-bladed propeller
structure (Figure 17). The strong intermolecular interaction
contributed to a high electron mobility, which further led to an
excellent FF (71.5%) in the device when blended with PDBT-
T1. In addition, a broad and strong absorption enabled a high
Figure 12. Calculated optimal molecular geometries of Ta-PDI and JSC of 12.99 mA cm−2, and properly matched energy levels
Ph-PDI (top). Bottom: (a,b) atomic force microscopy (AFM) height contributed to a high VOC (1.0 V). Finally, an outstanding PCE
images and (c,d) phase images of PTB7-Th:Ta-PDI (a,c) and PTB7-
of 9.28% was achieved. Two other reports used a similar
Th:Ph-PDI films. (b,d) Image size: 5 × 5 μm2. Reprinted with
permission from ref 38. Copyright 2017 John Wiley and Sons. strategy. The first example was an acceptor βTPB6-C (Figure
13, molecule 4.38) with four PDI units fused to a BDT-Th unit,
which also showed a 3D structure.163 Compared to its nonfused
neighboring PDIs.159 Shown in Figure 16 below, different analogue acceptor, βTPB6-C exhibited a larger bandgap, higher
ribbon lengths corresponded to different conformations. For VOC and JSC, and a higher PCE with PTB7-Th as the donor
instance, the PDI tetramer showed two conformers while the polymer. The second example was the fused PDI trimer with
PDI tetramer had four isoenergetic conformations. Electroni- benzotrithiophene as the core unit developed by Wang et al.164
cally, ring-fusion not only narrowed the bandgap (Figure 16g,h) It was shown that the fused-TriPDI (Figure 13, molecule 4.39)
3461 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 13. Chemical structures of selected fused-ring PDIs. The blue color highlights the nonfused structure, while the green highlights the fused
structure.

molecule 4.40) with a “double-decker” molecular shape, which


demonstrated an outstanding PCE of 10.58% when combined
with the polymer P3TEA.165 Compared to the nonfused PDI
tetramer, ring-fusion provided FTTB-PDI4 with a strong
intermolecular stacking, a high electron mobility, a blue-shifted
absorption (more complementary with the polymer), and
Figure 14. Exciton generation and charge separation in the PTB7:1
blend at high excitation energy. Adapted with permission from ref 155. energy levels that are more compatible to the polymer. This
Copyright (2014) American Chemical Society. result further highlights the effectiveness of the ring-fusion
strategy in modulating the intramolecular property.
exhibited a larger bandgap, higher mobility, and better solar cell 4.1.6. α-Substituted PDI. The α-position, or sometimes
device performance than the nonfused analogue molecule. referred as the nonbay or ortho-position, is another grafting site
More recently, Zhang et al. used the ring-fusion strategy to that has been shown to have a different effect on the coplanarity
obtain a novel PDI tetramer named FTTB-PDI4 (Figure 13, of the functionalized molecule in comparison to the bay-
3462 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

addition, 2,2′-bi(PDI) exhibited a higher LUMO level and thus


a higher VOC of its BHJ solar cell. The overall PCEs were
similar for the two NFAs, when PTB7-Th was used as the
donor polymer.
4.2. Other Rylene-Based NFAs
In addition to PDI, its lower homologue NDI and higher
homologue such as TDI have also been employed to build
NFAs.169−172 A recent NDI-based NFA example was reported
by Liu et al. in 2015.173 The authors connected two NDI units
with a vinyl group. The NDI dimer, named BiNDI (Figure 19,
molecule 4.44), exhibited a near-planar structure and achieved
Figure 15. UV−vis absorption of Ph1, Ph2a, and Ph2b in ref 157. a PCE of 2.41% with PTB7 as the donor polymer.
Reproduced from ref 157 with permission from the Royal Society of In 2017, Feng et al. reported the first application of terrylene
Chemistry. derivatives as NFAs in organic solar cells.174 Similar to the
twisting strategy used to decrease the coplanarity of PDI-based
position.147 Introducing substitution at the α-positions has NFAs, a dimerized TDI and a trimerized TDI were synthesized
therefore been considered as an alternative route to address the with a best PCE of 3.64% achieved using the BTDI3 (Figure
twisting-coplanarity problem and develop efficient PDI-based 19, molecule 4.45) as the acceptor and PBDT-TS1 as the
NFAs (Figure 18). donor polymer.
Two pairs of PDI dimers, with benzodithiophene (BDT) or Compared to PDI, NDI-based NFAs typically exhibit a low
pyrene diimides (PID) as the spacers connected at either bay or absorption coefficient in the visible as well as low electron
α-position, were synthesized and studied by Zhao et al.166 For mobilities due to the shortened conjugated backbone. There-
both cases, the α-substituted PDIs (Figure 18, molecule 4.41) fore, NDI is often used as the electron-withdrawing moiety to
showed a stronger aggregation, a higher electron mobility, and design polymeric acceptors. On the other hand, TDI consists of
finally a better PV performance. The best efficiency (4.92%) more conjugated aromatic rings than PDI, which typically leads
was achieved with αPBDT when PTB7-Th was used as the to red-shifted, intense absorption, and a high electron mobility.
donor polymer. Recently, a BDT based core named TPB However, too strong an aggregation tendency may be
(Figure 18, molecule 4.42) was employed by Wu et al. to detrimental for efficient free carrier generation. Besides, the
construct α-substituted PDI tetramers.167 Due to a comple- synthetic complexity may be another reason limiting TDI’s
mentary absorption and carefully optimized film morphology, application.
the JSC of the device surpassed 18 mA cm−2. Although the low
4.3. Summary and Challenges
mobilities led to a moderate FF, the overall efficiency reached
8.47%. In another report, Fan et al. compared PDI dimers Rylene diimides and their derivatives are the earliest electron
connected directly to each other via different positions.168 DFT acceptors applied in organic solar cells and are still being widely
calculations showed that the dimer connected at the α-position studied by many research groups. Among various rylene
(2,2′-bi(PDI), Figure 18, molecule 4.43) had a higher dihedral diimides, PDI is the workhorse material to construct small
angle between the two PDI units than that of the dimer molecular NFAs owing to their capability of being function-
connected at the bay-position (1,1′-bi(PDI)). The absorption alized into derivatives that could be solution-processed into
spectrum of 1,1′-bi(PDI) was broader than that of 2,2′- nanoscale structures while retaining a relative high electron
bi(PDI), indicating a stronger inter-PDI electronic coupling. In mobility and a strong absorption coefficient. Most PDI

Figure 16. DFT molecular geometries of PDI helical ribbons for the PDI dimer 2 (a), trimer 3 (b, c), and tetramer 4 (d, e, f) from ref 159. UV−vis
absorption (g) and photoluminescence spectra (h) for 2−4 along with a PDI monomer in solution. Adapted with permission from ref 159.
Copyright (2014) American Chemical Society.

3463 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 17. (a) X-ray molecular structures of TPH 4b and TPH-Se 6a (top view and side view). (The alkyl chains and hydrogen atoms are omitted
for clarity.) (b) The slipped 3D stacking mode of TPH 4b (top) and TPH-Se 6a (down). (c) Se···O (3.0 Å) interactions between different PBI
subunits of TPH-Se 6a. Adapted with permission from ref 162. Copyright (2016) American Chemical Society.

acceptors have bandgaps over 1.9 eV, and therefore low exciton dissociation. Meanwhile, most PDI-based organic solar
bandgap donor polymers were generally used to complement cells (OSCs) need to be processed with solvent additives.
the absorption. Electrically, electron mobilities of recent high- Several reports showed that the device performance was
performance PDI acceptors are typically on the order of 10−5 to sensitive to the amount of additives, which is an issue to be
∼10−3 cm2 V−1 s−1, which, despite further enhancement still addressed since solvent additives may affect the large-scale
being required, are able to enable a high FF in solar cell devices production and device longevity.
in general. Another advantage of PDI-based acceptors is the
facile funtionalization of PDI. Substitution at different positions 5. A-D-A TYPE ACCEPTORS
has been realized to construct different acceptors. Nevertheless, In addition to the development of rylene-based NFAs, the
an important consideration of PDI acceptors is to control the successful strategy of using a conjugated “push-pull” structure
coplanarity, which is important to balance electron mobility and in designing semiconducting polymers has also been applied to
3464 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

easily extended to the near-infrared (NIR) that could


substantially improve photocurrent generation and in the
meantime allows the design of semitransparent devices. To
tune the solubility and film morphology, alkyl or aromatic side
chains are typically introduced on the central fused-rings.
Overall, the synthetic flexibility, easily tuned electronic/optical
property, and rigid backbone make the A-D-A type small
molecules promising acceptor materials for nonfullerene OPV
applications. Chemical structures and properties of representa-
tive materials are summarized in Figure 21 and Table 2.
5.1. Early Reports of A-D-A Type Acceptors
One of the first NFAs of this type, named FEHIDT (Figure 21,
Figure 18. Chemical structures of selected α-substituted PDIs. molecule 5.1), was reported in 2013 by Winzenberg et al.,
where the “A” and “D” units are indan-1,3-dione and fluorene,
respectively, and a thiophene was used as the spacer.177 The
spacer thiophene and the flanking indan-1,3-dione are coplanar,
but they have a dihedral angle of ∼20° with the central fluorene
unit. The acceptor molecule has a large bandgap similar to that
of PDI, with a strong absorption in the range of 400−600 nm.
A PCE of 2.43% was obtained by using FEHIDT in
combination with P3HT (Figure 6) as the donor polymer. In
a later study by Kim et al., the indan-1,3-dione (“A” unit) of
FEHIDT was replaced by 3-ethylrhodanine, and a new acceptor
named Flu-RH (Figure 21, molecule 5.2) was obtained.178 In
addition, the alkyl chains on the central fluorene unit were n-
octyl instead of 2-ethylhexyl. Overall, Flu-RH showed a similar
optical bandgap and absorption range to FEHIDT, but a higher
LUMO level offered it a high VOC of 1.03 V with P3HT, which
led to a best efficiency of 3.08% despite a lower FF (0.52).
Another early work on push−pull NFA was represented by
Figure 19. Chemical structures of the NDI and TDI-based NFAs.
the report of FBR (Figure 21, molecule 5.3) by Holliday et al.
constructing NFAs. The electron-rich and electron-deficient FBR was obtained by replacing the thiophene spacers in Flu-
moieties combined can extend conjugation and reduce RH by benzothiadiazole units.179 The bandgap and absorption
bandgap. An A-(π)-D-(π)-A backbone (hereafter denoted as range of FBR was similar to those of Flu-RH. The spacer,
A-D-A) is typically employed when designing novel NFAs, benzothiadiazole, and flanking rhodanine were coplanar, while a
where “A” and “D” represent the electron-withdrawing and dihedral angle of ∼35° between them and the central fluorene
electron-donating moieties, respectively, which are sometimes unit was exhibited. The nonplanar structure could provide the
linked by a π conjugated spacer unit or a second “A” or “D” molecule with nonanisotropic electron transport and reduced
moiety, to further extend the conjugation (Figure 20). Despite tendency to overaggregate, which contributed to an excellent
PCE of 4.1% with P3HT as the donor polymer.
A major breakthrough in the design of A-D-A type acceptor
is the introduction of an indacenodithiophene (IDT) building
block as the central “D” unit. The first example, a NFA named
DC-IDT2T (Figure 21, molecule 5.4), was reported by Bai et
al., where the “A” unit was 1,1-dicyanomethylene-3-indanone
(DCI).180 A small optical bandgap with a strong absorption
extended to the NIR region was demonstrated. The overall
Figure 20. Cartoon representing a chemical structural template for device performance with a low bandgap donor polymer
designing the A-D-A type push−pull NFAs. PBDTTT-T-C was 3.93%.
Two milestone A-D-A type acceptors were soon reported by
the fact that other types of NFAs have also been developed, modifying DC-IDT2T. The first one, IEIC (Figure 21,
e.g., a D-A-D small molecular acceptor,175,176 most high molecule 5.5), reported by Lin et al., was obtained by adding
performance NFAs are synthesized using this structural a 2-ethylhexyl alkyl chain on each spacer thiophene.181 Properly
template. Besides the easily tuned absorption for spectral matched energy levels and strong absorption ensured a high
breadth and the rigid backbone for reduced reorganization VOC of 0.97 V and a good JSC of 13.55 mA cm−2, which led to
energy that benefits charge conduction, the A-D-A molecules an excellent PCE of 6.31% when IEIC was blended with the
have the electron-rich and electron-deficient moieties located at low bandgap polymer PTB7-Th. Later, Lin et al. reported
different parts of the molecule, allowing possible contact another milestone molecule named ITIC (Figure 21, molecule
between the electron-deficient part of the donor material and 5.6), where the IDT core was fused with the thiophene spacer
the LUMO of the NFA to facilitate charge transfer. Owing to in DC-IDT2T.182 This modification slightly downshifted the
their strong intramolecular charge transfer, the A-D-A back- energy levels and enlarged the bandgap. In combination with
bone typically exhibits strong and broad absorption that can be PTB7-Th, the BHJ solar cell (PTB7-Th:ITIC) exhibited PCE
3465 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 21. Chemical structures of A-D-A type NFAs.

3466 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Table 2. Summary of Representative A-D-A Type Acceptors Materials
mobility (cm2 V‑1 s‑1)
internal LUMO HOMO Eg,opt D:A processing PCE VOC JSC FF device
acceptor ref (eV) (eV) (eV) donor ratio solvent (%) (V) (mA cm‑2) (%) E (neat) E (blend) H (blend) structure ref
FEHIDT 5.1 −3.95d −5.95c 2.00 P3HT 1.2:1 DCB 2.43 0.95 3.82 67 NR NR NR conventional 177
Flu-RH 5.2 −3.53a −5.58a 2.10 P3HT 1:1.5 DCB 3.08 1.03 5.70 52 NR NR NR conventional 178
Chemical Reviews

FBR 5.3 −3.57b −5.70b 2.14 P3HT 1:1 CF+DCB 4.11 0.82 7.95 63 NR 2.6 × 10−5 NR inverted 179
(4:1)
DC-IDT2T 5.4 −3.85a −5.43a 1.55 PBDTTT-C-T 1.2:1 DCB+15%CF 3.93 0.90 8.33 52.3 3.3 × 10−4 1.5 × 10−3 2.0 × 10−3 conventional 180
IEIC 5.5 −3.82a −5.42a 1.57 PTB7-Th 1:1.5 NR 6.31 0.97 13.55 48 2.1 × 10−4 1.0 × 10−4 4.5 × 10−4 conventional 181
ITIC 5.6 −3.83a −5.48a 1.59 PTB7-Th 1:1.3 NR 6.80 0.81 14.21 59.1 3.0 × 10−4 1.1 × 10−4 4.3 × 10−5 conventional 182
IDTT-2BM 5.7 −3.8a −5.5a 1.54 PBDTTT-C-T 1.5:1 DCB+CF 4.81 0.851 9.87 57.2 1.0 × 10−5 1.3 × 10−5 4.1 × 10−4 conventional 183
(6:4)+3%
DIO
IDT-2BR 5.8 −3.69a −5.52a 1.68 P3HT 1:0.6 DCB+3%CN 5.12 0.84 8.91 68.1 3.4 × 10−4 2.6 × 10−4 2 × 10−4 conventional 184
O-IDTBR 5.9 −3.88a −5.51d 1.63 P3HT 1:1 CB 6.30 0.72 13.9 60 NR 3−6 × 10−6 NR inverted 25
IDTIDT- 5.10 −3.82a −5.42a 1.53 PTB7-Th 1:1.5 DCB 6.48 0.94 14.49 47.5 NR 4.54 × 10−5 NR conventional 185
IC
IDTIDSe- 5.11 −3.81a −5.41a 1.52 J51 1:1 CF 8.02 0.91 15.16 58.0 1.27 × 10−5 7.87 × 10−5 7.21 × 10−5 conventional 186
IC
IDSe-T-IC 5.12 −3.79a −5.45a 1.52 J51 1:1 CF 8.58 0.91 15.20 62.0 NR 7.72 × 10−5 8.25 × 10−5 conventional 187
IC-C6IDT- 5.13 −3.91a −5.69a 1.62 PDBT-T1 1:1 CF 8.71 0.89 15.05 65 1.1 × 10−3 2.9 × 10−4 5.1 × 10−5 inverted 188
IC
IC-1IDT- 5.14 −3.83a −5.91a 1.70 PDBT-T1 1:1 CF 7.39 0.92 13.39 60 4.5 × 10−4 1.4 × 10−4 3.4 × 10−5 inverted 189

3467
IC
ITIC-Th 5.15 −3.93a −5.66a 1.60 PDBT-T1 1:1 CF+1%CN 9.6 0.88 16.24 67.1 6.1 × 10−4 4.2 × 10−4 3.0 × 10−4 inverted 190
m-ITIC 5.16 −3.82a −5.52a 1.58 J61 1:1 CF 11.77 0.912 18.31 70.55 2.45 × 10−4 1.30 × 10−4 1.54 × 10−4 conventional 191
IDT-BOC6 5.17 −3.78a −5.51a 1.63 PBDB-T 1:1 DCB+1%DIO 9.60 1.01 17.52 54 NR 4.99 × 10−4 5.31 × 10−4 inverted 192
ATT-1 5.18 −3.63b −5.50b 1.54 PTB7-Th 1:1.5 CB+1%DIO 10.07 0.87 16.48 70 NR 2.40 × 10−4 5.13 × 10−4 conventional 190
ATT-2 5.19 −3.90b −5.50b 1.32 PTB7-Th 1:1.8 CB+2%CN 9.58 0.73 20.75 63 NR 3.69 × 10−4 5.10 × 10−4 inverted 191
IEICO 5.20 −3.95a −5.32a 1.34 PBDTTT-E-T 1:1 CB+2%DIO 8.4 0.82 17.7 58 1.40 × 10−4 4.6 × 10−4 1.5 × 10−3 conventional 193
IEICO-4F 5.21 −4.19a −5.44a 1.24 PTB7-Th 1:1.5 CB 10.0 0.739 22.8 59.4 1.14 × 10−4 NR NR conventional 194
IT-M 5.22 −3.98a −5.58d 1.60 PBDB-T 1:1 CB+1%DIO 12.05 0.94 17.40 73.5 NR 1.10 × 10−4 3.33 × 10−4 inverted 195
INIC3 5.23 −4.02a −5.52a 1.48 FTAZ 1:1.5 CF+0.25% 11.5 0.852 19.68 68.5 1.7 × 10−4 1.4 × 10−4 2.0 × 10−4 inverted 196
DIO
ITIC-Th1 5.24 −4.01a −5.74a 1.55 FTAZ 1:1.5 CF+0.25% 12.1 0.849 19.33 73.73 NR 7.6 × 10−3 2.7 × 10−2 inverted 33
DIO
IT-4F 5.25 −4.14c −5.66c 1.52h PBDB-T-SF 1:1 CB+0.5%DIO 13.0 0.88 20.50 71.9 5.05 × 10−4 4.32 × 10−4 3.25 × 10−4 inverted 19
ITCC 5.26 −3.76a −5.47a 1.67 PBDB-T 1:1 CB+1%DIO 11.4 1.01 15.9 71 9.26 × 10−4 6.74 × 10−4 NR conventional 197
ITCPTC 5.27 −3.96b −5.62b 1.58 PBT1-EH 1:1 CF 11.8 0.95 16.5 75.1 3.2 × 10−3 2.69 × 10−3 1.71 × 10−3 conventional 198
DTBTF 5.28 −3.62b −5.68b 2.03 DR3TSBDT 1:0.5 CF 3.84 1.15 7.42 45 NR 4.13 × 10−5 1.14 × 10−4 conventional 199
DICTF 5.29 −3.79b −5.67b 1.82 PTB7-Th 1:1.4 CF 7.93 0.85 16.33 55 NR 1.93 × 10−4 3.82 × 10−4 conventional 200
FDICTF 5.30 −3.71b −5.43b 1.63 PBDB-T 1:1.2 CB+0.3%DIO 10.06 0.94 15.81 66 NR 2.40 × 10−5 3.37 × 10−5 conventional 201
CBM 5.31 −4.13b −6.05b 2.02 PTB7-Th 3:7 CB+2%DIO 5.3 0.88 10.6 53 NR 1.9 × 10−6 1.0 × 10−4 inverted 202
IDFBR 5.32 −3.70a −5.75a 2.0 P3HT 1:1 CB 4.5 0.89 7.4 68 NR NR NR inverted 26
SFBRCN 5.33 −3.86a −5.93a 2.05 PTB7-Th 1:1.2 CB+0.75% 10.12 0.90 17.25 65.2 NR 2.2 × 10−4 1.5 × 10−4 inverted 203
DIO
Review

DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
NFBDT 5.34 −3.83b −5.40b 1.56 PBDB-T 1:0.8 CF 10.42 0.868 17.85 67.2 NR 1.38 × 10−4 3.68 × 10−4 conventional 204
Chemical Reviews Review

up to 6.8%, comparable to PC61BM-based OSCs. A remarkable

Measured by CV using NFA films. bMeasured by CV in solutions. cMeasured by other methods. dCalculated using measured HOMO/LUMO and the optical bandgap. eDetailed method was not
205
ref
improvement was recently reported by using a large bandgap
polymer PBDTBDD for ITIC.29 Benefiting from suitable

conventional
structure energy levels, complementary absorption, and high mobilities, a
device

high VOC of 0.899 V, an excellent JSC of 16.81 mA cm−2, and a


FF of 74.2% were obtained, respectively, which resulted in an
outstanding efficiency of 11.21%.
5.2. A-D-A Type Acceptors with a Core Unit of IDT or Its
1.07 × 10−3
H (blend)

Derivatives
5.2.1. Design of the Core Units. Following the success of
IEIC and ITIC, new acceptors based on IDT and its derivatives
mobility (cm2 V‑1 s‑1)

have been developed. On the basis of ITIC, an NFA with


1.86 × 10−3
E (blend)

indacenodithieno[3,2-b]thiophene (IDTT) as the core unit,


named IDTT-2BM (Figure 21, molecule 5.7) was reported by
Bai et al.183 Compared to ITIC, the flanking units were changed
to 2-(benzo[c][1,2,5]-thiadiazol-4-ylmethylene)-malononitrile
(BM), leading to a reduced bandgap and a rigid, coplanar
2.17 × 10−3

backbone along the entire molecule. However, the electron


E (neat)

mobility of IDTT-2BM appeared to be lower than that of ITIC,


which caused a largely imbalanced hole/electron mobility and
reported. fMeasured using OFET devices. hDetermined from the film absorption onset by the authors of this review; NR Not reported.

could partially explain the modest FF (∼57%) and PCE


(4.81%) with PBDTTT-C-T as the donor polymer.
56.2
(%)
FF

The IDT core has also been introduced to FBR, and the
resulting NFA was named IDT-2BR (Figure 21, molecule
(mA cm‑2)

5.8).184 It adopted a planar backbone, while the hexylphenyl


11.23
JSC

side groups displayed a dihedral angle of 115° (Figure 22),


which facilities charge transport while preventing large phase
separation in the blend film. With P3HT as the donor polymer,
0.95
VOC
(V)

an excellent performance (5.12%) was achieved.


PCE
(%)
6.0
processing

CF+1%CN
solvent
ratio
D:A

1:1
donor
PTB7-Th

Figure 22. (a, b) Optimized molecular geometries of IDT-2BR. (c, d)


Molecular orbitals of IDT-2BR. Reproduced from ref 184 with
Eg,opt
(eV)
1.59

permission. Copyright 2015 Royal Society of Chemistry.

Later, Holliday et al.25 reported another effort to introduce


HOMO

−5.50a
(eV)

IDT core to FBR with n-octyl as the side chains on the IDT
core instead of the n-hexylphenyl used in IDT-2BR. This
acceptor, O-IDTBR (Figure 21, molecule 5.9), exhibited a red-
LUMO

−3.87a

shifted absorption and a more ordered nanostructure than both


(eV)

FBR and IDT-2BR due to an enhanced intermolecular packing,


especially after thermal annealing, which emphasized the
Table 2. continued

importance of side chains. The downshifted LUMO level


internal

5.35
ref

resulted in a smaller VOC of 0.72 V with P3HT, but the


broadened absorption boosted the JSC to 13.9 mA cm−2. An
overall PCE of 6.30% was achieved for P3HT:O-IDTBR, which
DTCC-IC
acceptor

was much higher than P3HT:PCBM based solar cells.


Li et al. showed an acceptor, IDTIDT-IC (Figure 21,
molecule 5.10), with a fused IDTIDT as the central donor unit
a

3468 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

and DCI as the end-capping.185 The 10-heterocyclic fused ring general, the “D” moiety consists of multiple conjugated
backbone of IDTIDT-IC extended its absorption into the NIR aromatic rings and is typically larger in size than the “A”
(λedg ∼ 810 nm) region, which gave rise to a high JSC of 14.49 moiety. Therefore, in most high-performance A-D-A type
mA cm−2 in combination of PTB7-Th. In spite of the high JSC, NFAs, the side chains are substituted on the “D” moiety, to
the BHJ device was able to achieve a high VOC of 0.94 V with a prevent excessive steric hindrance between the “A” units of the
low energy loss (∼0.59 eV, calculated by subtracting eVOC from acceptor and the electron deficient fragment of the polymer.
the optical bandgap of IDTIDT-IC) and a high IQE (>80%) Drawing on the accomplishment of the IDT-based backbones,
throughout nearly the entire absorption range. The electron engineering the side chains could further improve the device
mobility for IDTIDT-IC in the blend film was lower than that performance and/or offer wider applicability to more donor
of IEIC or ITIC, which was likely the source of a relatively low materials. For instance, Lin et al. replaced the alkylphenyl
FF (47.5%). Overall, the best efficiency of PTB7-Th:IDTIDT- chains of ITIC with alkylthienyl chains (ITIC-Th, Figure 21,
IC reached 6.48%. On the basis of this work, Li et al. replaced molecule 5.15).190 Compared to phenyl, the thienyl side chains
the two sulfur atoms at the end of the IDTIDT core with could enhance the nanoscale molecular order induced by the
selenium (Figure 21, molecule 5.11).186 The large Se atoms sulfur−sulfur intermolecular interaction, which was beneficial
increased the ground state quinoid resonance character, which for electron transport. As a result, a PCE as high as 9.6% was
improved the electron mobility and lowered the bandgap. achieved in an optimized PDBT-T1:ITIC-Th BHJ, with an
When a medium bandgap donor polymer (J51; Figure 6) with excellent JSC of 16.24 mA cm−2 and a high FF of 67.1%.
an intense absorption in the range of 400−600 nm was used to In addition to using structurally different side chains, the
complement the absorption, and a high efficiency of 8.02% was position of the side chain can also dramatically vary material
achieved. properties and device performance. For example, Yang et al.
The strategy of replacing sulfur with selenium has also been relocated the alkyl chains on the phenyl rings of ITIC, forming
applied to modifying IEIC, which produced a planar acceptor an isomer named m-ITIC (Figure 21, molecule 5.16).191 The
IDSe-T-IC (Figure 21, molecule 5.12) with a slightly lower isomerization caused little effect on the electronic structure of
bandgap than IEIC.187 In combination with J51, a high the acceptor but enhanced its intermolecular self-assembly. In
efficiency up to 8.58% was demonstrated. the two-dimensional grazing-incidence wide-angle X-ray
One marked advantage of the A-D-A type NFA is their facile scattering (2D-GIWAXS) patterns (Figure 23), sharper and
synthesis. The “A” and “D” units could be distinctly prepared
before being connected together. As an example, Lin et al.
synthesized a facile planar fused ring molecule named IC-
C6IDT-IC (Figure 21, molecule 5.13), where an IDT core was
directly flanked with DCI units, and the side chains were simple
alkyl instead of alkylphenyl groups.188 The BHJ of PDBT-
T1:IC-C6IDT-IC, where the polymer and NFA showed
complementary absorption and compatible electronic property,
provided a PCE as high as 8.71%. To understand the
structure−property relationship, the authors synthesized a
series of structurally similar derivatives with 1−3 IDT as
cores and studied the morphological evolution of the series.189
An acceptor IC-1IDT-IC (Figure 21, molecule 5.14) was first
obtained by removing the alkyl thiophene spacer in IEIC, and
two analogues with 2 or 3 IDT in the cores were developed. It
was shown that, with increasing number of IDT in the core, the
bandgap as well as the tendency to crystallize of the acceptor
decreased, which was consistent with the trend for the electron
mobility. Two donor polymers, PTB7-Th and PDBT-T1, were
used to construct 6 different BHJs, among which, the best
performance (7.39% efficiency) was obtained with PDBT-
T1:IC-1IDT-IC due to a fine balance among energy level,
carrier mobility and film morphology. The aforementioned IC-
C6IDT-IC was achieved by replacing the alkylphenyl side Figure 23. (a) Line cuts of the 2D-GIWAXS patterns of neat m-ITIC
chains with simple alkyl chains on the IDT core unit. and ITIC films. 2D-GIWAXS patterns of (b) m-ITIC film and (c)
ITIC film. Adapted with permission from ref 191. Copyright (2016)
Compared to IC-nIDT-IC (n = 1−3), IC-C6IDT-IC showed
American Chemical Society.
a bathochromic shift in its absorption, ordered crystal packing,
and an improved electron mobility of 1.1 × 10−3 cm2 V−1 s−1 in
neat films and 2.9 × 10−4 cm2 V−1 s−1 in the blend with PDBT- more intense π−π stacking peaks were observed in the out-of-
T1. These improvements resulted in a high FF and JSC, and an plane direction for m-ITIC in both pure and blend films with a
excellent overall performance of 9.20%. It is also worth noting medium bandgap polymer (J61; Figure 6). The enhanced
that all of these devices were fabricated from as-cast films intermolecular interaction resulted in an increased electron
without solvent additives in the solution or solvent/thermal mobility, which in turn gave rise to an outstanding OPV
annealing after film formation. performance with a PCE of 11.77%, a VOC of 0.912 V, a JSC of
5.2.2. Side Chain Effects. Akin to its role in the design of 18.31 mA cm−2, and a FF of 70.55%. In addition, the authors
polymeric donor materials, the side chains of A-D-A type NFA demonstrated that the PCE of the device was less sensitive to
are critical in determining the solubility and morphology. In thickness of the active layer, e.g., a PCE of >8% was maintained
3469 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

even at a thickness of 360 nm, which further illustrated the g − eVOC) of ∼0.5 eV, indicating a high potential of the
(Eopt
strong impact of a seemingly minor structural modification. PCE.
5.2.3. Effect of the Spacer Unit. Besides creating large The increasingly smaller energy loss observed in these recent
fused ring systems, another strategy to rigidify the backbone of NFA-based systems triggered research to further reduce the
the acceptor was to use “conformational locks”, which utilizes energetic offsets between the donor and the NFA. Due to the
noncovalent interaction such as intramolecular hydrogen high PCE and wide applicability of the polymer PBDB-T in the
bonding to lock the coplanar conformation. This strategy has ITIC family NFAs, in order to further reduce energy loss, one
previously been employed to design polymers with high charge facile idea would be to slightly increase the energy levels of the
carrier mobilities. Since the neighboring aromatic rings are NFA on the basis of ITIC. By introducing one or two methyl
connected by a single bond, the synthesis of such conforma- groups as the substituents to the end-capping groups
tional locked NFAs should be more facile than ring fusion. The (dicycanovinylindan-1-one) of ITIC, Li et al. were able to
first attempt was reported by Liu et al.,192 who used obtain two novel methyl-modified NFAs named IT-M (Figure
dihexyloxybenzene as the spacers to replace the thiophene in 21, molecule 5.22) and IT-DM, respectively.195 Compared to
IEIC. The new NFA, IDT-BOC6 (Figure 21, molecule 5.17), ITIC, incorporation of the methyl groups was found to slightly
was compared with an analogue acceptor without oxygen (IDT- upshift both the HOMO and LUMO levels, which enhanced
BC6). It was shown that IDT-BOC6 possessed a more planar the VOC with the same donor polymer and improved the
structure than IDT-BC6, which offered it a smaller bandgap, a nanoscale morphology in the meantime. Taking all of these
higher electron mobility, and a higher fluorescence quantum effects into account, the best device performance was achieved
yield. When PBDB-T (Figure 6) was used as the donor using PBDB-T:IT-M, which showed a PCEmax of 12.05%, with a
polymer, a high PCE of 9.6% was achieved. VOC of 0.94 V, a JSC of 17.40 mA cm−2, and a FF of 73.5%.
Liu et al. designed two acceptors with thieno-[3,4-b]- Inspired by the success of IDTT as the core and fluorinated
thiophene as the spacer to replace the thiophene in IEIC, DCI as the “A” moiety, Dai et al. constructed a series NFAs
namely ATT-1 (Figure 21, molecule 5.18)190 and ATT-2 using 6,6,12,12-tetrakis(4-hexylphenyl)-indacenobis(dithieno-
(Figure 21, molecule 5.19).191 ATT-2 had the same electron [3,2-b;2,3-d]thiophene) (IBDT) as the core, which had a
accepting moiety (DCI) as IEIC, while ATT-1 featured a new further extended, rigid and coplanar structure, as well as a
“A” moiety, 2-(1,1-dicyanomethylene)rhodanine. In addition, stronger electron donating ability than IDT and IDTT. DCI or
the direction of the thieno-[3,4-b]thiophene spacer was fluorinated DCI were used as the end-capping groups.196 The
different from each other for ATT-1 and ATT-2. Due to a resulting acceptor INIC had a larger rigid plane and stronger
stronger electron-withdrawing ability of DCI, ATT-2 exhibited electron-donating ability, which enhanced the optical absorp-
a smaller optical bandgap, with absorption edge extended to tion and electron transport. Fluorination on the DCI end
∼900 nm. Both acceptors were combined with PTB7-Th and groups further enhanced charge transport ability, and lowered
achieved PCEs of ∼10% in their BHJ solar cells. In addition, the LUMO levels without significantly affecting the HOMO
benefiting from efficient photon harvesting in the NIR region, levels. With two fluorine atoms on each DCI unit, the acceptor
semitransparent PTB7-Th:ATT-2 devices were fabricated and a INIC3 (Figure 21, molecule 5.23) showed the strongest
very promising PCE of 7.74% was achieved, while maintaining tendency to crystallize, which led to the tightest polymer
an average transmittance of 37% in the visible with good color packing during the cocrystallization of FTAZ (Figure 6) and
rendering property. INIC3. More importantly, the extended coherence length did
The attempt to further reduce the optical bandgap of these not result in an excessively large phase separation and a
IDT-based A-D-A acceptors and harvesting more photons in reasonably small domain size in the FTAZ:INIC3 blend was
the NIR region was reported by replacing alkyl side chains with obtained. All of these factors contributed to excellent JSC and
alkoxy chains on the spacer units of IEIC.193 The new acceptor FF of 19.68 mA cm−2 and 68.5%, respectively. Although the
IEICO (Figure 21, molecule 5.20) has an absorption edge of lower LUMO level led to a relatively smaller VOC of 0.852 V
∼925 nm due to an upshifted HOMO level compared with among the acceptors, the overall efficiency of FTAZ:INIC3
IEIC. Consequently, when combined with a low bandgap reached 11.5%.
donor polymer PBDTTT-E-T, IEICO achieved a much higher Furthermore, fluorination of the DCI unit was also applied to
JSC (17.7 mA cm−2) and PCE (8.4%) than IEIC (JSC = 11.7 mA ITIC-Th and ITIC.33 Similar to the previous reports, a
cm−2 and PCE = 4.9%). Using the optical bandgap of IEICO, monofluorinated acceptor, namely ITIC-Th1 (Figure 21,
g − eVOC) was as low as 0.52
the energy loss of this system (Eopt molecule 5.24), exhibited lower energy levels, slightly red-
eV. shifted absorption, improved intermolecular interaction, and
5.2.4. Design of the “A” Unit. In addition to the central better electron mobility than ITIC-Th. When FTAZ was used
“D” moiety, the side chains, and the spacer unit, the property of as the donor polymer, the blend FTAZ:ITIC-Th1 showed a
the end-capping “A” moieties is essential for determining the slightly higher crystallinity and a smaller domain size than
overall electron affinity and bandgap. FTAZ:ITIC-Th. As a result, a high JSC of 19.33 mA cm−2 and a
To be better energetically compatible with narrow bandgap FF of 73.73% of FTAZ:ITIC-Th1 led to an outstanding
donor polymers, the energy levels of IEICO were downshifted efficiency of 12.1%, which was much higher than that base on
by introducing fluorine atoms on the end-capping DCI units, ITIC-Th. In a recent report, difluorination of each DCI unit of
which also narrowed its optical bandgap to 1.24 eV as a result ITIC led to a new acceptor IT-4F (Figure 24).19 To match the
of the enhanced intramolecular charge transfer effect.194 When downshifted energy levels of IT-4F, a derivative of PBDB-T was
the new acceptor, IEICO-4F (Figure 21, molecule 5.21), was developed by fluorinating the BDT building block. In
combined with the low bandgap polymer (PTB7-Th), an comparison to fullerene-based organic cells where the newly
outstanding JSC of 22.8 mA cm−2 was achieved, resulting in an designed polymers must meet the energy level requirement set
overall efficiency of 10.0%. Similar to IEICO, the device based by PCBM, the capability of synergistically adjusting the energy
on PTB7-Th:IEICO-4F also demonstrated a small energy loss levels for both donor and acceptor clearly shows the design
3470 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

al. used the weak electron-donating unit fluorene as the


building block and introduced thiobarbituric acid as the end-
capping “A” unit to replace the rhodanine in Flu-RH.199 The
acceptor DTBTF (Figure 21, molecule 5.28) showed a large
optical bandgap with strong absorption in the range of 400−
600 nm. When blended with a medium bandgap molecular
donor DR3TSBDT (Figure 6), the BHJ device demonstrated a
high VOC of 1.15 V benefiting from the high LUMO level of the
acceptor. However, both the JSC and FF of the optimized device
were relatively low, which could be attributed to the relatively
low electron mobility and large domains observed in the
transmission electron microscopy (TEM) images, leading to an
efficiency of 3.84%. In a later report by Li et al., the “A” unit
was further replaced by DCI, and a new acceptor DICTF
(Figure 21, molecule 5.30) was obtained. The DCI end-group
lowered the LUMO level without significantly altering the
HOMO level.200 With PTB7-Th as the donor material, the
blend PTB7-Th:DICTF showed a higher JSC and FF than those
of DR3TSBDT:DTBTF. Overall, the PCE was improved to
7.93%. On the basis of these NFAs, Qiu et al. synthesized a new
acceptor FDICTF by rigidifying the core unit and the spacer in
DICTF, forming a ladder type large core unit.201 FDICTF
possessed two additional alkyl chains on each “lock” sitting
Figure 24. (a) Chemical structures of ITIC and PBDB-T, along with between the original core unit and spacer of DICTF. This
their fluorinated derivatives IT-4F and PBDB-T-SF, respectively. (b) structure evolution flattened the molecule, slightly up-shifted
Absorption spectra of neat polymer and NFA films. (c) Energy the LUMO level, and significantly elevated the HOMO level,
diagrams of the materials. Adapted with permission from ref 19. together causing the absorption to red-shift by ∼65 nm. To
Copyright (2017) American Chemical Society. complement the red-shifted absorption and modified energy
levels, the donor polymer PBDB-T was chosen to form a BHJ.
Compared to PBDB-T:DICTF, a significantly improved
flexibility of materials for nonfullerene organic solar cells. efficiency up to 10.06% was achieved by PBDB-T:FDICTF.
Compared to ITIC, the red-shifted absorption edge of IT-4F Using 2-(benzo[c][1,2,5]thiadiazol-4-ylmethylene)-
allowed photon harvesting in the NIR, which resulted in a JSC of malononitrile (BM) as the end-capping groups, Wang et al.
20.50 mA cm−2 for the PBDB-T-SF:IT-4F BHJ and a designed A-D-A NFAs with three core units including fluorine,
benchmark PCE of 13.0%. In addition, the device exhibited a carbazole, and cyclopenta-[2,1-b:3,4-b′]dithiophene.202 The
remarkable PCE of >12% at an active layer thickness of 200 three acceptors were named FBM, CBM (Figure 21, molecule
nm. 5.31), and CDTBM, respectively. They exhibited comparable
Besides fluorination, Yao et al. reported another strategy to solar cell performance when PTB7-Th was used as the donor
modify the “A” unit by tailoring the aromatic ring of the DCI polymer, with CBM showing the best PCE of 5.3%.
building block.197 The new acceptor ITCC (Figure 21, Baran et al. replaced the IDT core in O-IDTBR with
molecule 5.26) had a fused thiophene ring, instead of a phenyl indacenodibenzene, which enlarged the bandgap.26 Although
ring (ITIC), at each end of the molecule. The stronger electron the new acceptor IDFBR (Figure 21, molecule 5.32) showed
donating nature of the thiophene than phenyl provided ITCC only a 4.5% PCE with P3HT, it showed strong and
with a higher lying LUMO level than ITIC. Morphological complementary absorption to O-IDTBR. The ternary blends
characterization revealed a denser π−π stacking for ITCC than using IDFBR and IDTBR as the acceptors and P3HT or PTB7-
ITIC, which led to a higher electron mobility. With PBDB-T as Th as the donor demonstrated largely improved photovoltaic
the donor polymer, the blend of PBDB-T:ITCC exhibited a performance (see section 9.3.1 for details) and device stability
higher domain purity than PBDB-T:ITIC, which was (see section 10.1 for details).
responsible for the suppressed bimolecular recombination and Recently, Zhang et al. introduced spirobifluorene as the core
enhanced FF of the former blend. Overall, the higher FF and unit to FBR. Flanked with BT, the new acceptor, namely
VOC led to a PCE of 11.4% for the PBDB-T:ITCC BHJ device, SFBRCN (Figure 21, molecule 5.33), contained a strong
despite a slightly reduced JSC due to the narrow absorption electron withdrawing unit, 2-(1,1-dicyanomethylene)rhodanine,
range relative to ITIC. In another report by Xie et al., the as the end-capping “A” moiety.203 The acceptor maintained a
thiophene ring on the “A” unit was fused with the cyclopentane similarly large bandgap (2.03 eV) to FBR, while the strong
ring at its 3- and 4-positions, which altered the double bond electron deficient nature of BT and the end-groups afforded
arrangement and thus the electronic structure of the whole SFBRCN a deep LUMO. Furthermore, the authors showed
molecule.198 The molecule ITCPTC (Figure 21, molecule that the 3D structure of the BT flanked spirobifluorene core
5.27) exhibited a slightly smaller bandgap and higher mobility prevented the molecule from overaggregation. In combination
than ITIC. In combination with a large bandgap donor polymer with the low bandgap polymer PTB7-Th, which provided a
PBT1-EH (Figure 6), the overall PCE reached 11.8%.
complementary absorption to SFBRCN, the BHJ device
5.3. A-D-A Type Acceptors with Other Core Units showed a high PCE of 10.12%.
In addition to IDT and IDTT, A-D-A type molecules with Replacing the benzene unit in the IDT core by a BDT was
other core units were also developed as promising NFAs. Ni et reported by Kan et al.204 The acceptor NFBDT (Figure 21,
3471 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 25. Chemical structures of DPP-based NFAs.

molecule 5.34) possessed the same conjugation length as NFAs with other cores and shapes had never halted. For
IDTT. Consequently, they showed similar bandgaps. When example, the diketopyrrolopyrrole (DPP) based NFAs continue
PBDB-T was used as the donor polymer, a high PCE of 10.42% to emerge, and the subphthalocyanine (SubPc) type NFAs have
was achieved by performing solvent vapor annealing to the demonstrated planar heterojunction OPV devices with PCEs
active layer prior to electrode deposition. over 8%.65 We note that phthalocyanine (Pc) dyes are widely
In another report by Cao et al., the IDT core in IC-1IDT-IC employed to design NFAs due to their high extinction
was replaced by DTCC, which had seven fused rings and five coefficient and relatively long exciton diffusion length. Most
side chains.205 The acceptor DTCC-IC (Figure 21, molecule of the Pc-based NFAs were deposited using vacuum deposition
5.35) had four alkoxyphenyl and one linear octyl side chains. techniques, and the OPV devices were based on planar
DTCC-IC had a smaller bandgap than IC-1IDT-IC and heterojunction structures. Since this review focuses on solution-
showed a high SCLC mobility (∼10−3 cm V−1 s−1), which processed BHJ solar cells, the details of the Pc-based
gave rise to a 6.0% PCE in its BHJ solar cell with PTB7-Th as nonfullerene planar heterojunction solar cells are not discussed.
the donor. Reviews on this topic are available.39,43,206−208 Nonetheless,
5.4. Summary other NFAs with different structures can inspire the design and
application of new moieties in the state-of-the-art molecules.
Although A-D-A type NFAs were developed much later than This section surveys these NFA classes and summarize their
PDIs, tens of new structures with excellent performance have performance in BHJ OPVs.
been reported to date. The best performance of this type has 6.1. DPP-Based NFAs
surpassed those of both fullerenes and PDIs. Similar to PDIs,
A-D-A type NFAs are easy to synthesize and modify. The DPP, with a high polarity and a strong electron-withdrawing
functionalization of the end-groups, spacers, core units, and side ability, has been commonly used to construct NFAs.
chains could be independently realized, which enables Conjugated NFAs based on DPP typically features (i) easy
numerous structure possibilities. In addition, the HOMO and synthesis and modification, (ii) strong absorption profiles and
LUMO levels of A-D-A type NFAs can be easily tuned by readily tunable energy levels, (iii) a strong tendency to
adjusting the electron rich and deficient strength of the “D” and crystallize, and (iv) high carrier mobilities. Here we summarize
“A” building blocks, respectively. As a result, a large number of recent reports on DPP-based NFA materials (Figure 25 and
NFAs with absorption edges in the NIR region have been Table 3). It is worth noting that most of the DPP-based NFAs
developed by using strong electron donating “D” units and/or were studied in combination with P3HT in BHJ solar cells,
withdrawing “A” units, which is a challenging objective for PDI- partially due to the NFAs’ high-lying LUMO levels.
based NFAs. Similar to the development of PDI dimers, structure
engineering of the spacers of DPP dimers is an effective
method to construct different acceptor materials. The first
6. OTHER NFAS DPP-based acceptor with a PCE above 2% was reported in
Despite the fact that most top-of-the-line PCEs were achieved 2013 by Lin et al. with a structure of 5H-dibenzo[b,d]silole
by either PDI-based or IDT/IDTT-based NFAs, research on (DBS) flanked by two DPP units (DBS-2DPP, Figure 25,
3472 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Table 3. Summary of Other Representative NFAs
mobility (cm2 V‑1 s‑1)
internal LUMO HOMO Eg,opt D:A processing PCE VOC JSC FF device
acceptor ref (eV) (eV) (eV) donor ratio solvent (%) (V) (mA cm‑2) (%) E (neat) E (blend) H (blend) structure ref
DBS-2DPP 6.1 −3.28a −5.30a 1.83 P3HT 1.2:1 DCB 2.05 0.97 4.91 43 3.3 × 10−4 2.8 × 10−5 6.1 × 10−4 conventional 209
Chemical Reviews

F(DPP)2B2 6.2 −3.39b −5.21b 1.77g P3HT 1:1 CF 3.17 1.18 5.35 50.2 2.8 × 10−4 NR NR conventional 210
F8-DPPTCN 6.3 −3.65b −5.31b 1.64 P3HT 1:3 CF+0.4% 2.37 0.97 6.25 39 NR 2.68 × 10−3 9.49 × 10−5 conventional 439
DIO
DPP 1b 6.4 −3.53b −5.29b 1.70 P3HT 2:1 NR 2.69 0.90 5.88 51 NR 2.05 × 10−5 5.87 × 10−4 conventional 211
SF-DPPEH 6.5 −3.60b −5.26b 1.79 P3HT 1:1 CF 3.63 1.10 6.96 47.5 NR NR NR conventional 212
SF(DPPB)4 6.6 −3.51b −5.26b 1.77g P3HT 2:1 CF 5.16 1.14 8.29 55 NR 1.292 × 10−4 1.480 × 10−4 conventional 36
DTDfBT(TDPP)2 6.7 −4.33d −5.85e 1.52 PTB7 1:2 DCB+1% 5.00 0.81 12.10 51 NR 8.2 × 10−4 3.1 × 10−3 inverted 213
CN
DTI 6.8 −3.9e −6.1e 2.21g PBDTTPD 1:1 CB+2% 1.47 1.051 4.6 30 NR NR NR conventional 214
CN
Cor-NI 6.9 −3.24b −6.28d 3.04 P3HT 1:1 NR 1.03 0.82 2.75 46 NR 1.32 × 10−4 NR inverted 215
DIR-2EH 6.10 −3.30b −5.38b 1.97 P3HT 1:4 DCB 3.05 1.22 4.29 58.2 NR 1.26 × 10−4 NR conventional 216
BNIDPA-BO 6.11 −3.6a −5.8d 2.22 PTB7 1:3 CB+DCB 3.08 0.98 8.13 39 NR 5.2 × 10−5 2.4 × 10−4 inverted 217
(9:1)
SubPc (2) 6.12 −3.5c −5.6c 2.03g PTB7 1:1.5 DCB+1% 3.51 0.935 7.8 48 NR NR NR conventional 207
DIO
BT(TTI-n12)2 6.13 −3.53e −5.99e 1.89 DPP-Py 1:2 CF 2.40 1.05 3.72 60 NR NR NR conventional 175
NIDCS 6.14 −3.42b −5.90b 2.22 P3HT 1:2.5 CF 2.71 0.73 8.04 46 NR 5.65 × 10−8 2.00 × 10−5 conventional 218

3473
S4 6.15 −3.73b −5.61b 1.74 p- 1:1 CB+0.4% 1.93 0.87 5.27 42 NR NR NR inverted 220
DTS(FBTTh2)2 DIO
NTz-Np 6.16 −3.60b −6.01c 1.73 P3HT 1:1 CB+DCB 2.81 0.90 5.18 60 NR 1.6 × 10−5 3.8 × 10−5 conventional 221
(4:1)
Ph-MH 6.17 −3.31b −6.01d 2.70 P3HT 1:1 CF 2.05 0.76 5.59 48 5.0 × 10−7 3.3 × 10−6 1.1 × 10−5 conventional 222
Zn(WS3)2 6.18 −3.85b −5.60b 1.55 P3HT 1:0.7 DCB 4.10 0.77 9.1 59 NR 1.9 × 10−4 2.1 × 10−4 inverted 227
DBFI-T 6.19 −3.8a −5.8a 1.73 PSEHTT 1:2 CF 5.04 0.86 10.14 58 0.006f 1.2 × 10−4 2.8 × 10−4 inverted 224
DBFI-DMT 6.20 −3.66a −5.82a 1.91 PSEHTT 1:2 CB+3% 6.4 0.92 12.56 55 0.056f 3.32 × 10−2 1.57 × 10−4 inverted 225
DIO
DBFI-EDOT 6.21 −3.65a −5.72a 1.70 PSEHTT 1:2 CB 8.10 0.93 13.82 63 NR 6.80 × 10−2 7.53 × 10−3 inverted 226
SF-OR 6.22 −3.25b −5.50b 2.15 P3HT 1:0.5 CF 4.66 0.96 7.44 65.0 NR 6.71 × 10−6 8.49 × 10−5 conventional 228
a
Measured by CV using NFA films. bMeasured by CV in solutions. cMeasured by other methods. dCalculated using measured LUMO/HOMO and the optical bandgap. eDetailed method was not
reported. fMeasured using OFET devices. gDetermined from the film absorption onset by the authors of this review; NR = not reported.
Review

DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 26. Chemical structures of selected NFAs with other core units.

molecule 6.1).209 The high LUMO level of DBS-2DPP led to a unit (1b, Figure 25, molecule 6.4) was developed by Yang et
high VOC (0.97 V) for the P3HT:DBS-2DPP device, and an al., showing a PCE of 2.7% with P3HT.211 Another DPP
optical bandgap of 1.83 eV was beneficial for harvesting the tetramer with spirobifluorene as the core unit was reported later
low-energy photons. By changing the spacer to dipropyl by Wu et al.212 Among several acceptors with different alkyl
fluorene and adding a phenyl end-group at each side of the chains, SF-DPPEH (Figure 25, molecule 6.5) with a branched
dimer, a new DPP acceptor F(DPP)2B2 (Figure 25, molecule alkyl chain showed the highest crystallinity after annealing and a
6.2) was developed.210 F(DPP)2B2 had a similar absorption PCE of 3.63%. In a later study, a phenyl ring was added at the
range with DBS-2DPP. With an even higher VOC of 1.18 V, the end of each DPP unit in SF-DPPEH.36 The new acceptor
overall PCE reached 3.17%. Later, Li et al. used cyanothio- SF(DPPB)4 (Figure 25, molecule 6.6) had an upshifted LUMO
phene to replace the benzene as end-group and used octyl alkyl level but a reduced optical bandgap compared to SF-DPPEH.
chains to replace the propyl ones on the central fluorene Therefore, both the JSC and VOC were enhanced, and an
core.439 The new acceptor F8-DPPTCN (Figure 25, molecule excellent performance of 5.16% was achieved with P3HT. The
6.3) showed a lower LUMO level and a narrower bandgap than morphological stability was also demonstrated to be better than
F(DPP)2B2. As a result, in comparison with P3HT:F(DPP)2B2, the fullerene-based devices by a long-time thermal annealing
an increased JSC but a decreased VOC and FF were observed in test.
P3HT:F8-DPPTCN, leading to a slightly reduced PCE. Jung et al. synthesized two DPP-based NFAs. DTBT-
Constructing tetramers with a 3D or quasi-3D structure was (TDPP) 2 had a 4,7-di-2-thienyl-2,1,3-benzothiadiazole
also reported for DPP-based NFAs. For example, a two-layer (DTBT) core flanked by two DPP units, and DTDfBT-
structured DPP tetramer with [2,2]paracyclophane as the core (TDPP)2 (Figure 25, molecule 6.7) had a fluorinated core.213
3474 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

The two NFAs both displayed a bandgap of ∼1.5 eV, but important for tuning the solubility and crystallinity. Using a
fluorination provided DTDfBT(TDPP)2 with downshifted molecular donor DPP-Py (Figure 6) to complement the
HOMO and LUMO levels and stronger intermolecular absorption, a high VOC of 1.05 V and a PCE of 2.40% were
interaction. In combination with the relatively low-crystallinity demonstrated. Another example of linear oligomer was
low-bandgap polymer PTB7-Th, DTBDfT(TDPP)2 showed reported by Kwon et al. The molecule NIDCS (Figure 26,
more crystalline features with reasonably small domains than molecule 6.14) consisted of a phenyl ring as the core unit,
DTBT(TDPP)2 in their respective blends. The higher hole and naphthalimides as the flanking units and a thiophene-ethylene
electron mobilities measured in the PTB7-Th:DTDfBT- spacer.218 Three acceptors with different side chains on the
(TDPP)2 blend gave rise to an improved PCE (5.0%) over phenyl rings were studied. It was shown that the spacers and
the BHJ based on the nonfluorinated molecule (3.0%). It is the central phenyl ring were nearly coplanar, while a dihedral
worth noting that both efficiencies were obtained through the angle of ∼45° was observed between the thiophene and the
aid of 1.0% CN as solvent additive in the dichlorobenzene flanking naphthalimide unit. All three acceptors showed large
(DCB) solution. All of the devices fabricated without CN bandgaps, and the one without any side chains on the central
showed excessively large domains (PCEs below 1.0%), which unit exhibited the best OPV performance (2.71%) with P3HT
indicated that the strong aggregation tendency of the NFAs as the donor. The low electron mobility in the blend (5.65 ×
favored a largely separated phase even with the low crystallinity 10−8 cm2 V−1 s−1) was the limiting factor for achieving high FF
polymer. This is consistent with the crystalline nature of DPP in these blends. The other two derivatives were further studied
and suggests that solvent additive or other methods to with different donors. Kwon et al., in a later study, utilized a
kinetically lock the morphology during film drying are needed polymer donor PPDT2FBT (Figure 6) and a NIDCS derivative
to achieve the small domain sizes. with longer alkoxy side chains.219 With more compatible energy
6.2. Others levels between the donor and acceptor, the VOC was improved
to 1.03 V. A thermal annealing step at 90 °C was also found
Besides the NFAs above, many other structures were also crucial for achieving the high performance, which was explained
explored (Figure 26). Here we categorize them into fused-ring by the increased crystallinity of the acceptor. The best efficiency
aromatics, linear oligomers, twisted dimers, and 3D structures. was 7.64% benefiting from a high JSC of 11.88 mA cm−2.
Their properties along with device performance are listed in Another linear acceptor (Figure 26, molecule 6.15) with
Table 3. chlorinated isoindigo (IID) as the central unit, thiophene as the
6.2.1. Fused-Ring Aromatics. Decacyclene triimides spacer, and naphthalimide as the flanking unit was later
(DTI, Figure 26, molecule 6.8) were fused-ring aromatics reported by McAfee et al.220 The design strategy was that the
reported by Pho et al. in 2013, which consisted of a benzene twisted naphthalimide unit could suppress the overcrystalliza-
core and three fused naphthalimide units.214 The best efficiency tion tendency, and chlorination on the IID unit could improve
was 1.47% with PBDTTPD (Figure 6) as the donor polymer. A charge transport. With a molecular donor p-DTS(FBTTh2)2
possible factor limiting the efficiency was the large bandgap of (Figure 6), the best efficiency was 1.93%. Chatterjee et al.
DTI, which constrained light harvesting in the blue region. In a synthesized a linear acceptor with a naphtho[1,2-c:5,6-c′]bis-
later report, an naphthalimide unit connecting to a corannulene [1,2,5]thiadiazole as the core, thiophene-ethyne as the spacer
(Cor-NI, Figure 26, molecule 6.9) was synthesized by Lu et and naphthalimide as the end-capping groups (Figure 26,
al.215 The main absorption of Cor-NI was also in the UV range, molecule 6.16).221 With P3HT as the donor polymer, the best
and the overall performance was 1.03% with P3HT. In another efficiency was 2.81%. Jinnai et al. demonstrated another
study, Chen et al. reported a rubicence derivative named DIR- example with a simple chemical structure (Figure 26, molecule
2EH (Figure 26, molecule 6.10), which had a strong absorption 6.17).222 The benzothiadiazole-based acceptor had ethyne as
in the range of 400−600 nm.216 Due to a high-lying LUMO the spacer and phthalimide as the flanking unit. The effect of
level, the VOC was as high as 1.22 V with P3HT, and the PCE alkyl chains on the phthalimide unit was investigated by
reached 3.05% at a low D:A ratio of 1:4. Another acceptor with comparing five different alkyl chains. The alkyl chains were
two naphthalimide units fused onto a diphenylanthrazoline unit found to have a strong influence on the energy levels and the
was reported by Li et al. with a name of BNIDPA-BO (Figure crystallinity of the materials and in turn the PV device
26, molecule 6.11).217 BNIDPA-BO had a planar backbone performance. Noticeably, the overall structural similarity
with an absorption edge at ∼550 nm, but the main absorption between the molecules allowed the authors to study the
was in the 300−400 nm region. When PTB7 was used as the structure−property relationship, and a good correlation
donor polymer, the VOC was as high as 0.98 V but the FF was between JSC and the dispersion component of the surface
only 39%, which resulted in an overall efficiency of 3.08%. energy (γd) of the acceptor molecule was observed. The
Ebenhoch et al. reported a derivative of SubPc in solution- authors hypothesized that the branched alkyl chain could
processed BHJ OSCs.207 The acceptor named 2 (Figure 26, enhance the fraction of the amorphous region and thus expose
molecule 6.12) was obtained by connecting SubPc with a more conjugated planes to the D:A interface. The best
phenoxy group, which had an absorption edge of ∼620 nm. performance was achieved using the acceptor with special 2-
Using PTB7 as the donor polymer, a PCE of 3.5% was methylhexyl alkyl chains. When P3HT was used as the donor
achieved. polymer, an efficiency of 2.05% was achieved.
6.2.2. Linear Oligomers. Linear oligomers can be 6.2.3. Twisted Dimers. Similar to the strategy used in
developed by connecting different building blocks together. constructing PDI dimers, other twisted dimers based on other
One example was BT(TTI-n12)2 (Figure 26, molecule 6.13), electron withdrawing building blocks were also developed.
where a thienoimide unit was connected to each side of the Spacers were frequently used to connect the building blocks. In
benzothiadiazole unit via a thiophene spacer.175 This acceptor 2013, the synthesis of a novel building block named
had an almost planar structure with an absorption edge at ∼650 tetraazabenzodifuoranthene diimides (BFI), which had a large
nm. The linear alkyl chain on the imide unit was proved to be planar backbone structure, was reported by Li et al.223 Different
3475 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

spacers were employed to construct dimers with the BFI of 4.10%, which was higher than that of P3HT:PCBM. Another
building block. The first example was DBFI-T (Figure 26, SF-core based 3D acceptor (SF-OR) was recently reported by
molecule 6.19). With thiophene as the spacer, the two BFI Qiu et al., which had a high LUMO level and a high VOC of 0.96
units showed a twisted structure.224 The acceptor had a large V was achieved with P3HT as the donor polymer.228 Despite a
bandgap and a deep LUMO level, and achieved a PCE of 5.04% low mobility, the FF still reached 65%, and a high PCE of
with PSEHTT (Figure 6) as the donor polymer. In a later 4.66% was thus obtained.
study, Li et al. explored four BFI dimers with different
spacers.225 The one with dimethylthiophene as the spacer 7. COMPARISON AMONG DIFFERENT CORE UNITS
(DBFI-DMT, Figure 26, molecule 6.20) showed the largest
dihedral angle between the two BFI units, and the best PCE of PDI-based NFAs exhibit outstanding thermal and chemical
6.4% with the same donor polymer PSEHTT. The improved stabilities, and they could achieve extraordinary electron
PCE for DBFI-DMT over DBFI-T was attributed to the mobilities, proved by their successful application in OTFTs.
enhanced electron mobility of DBFI-DMT and its upshifted However, to be solution processable into nanoscale phase
LUMO level and high EQEs. A subsequent research by Hwang separated grains, approaches such as molecular twisting are
et al. demonstrated a dimer named DBFI-EDOT, where EDOT introduced to overcome the overaggregation caused by
was used as the spacer to further enlarge the dihedral angle to excessively strong intermolecular interaction, which dramati-
76° between the two BFI units.226 With a further enhanced cally reduced the intrinsic high mobility of PDI. Broadly
electron mobility and higher energy levels, the overall speaking, most PDI-based NFAs show an electron mobility in
performance for DBFI-EDOT reached 8.1% with PSEHTT. the range of 10−6 to 10−4 cm2 V−1 s−1, compared to 10−4 to
6.2.4. Three Dimensional Structures. Constructing a 3D 10−3 cm2 V−1 s−1for the IDT-based counterparts in blend films
structure is a challenging but promising strategy to develop with reasonable domain sizes for BHJ solar cell applications.
high-performance acceptors due to the success of fullerenes. An For DPP-based NFAs, an important limitation of their
azadipyrromethene dimer with an zinc atom as the core application is their relatively high-lying LUMO levels, which
(Zn(WS3)2) was reported by Mao et al. in 2014 (Figure 27), is not compatible with most low-bandgap donor polymers.
Regarding the recent rapid development of the A-D-A type
NFAs, besides their own distinctive advantages, the other
nontrivial reason is the existence of a large library of high-
performance large bandgap polymers as electron donor/hole
acceptor that could be chosen to match with the NFAs, which
generally have strong absorption in the red or even NIR region
of the solar spectra. Moreover, the design strategy of large/
medium bandgap polymers is relatively flexible. In contrast,
there are less high performance narrow bandgap polymers,
especially those with absorption in the NIR region, which limits
the maximum current and increases the difficulty in finding the
best match with novel PDI-based NFAs.
Besides PDIs and A-D-A type NFAs, many other structures
Figure 27. 3D structure of Zn(WS3)2. Reprinted with permission were also explored by different groups. Among them, only the
from ref 227. Copyright 2014 John Wiley and Sons. BFI-based dimers and the two 3D acceptors achieved PCEs
higher than 4%. However, one should not underestimate the
showing a 3D structure.227 In spite of the amorphous nature of potential of these new structures because PDIs and A-D-A type
the molecule, the electron mobility in the blend with P3HT NFAs also struggled from limited efficiencies only a few year
could reach 1.9 × 10−4 cm2 V−1 s−1. With an absorption edge at ago. Further molecular design of the new acceptors could lead
∼800 nm, Zn(WS3)2 had a complementary absorption to to continuous improvement of their performance. In addition,
P3HT. The BHJ solar cell of P3HT:Zn(WS3)2 showed a PCE the structure−property relationships obtained in these NFAs

Figure 28. (a) Computed deviation from the molecular plane versus experimental PCE. (b) Computed difference between LUMO+1 and LUMO
energy versus experimental PCE for a device containing a given electron acceptor. The acceptors are classified as described in ref 229. (c) (Lack of)
correlation between molecular size and measured PCE for the considered data set. Reproduced from ref 229 with permission. Copyright 2017 Royal
Society of Chemistry.

3476 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

may also be valuable for the development of any other high- developed, the PCE of this device platform has increased from
performance acceptor materials. ∼5% to >10% within 3 years.199,233−241
As there are increasingly more NFAs emerging each year, it is In 2013, Sharenko et al. combined the molecular donor p-
essential to systematically investigate the common features DTS(FBTTh2)2 with a monomeric PDI acceptor and obtained
shared by the high-performance molecules and seek the a PCE of 3.0% for the solution processed BHJ solar cell using
underlying chemical and physical principle to guide future 0.4 vol % DIO as solvent additive in the chlorobenzene
molecular design. In early 2017, Kuzmich et al. published an solution.120 Later, by tuning the solvent additive in the solvent
analysis article by constructing a database of 80 NFAs (with mixture, Chen et al. achieved a PCE of 3.7% using a
PCE ≥ 3%) with different core units and molecular geometries combination of DTS(FBTTh2)2 and bis-PDI-T,241 while the
and investigated whether common electronic and geometric efficiency was boosted to 5.4% by Kwon et al. using the same
properties of the NFAs such as nonplanarity and molecular size molecular donor and an naphthalimide based NFAs (NIDCS-
show any correlation with the overall device performance.229 In MO).240 The same NFA has been employed by Min et al. and a
line with our discussion, the authors revealed that, despite that PCE of 5.3% was demonstrated with a BDTT based p-type
the highest PCEs for each NFA class were achieved at an molecular donor (BDTT-S-TR).238 The PCE was obtained by
RMSDp between ∼1 and 2, there lacked a direct correlation surveying a series of solvent additive at various concentrations.
(Figure 28a) between the PCE and the nonplanarity of the With the emergence of NFAs such as IEIC and O-IDTBR,
NFA (represented by the root-mean-square distance of the sp2 the PCE of all SMSC has been further enhanced.239 As an
carbon atoms from the best plane passing across them, example, Badgujar et al. used O-IDTBR and a rhodanine donor
RMSDp). Although there were only ∼13 data points with BDT3TR to obtain device with a PCE >7% and an enhanced
high nonplanarity (RMSDp ≥ 2), the fact that all of them thermal stability.237
showed PCE ≤ ∼6% suggests that excessive molecular twisting Recently, the rapid PCE growth in IDT/IDTT-based
is detrimental to charge transport that limits the overall PCE. In polymer:NFA solar cell field has also promoted the develop-
addition, there appears to be no correlation between molecular ment of relevant all SMSCs. Yang et al. synthesized an A-D-A
size (represented by the total number of atoms in a molecule) molecular donor (DRTB-T) using an alkylthienyl-substitued
and device PCE (Figure 28b). However, the authors calculated BDT trimer as the core unit capped with 3-ethylrhodanines.235
the difference between LUMO and LUMO+1 (ΔLUMO) for The wide bandgap donor was combined with IC-C6IDT-IC,188
the neutral molecule and found that all of the 22 NFAs with a narrow bandgap NFA, to obtain an efficient BHJ device with
PCE >6% showed <0.5 eV ΔLUMO (Figure 28c). Despite the efficiency up to 9.08%. The high PCE was enabled by a solvent
fact that the low-lying LUMO+1 level could facilitate charge vapor annealing (SVA) process: a 30 s dichloromethane SVA
separation at the donor/acceptor interface, the discovery of increased the PCE from 5.03% (as-cast) to 7.47% and a 60 s
such an apparent correlation between an isolated property of dichloromethane SVA further enhanced the PCE to 9.08%.
the NFA and the overall device efficiency is still intriguing. This indicated a dramatic morphology evolution during SVA
More recently, Lopez et al. conducted a statistical analysis of and was confirmed by morphological characterization, which
∼51 000 NFAs. Using DFT calculations, the authors computed revealed a clear increase in crystallinity of the blend film, mainly
the frontier molecular orbital energies and related them to the caused by the enhanced intermolecular order of the acceptor.
PCEs through the Scharber model.230 Through machine In a more recent work by Bin et al., the authors designed a
learning methods, they calibrated theoretical data to exper- pair of small molecular donors inspired by the success of the
imental results, which allowed them to provide top NFA BDTT-alt-FBTA D−A copolymers. The two small molecular
candidates and design rules for future studies. donors have different conjugation lengths enabled by the
As more and more NFAs are developed, the top-down difference in their side chains. Donor H11 had thienyl
analysis, i.e., building correlations from a large set of conjugated side chains while the other donor H12 had alkoxy
experimental data,229,230 could serve as a supplement to the side chains on the BDT unit. In combination with a highly
bottom-up approaches, i.e., using fundamental working crystalline low bandgap NFA IDIC, the blend of H11:IDIC and
principle to guide molecular design and device fabrication. H12:IDIC yielded the best PCEs of 9.73% and 5.51%,
respectively, with a thermal annealing at 120 °C for 10 min
8. NONFULLERENE ALL SMALL MOLECULE OPV performed prior to electrode deposition. Compared to the
Nonfullerene all small molecule solar cells (all SMSCs) PCEs of the as-cast samples (5.37% for H11:IDIC and 1.83%
consisting of a p-type molecular donor and an n-type NFA for H12:IDIC), thermal annealing caused a remarkable
inherit most of the advantages of polymer:NFA solar cells while enhancement to the device. Through a series of morphological
improve batch-to-batch reproducibility. Historically, due to characterizations (Figure 29), thermal annealing has been
their vacuum deposition ability and poor solution process- shown to induce prominent self-aggregation of the small
ability, evaporated planar-heterojunction all-SMSC is the main molecules and cause enhanced and more balanced hole and
architecture to utilize the small molecules.65 Whereas, the electron mobilities for both material sets. Overall, the extended
development of solution processed BHJ-type nonfullerene all conjugation of H11 and the extra crystallinity introduced by
SMSCs has been limited by two main restrictions including (i) thermal annealing provided a huge enhancement to the
overmixing between small molecular donor and acceptors in performance of the device, which suggested a lack of phase
the blend and (ii) low electron mobility of the NFA.120,231,232 separation between the components in these two systems
Since both components of this device platform do not possess originally. Furthermore, on the basis of this work, Qiu et al.
long backbone (like a polymer does) as the charge carrier designed another molecular donor using cyano end-capping
transport highway, a lack of phase separation between the groups and improved the donor:IDIC photovoltaic device
components would likely induce free carriers to be confined performance to 10.11%.234
within a limited space by grain boundaries. In spite of these Most of these reports utilize solvent additives or post-
limitations, thanks to the increasingly numbers of NFAs being treatments such as SVA or thermal annealing to tune the film
3477 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

be surprising to observe further breakthroughs upcoming in the


near future.

9. TERNARY BLEND NONFULLERENE OPV


9.1. Motivations to Use Ternary Blend
To maintain the simplicity of the single-junction structure and
in the meantime broaden the absorption range of the OSCs,
ternary blend bulk heterojunction (BHJ) OSCs have been
developed where two donors and one acceptor or one donor
and two acceptors are blended together during device
fabrication. The original motivation for introducing a third
component into the active medium is to broaden the
absorption spectrum. Despite this beneficial effect on JSC, the
VOC of ternary blend solar cells has originally been predicted to
be pinned to the smaller VOC of the corresponding binary
blends of the constituent components,242 thus increasing the
overall energy loss of the device. However, ternary OSCs have
gone beyond expectations since the demonstration of devices
whose VOC change continuously with the film composition
(Figure 30a,b).243−245
The origin of the tunable VOC is complicated.247−251
Mainstream models such as organic alloy formation246,252 and
parallel-like BHJ,245,253 were used to explain such composi-
tional dependence of VOC. Street and co-workers revealed that
the charge transfer state energy also showed a dependence on
film composition (Figure 30c, d), which indicated the
formation of an organic alloy between the two acceptors (or
two donors).246 By comparing two donor:donor:acceptor
ternary systems where two different pairs of donor polymers
have either large or small interfacial tension, Gobalasingham et
al. proposed that the minimized interfacial tension between the
two donors, which allows the donors to better intermix, is the
prerequisite for the alloy formation and tunable VOC.254
Morphological studies and ionization potential measurements
Figure 29. (a) Line cuts of the 2D-GIWAXS patterns of (b) H11, (c) on these ternary systems by Khlyabich et al. further confirmed
IDIC, (d) H11:IDIC (as-cast), and (e) H11:IDIC (thermal annealed). the effectiveness of using surface tension as a proxy to predict
Adapted with permission from ref 233. Copyright (2017) American organic alloy formation (Figure 31).249 The same correlation
Chemical Society. has been observed by Jiang et al. on a series of nonfullerene
ternary blends with one polymer donor and two small
molecular acceptors.255 When the two acceptors (with different
LUMO levels) had a small interfacial tension (γ), the VOC of
morphology. For most of them, the result of these
the ternary devices demonstrated a tuned VOC with respect to
morphological tuning techniques is the enhanced crystallinity
the weight percentage of the acceptor, whereas the devices
for at least one of the components and enhanced phase
showed a pinned VOC when a NFA (ITIC-Th) was combined
separation between the donor and acceptor. The general trend
with PC71BM due to the large interfacial tension between these
for all SMSC is the propensity of forming an overmixed phase
acceptors. In spite of these explanations, the result is that,
due to the absence of polymerization that reduces the entropy
through the introduction of the third component, the ternary
gain during mixing. Therefore, designing a BHJ of this type
device’s VOC could be tuned between the VOC of corresponding
should start with a NFA with a high crystallization tendency. In
binary blend devices, which is a considerable bonus effect in
fact, Figure 29 shows that even IDIC, whose pure film showed a
addition to the absorption spectrum broadening. This
high crystallinity, lost most of the crystalline behavior in a blend
significantly increases the perceived impact of this device
film with H11, which further demonstrated the high propensity
platform.
for two small molecules to mix. With a careful optimization of
solvent additives, SVA, or thermal annealing, the morphology 9.2. Benefits of NFAs in Ternary Blend OPVs
of the film can be modified to a state where charge transport is Similar to the binary blend BHJ, a large fraction of previously
enhanced while the initial intermixing is not totally destroyed. reported ternary blends contain a significant amount of
These recent advances demonstrate the great potential of the fullerenes,256−263 whose absorption cannot be easily tuned
all SMSC. The flexible synthesis and readily tuned and is relatively weak. Hence, in order to broaden the
spectroscopical and electrochemical properties of both absorption, the majority of the effort was focused on exploring
components in this device platform has provide numerous the option of two polymer donors and one fullerene
potential combinations to further improve the device perform- acceptor.254,261,264−271 However, polymers are large molecules.
ance. As both branches (small molecule donor and small From Flory−Huggins theory, the entropy gain for mixing
molecule acceptor) are developing in a fast pace, it would not monomers is S = −R(ϕ1 ln ϕ1 + ϕ2 ln ϕ2), where ϕ1 and ϕ2 are
3478 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 30. (a) Structures and corresponding HOMO and LUMO energy levels of P3HT, ICBA, and PC61BM. (b) VOC for the ternary blend BHJ
solar cells as a function of the amount of ICBA in the blend. (a, b) Adapted with permission from ref 244. Copyright (2011) American Chemical
Society. (c) Photocurrent spectra response data for the P3HT:PCBM:ICBA (D:A1x:A2(1−x)) ternary blend solar cells plotted as a function of ICBA
fraction in PCBM:ICBA pair. The inset indicates the CT transition or interface band gap that is being measured, and the pair of dashed lines
indicates the range over which the interface band gap energy is extracted. (d) Expanded plot of the peaks near 1.7 eV with the background
subtracted. The peak centered above 1.7 eV corresponds to PCBM absorption, and the peak centered below 1.7 eV corresponds to ICBA. (c, d)
Adapted with permission from ref 246. Copyright (2013) American Chemical Society.

the volume fractions of monomer 1 and 2. In contrast, when still possess the aforementioned limitations such as weak and
mixing two polymers, the entropy gain becomes less tunable absorption, which reduces the potential for their
ϕ1 ϕ2 application in ternary blend OPVs.
S = −R ( N1
ln ϕ1 + N2 )
ln ϕ2 , where N1 and N2 are the degree
NFAs, in contrast, do not have such a strong phase
of polymerization of polymer 1 and polymer 2, respectively. separation tendency as two polymers or the common problems
This suggests that the entropy gain for mixing two polymers is possessed by fullerenes. Their feasible synthesis and easily
much less than mixing two monomers/small molecules, which tunable optical properties make them ideal candidates for
is also the main reason for the enhanced difficulty to reduce enhancing JSC and VOC simultaneously in ternary blends.
phase separation in all-polymer BHJ solar cells. Therefore, to Therefore, designing a pair of complementarily absorbing
reduce ΔGmixing, the enthalpy of mixing, RTχϕ1ϕ2, must be miscible NFAs as acceptors in ternary blends is a promising
minimized, which creates a criterion that the Flory−Huggins route to improve the performance of single-junction OPVs. We
interaction parameter χ cannot be too large for achieving note that, despite being beyond the scope of the review, this
adequate mixing. This explains why the tunable VOC in the idea has already been proved effective in multiple all small
ternary blend with two polymers and one acceptor only occurs molecule solar cells where two small molecule donors with
when the two polymers have excellent molecular miscibility. decent miscibility and complementary absorption were blended
Compared to polymer blends, without the reduced entropy with a fullerene acceptor and provide improved PCE over the
gain during mixing, two small molecules are more likely to form binary devices.259,272,273 In addition to device efficiency
an intermixed phase. For example, when PC61BM and ICBA are improvement, it has also been suggested by multiple reports
used in a ternary blend of P3HT, the device VOCs display an that introducing the third component into the binary blend can
almost linear compositional dependence,244 whereas fullerenes enhance the device stability. This may have huge impacts on
3479 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 32. Visual illustration of the binary P3HT:IDTBR blend with


Figure 31. (a) Relationship between interfacial tension and VOC of the IDFBR presence, wherein the crystallinity of both P3HT and IDTBR
ternary blend device. Adapted with permission from ref 254. Copyright is preserved. Adapted with permission from ref 26. Copyright 2017
(2016) American Chemical Society. (b) VOC as a function of ITIC-Th Nature Publishing Group.
weight percentage for devices with a polymer:ITIC-Th:acceptor
ternary blend active layer. γ is the interfacial tension between the
two acceptors: a small γ corresponds to a quasi-linear relationship ance compared to the binary counterparts, owing mainly to the
while a large γ provides a pinned VOC. It is noted that PC71BM has a increased JSC and VOC. Besides efficiency, the NFA ternary
lower LUMO level than ITIC-Th while the other three NFAs have device demonstrated outstanding dark and photostability in air.
higher LUMO levels than ITIC-Th, which is accounted for the reverse Another work related to this design rule is reported by Liu et
trend in their VOCs. Reprinted with permission from ref 255. al., who also utilized two complementarily absorbing NFAs in a
Copyright 2016 John Wiley and Sons. ternary blend, one of which (SdiPBI-Se) absorbs strongly in
∼400−600 nm while the other (ITIC-Th) in ∼600−800 nm
reducing the energy payback time and the effective cost of (Figure 33).275 In combination with a polymer (PDBT-T1)
OPVs, which will be discussed in more details in the stability- that absorbs strongly at ∼500−700 nm, the ternary blend solar
focusing section 10. In section 9.3, we focus exclusively on the cells, at an optimal composition, were able to achieve
materials and photovoltaic device efficiency of ternary blend
OPVs with at least one NFA employed. For more details
regarding fundamentals of ternary OPV and discussions on
other types of ternary blend OPVs, review articles are
available.263−265,274
9.3. Utilizing NFAs in Ternary Blend OPVs
9.3.1. Pair of Miscible NFAs in Ternary Blend. A typical
example of employing two compatible NFAs with comple-
mentary absorption was given by Baran et al.26 The authors
employed IDTBR, a NFA working extremely well with the
prototypical polymer P3HT, and an indenofluorene analogue,
IDFBR, as acceptors in ternary blends with either P3HT or
PTB7-Th as the donor polymer. The absorption maxima for
IDTBR and IDFBR were at ∼690 nm and ∼530 nm,
respectively, which provided the ternary blend films with a
wider spectral range than the binary counterparts did. Despite
their similar chemical structures, differential scanning calorim-
etry (DSC) and morphological studies revealed different
behaviors between IDTBR and IDFBR in the ternary blend: Figure 33. (a) Chemical structures of PDBT-T1, SdiPBI-Se, and
the more diffusive and less crystalline IDFBR dispersed in both ITIC-Th. (b) Normalized UV−vis absorption spectra of neat PDBT-
crystalline polymer phase and crystalline IDTBR phase (Figure T1, SdiPBI-Se, and ITIC-Th films. (c) Schematic diagram of ternary
32), allowing the formation of a three-component redox organic solar cell device structure. (d) Energy level diagrams for
cascade. As a result, both P3HT based and PTB7-Th based PDBT-T1, SdiPBI-Se, and ITIC-Th. Reprinted with permission from
ternary blend devices showed enhanced photovoltaic perform- ref 275. Copyright 2016 John Wiley and Sons.

3480 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

significantly enhanced JSC compared to either binary device through tuning the blend film morphology. As a result, a PCE
based on the constituent components. In addition, despite the improvement from ∼6% binary to ∼7% ternary was achieved.
structural dissimilarity between the two NFAs, i.e., one based The same NFA, ITIC, has been utilized by Zhong et al. in a
on perylene bisimide while the other containing an IDT core, ternary blend with two conjugated polymer donors, J51 and
the two NFAs demonstrate decent miscibility confirmed by the PTB7-Th, and an improved ternary device PCE of ∼9.7% was
morphological characterizations, which led to a linear relation- observed by carefully tuning the film composition.280
ship between VOC and film composition. As a result of both In ref 194, to further utilize the photons in shorter
improved JSC and tunable VOC, the best ternary device wavelength region, ternary solar cells with a small amount of
demonstrated a PCE ∼10.3%, much higher than the binary a medium bandgap donor polymer J52 was added to the blend
counterparts (∼8.2% and ∼6.4%). of PBDTTT-EFT:IEICO-4F, and the efficiency was improved
In line with these work, Yu et al. employed two NFAs with from 10.0% to 10.9%, benefiting from the exceptionally high JSC
IDT cores, IT-M and IEIC-O,194 whose bandgaps were 1.58 of 25.3 mA cm−2.
and 1.37 eV, respectively, in a ternary blend in combination 9.3.3. Combining NFA and Fullerene in a Ternary
with a large bandgap polymer, J52.276 The similar LUMO levels Blend. NFAs have also found their application in ternary blend
between IDT and IT-M offered their binary BHJ devices, when OPVs in conjunction with fullerene to enhance photovoltaic
each was blended with J52, almost identical VOCs, but the performance.
elevated HOMO level of IEIC-O caused a red-shifted Chen et al. used a tiny amount of fullerene to sensitize the
absorption, which provided more complementary solar binary blend formed by a red-to-NIR absorbing NFA, IEIC-O
spectrum coverage and was the main source of the improve- and a medium bandgap polymer PBDTTT-E-T.281 With a
ment in JSC (from binary 17.1 to ternary 19.7 mA/cm2) and minute amount of only 5% (w/w) in the blend, the ternary
thus PCE (from 9.4% binary to 11.1% ternary) at a IT-M:IEIC- devices were able to achieve increased JSC and FF. Due to the
O weight ratio of 4:1. Besides, the structural similarity provides low fraction of fullerene, the enhanced current cannot be
the NFAs with compatible morphology, i.e., the blend simply explained by the broadened absorption. Instead, through
morphology of IT-M and IEIC-O can be viewed as the morphological investigation and charge transfer mechanism
compositional average of morphology formed by the individual studies (Figure 34), the authors concluded that the fullerenes
material. This is another factor for maintaining the high FF in that have higher LUMO levels than IEIC-O, serving the dual
the ternary blend device. roles of facilitating electron transfer by forming an energetic
Furthermore, Jiang et al. systematically studied the device cascade and adjusting the blend morphology toward the
performance of a series of polymer:NFA1:NFA2-based ternary direction of reducing geminate recombination. Another work
blends, where NFA1 and NFA2 are two different nonfullerene
reported by the same group also employed bis-PC71BM as the
acceptors.255 By comparing four different combinations of
third component and added it into the polymer:NFA blend.282
NFA1 and NFA2 including the structurally similar ITIC-Th
At a weight ratio of 1:1:0.2 (PBDB-T:IT-M:bis-PC71BM), the
and IEIC-Th, the structurally dissimilar ITIC-Th and SF-PDI4
ternary device showed an optimized PCE of 12.20% with
or TPE-PDI4, the authors not only confirmed the role of
simultaneously enhanced JSC, VOC, and FF. It was concluded
surface tension in influencing device behavior (tuned vs pinned
VOC) but also provided a morphology control method for that more effective dissociation of excitons at the polymer:-
constructing nonfullerene ternary blends. It was found that the acceptor interface was the main factor contributing to the
morphology of the ternary blend could be effectively controlled device performance enhancement.
through the use of two NFAs with a small interfacial tension In another report by Fan et al.,283 the authors also took
and a polymer with a temperature dependent aggregation advantage of the electronic property of the NFA, IFBR, and
property. An enhanced PCE (11.2%) was achieved despite the used it to design a charge transfer cascade. Both IFBR’s HOMO
similar absorption ranges of the two NFAs, which was and LUMO lie between the polymer and fullerene, which
attributed to the enhanced electron transport through (higher dramatically suppressed trap-assisted recombination in the
FF) and elevated LUMO level of the third component (higher ternary blend device. As a result, together with the broadened
VOC). Su et al. reported a ternary blend consisting ITIC and spectral coverage by IFBR, the best ternary blend exhibited an
IDIC as acceptors and a large bandgap polymer PSTZ as >40% increase in PCE compared to the polymer:fullerene
donor.277 However, ITIC and IDIC have almost identical binary device.
absorption range. The authors attribute the device performance Difference in molecular interaction between NFA and
enhancement to the improved morphology and charge fullerene in a ternary blend could cause variations in phase
transport by tuning the ratio of the two NFAs. These results separation. One report by Chen et al.284 employed the organic
imply that in addition to broadening absorption range and alloy formation between the NFA (TPE-4PDI) and PC71BM to
increase VOC, blending two NFAs also has a potential to fine- explain the tunable VOC in the ternary blend PTB7-Th:TPE-
tune the blend morphology. This clearly is another advantage 4PDI:PC71BM, despite the drastically different chemical nature
for this device platform as tuning the morphology278 in a between the two acceptors. The authors used DFT calculations
controllable fashion has significant implication for charge to simulate the composition dependence of VOC. At a mass
generation and transport of OPVs. electron DOS ratio of 2:1 between PC71BM and TPC-4PDI,
9.3.2. Blending NFA with a Pair of Polymer Donors. the calculated VOC curve fit well with the experimental data,
The idea of broadening spectral range using NFA has been which provides support to their model. In another ternary
extended to all-polymer solar cells. Ding et al. added 6 wt % system reported by Liu et al.,253 where ITIC-Th and PC71BM
ITIC into a polymer donor:polymer acceptor blend (PTP8:P- were blended together with a wide bandgap polymer (PDBT-
(NDI2HD-T)).279 Compared to both polymers, ITIC has a T1), the quasi-linear relationship between VOC and ITIC-Th
red-shifted absorption. Besides, the addition of ITIC played weight ratio was interpreted by a parallel-like BHJ model. The
another critical role of balancing hole and electron mobility authors carried out grazing-incidence X-ray diffraction
3481 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

blends. Additionally, studies on ternary device stability are


needed to figure out whether the introduction of fullerenes in
NFA OPVs decrease the lifetime of the devices.

10. ENHANCED DEVICE STABILITY OF


NONFULLERENE OPV
As the maximum reported PCE reached above 13%, OPV
device’s operational stability becomes increasingly vital for the
objective of commercialization. Possible degradation routes of
OPVs are manifold.17,285−287 For example, the morphology
optimized for device efficiency might not be in thermodynam-
ical equilibrium, which causes the nanoscale morphology to
change during operation, particularly for the fullerene involved
systems due to their high diffusivity.288−291 In addition,
dimerization of fullerenes caused by photochemical reaction
further increases device instability.288−291 Besides, molecular
oxygen has been shown to cause both reversible and irreversible
effects on performance of the devices based of π-conjugated
systems.17,286 As fullerenes play a role in most of these
degradation pathways, the replacement of fullerenes with NFAs
could be beneficial for device longevity.
10.1. Enhanced Stability of BHJ Solar Cells with IDT Based
A-D-A Type NFAs
On the basis of the improved device longevity for P3HT:O-
IDTBR demonstrated by Holliday et al.25 (section 5.2.1), Baran
et al. further revealed that an 800 h exposure to air made the
P3HT:PC71BM BHJ solar cell nonfunctional, whereas the
device retained 70% of the efficiency under the same condition
once the fullerene was replaced with IDTBR (Figure 35).26 The
same trend had been observed in these devices under 1.5 AM
illumination, unencapsulated in air.
Later, Gasparini et al. compared the “burn-in” effect (a rapid
decay in the early stage followed by a flattened slower decay at
longer times) between P3HT:PC71BM and P3HT:IDTBR
Figure 34. (a) Electron spin resonance (ESR) response of the devices.27 It was concluded that fullerene dimerization was the
PBDTTT-E-T:Bis-PC71BM:IEICO ternary blends. (b) Schematic of primary cause of the photoinduced burn-in. Replacing PC71BM
charge transfer in ternary blend films. (c) EQE spectra of PBDDTTT-
eliminated not only dimerization but also the disorder induced
E-T:IEICO binary and PBDTTT-E-T:Bis-PC71BM:IEICO ternary
blend devices. Reprinted with permission from ref 281. Copyright voltage loss.
2017 John Wiley and Sons. In a later work by Wadsworth et al., the stability of
nonfullerene solar cells fabricated using a low bandgap polymer
(PffBT4T-2DT) and an A-D-A type acceptor (EH-IDTBR)
(GIXRD) and R-SoXS studies and concluded that ITIC-Th and was systematically compared among various processing solvents
PC71BM form their own transport network. for the active layer.28 The authors showed that the devices
In general, NFA and fullerenes have rather different chemical whose active layers were processed from hydrocarbon solvents
structures and electronic and material properties. Different such as o-xylene, mesitylene, and trimethylbenzene showed
miscibility between NFA and fullerene may cause a more reproducible device performance. In addition, the
fundamentally different working principle of ternary blend mesitylene processed nonfullerene device maintained ∼92%
solar cells. For example, both organic alloy and parallel linkage of its initial PCE after 4000 h storage in a N2 atmosphere and
models have been used to explain device behavior in different ∼80% initial PCE after 250 h under illumination, which
systems, indicating dramatically different interaction between outperformed the devices processed from chlorobenzene.
fullerene and NFA in the blends. Meanwhile, miscibility and The workhorse A-D-A type NFA, ITIC, showed excellent
morphology characterization are highly demanded in con- thermal stability in the BHJ solar cells with PBDB-T as the
junction with the energy diagram and recombination analysis polymer donor.29 In comparison to PBDB-T:PC71BM, PBDB-
when the model of charge transfer cascade is used to explain T:ITIC based device exhibited little PCE decline after being
device performance since the energetically favorable transition thermally annealed at 100 °C for 250 h, while the former
may be spatially forbidden (or vice versa). Broadly speaking, dropped by >40% of its initial value (Figure 36). In line with
different mechanisms may coexist in determining the overall these reports, decent thermal stabilities were also presented in
device performance. Therefore, interpreting device result in other A-D-A type NFA based solar cells.30,31
NFA/fullerene blends requires a better understanding on the Choi et al. synthesized a novel conjugated polymer, 3MT-Th,
molecular interaction among polymer, fullerene, and NFA. and evaluated its performance in BHJ solar cells with ITIC as
Furthermore, systematic investigation of the voltage loss the electron acceptor.32 The active layer was fabricated using
mechanism is also needed to clarify the role of NFA in such nonhalogenated solvent. Compared to PTB7-Th:ITIC, 3MT-
3482 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 35. (a) Shelf storage lifetime (dark, in air) comparison of Figure 37. Evaluation of device performance stability under ambient
P3HT:IDTBR:IDFBR device efficiencies with other polymer:fullerene conditions up to 1100 h storage. (a) 3MT-Th:ITIC processed from
systems. Devices were exposed to ambient conditions over a 1200 h toluene with 0.25% diphenyl ether (DPE) and (b) PTB7-Th:ITIC
duration or until devices no longer showed any diode behavior. (b) processed from toluene with 0.25% DPE. The average photovoltaic
Photostability of P3HT:IDTBR:IDFBR device and polymer:fullerene parameters were obtained from four devices. Reprinted with
solar cells (in air, unencapsulated, under AM1.5 illumination at 1 sun) permission from ref 32. Copyright 2017 John Wiley and Sons.
for 90 h. Adapted with permission from ref 26. Copyright 2017 Nature
Publishing Group.
In the most recent highly efficient nonfullerene solar cells by
Zhao et al.33 and Zhao et al.,19 the ITIC derivatives ITIC-Th1
and IT-4F outperform PC71BM in terms of not only device
efficiency but also device longevity. The same observation has
also been made by Choi et al. on PTB7-Th:IDT(DCV)2
nonfullerene solar cells compared with the fullerene counter-
parts.34
In a later report by Guo et al., the authors presented a 10.5%
PCE for nonfullerene solar cells based on PTZ1:IDIC.
Preliminary evaluation revealed decent thermal stability for
the unencapsulated devices with a device area of 0.81 cm2. After
a 10 h thermal annealing at 100 °C, the PCE of device still
retains 81% of the original value.
10.2. Enhanced Stability of BHJ Solar Cells with PDI or DPP
Figure 36. PCEs of the PBDB-T:PC71BM and PBDB-T:ITIC devices Based NFAs
against time of thermal annealing at 100 °C. Reprinted with
permission from ref 29. Copyright 2016 John Wiley and Sons. Since most of the top-of-the-line PCEs of nonfullerene solar
cells were achieved using ITIC or other IDT core based NFAs,
there are relatively fewer studies on device longevity with
acceptors based on PDI or other core units. However, rylene
Th:ITIC based devices exhibited superior shelf lifetime with no diimides and DPP based molecules are well-known for their
obvious PCE decay after 1000 h (Figure 37), which was pronounced thermal and chemical stabilities. 103,143 As
attributed to the robust morphology formed between 3MT-Th improved PCEs are being progressively published for PDI-
and ITIC. By measuring the UV−vis spectra of both pure and based NFAs, systematic studies on PDI based nonfullerene
blend films under illumination, the authors observed a fast devices are expected to emerge. The work shown below
decay of the 0−0 vibronic peak in the neat PTB7-Th film, manifest the huge potential of using PDIs to improve device
which was not observed for 3MT-Th. In addition, from the longevity.
variation of the 649 and 703 nm peaks of the UV−vis spectra Lin et al. have evaluated the ambient stability of nonfullerene
upon illumination time (in air), the authors attributed solar cells based on a series of three-dimensional PDI based
enhanced stability of 3MT-Th:ITIC to mainly the better acceptors.35 It was shown that the intramolecular twisting of
photostability of the polymer in the blend. these acceptors not only produced a dramatic difference in
3483 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 38. (a) Normalized UV−vis−NIR absorption spectra of PTB7-Th and ATT-2 in thin films. (b) The standard human eye sensitivity spectra.
(c) Shelf stability of ATT-2:PTB7-Th in ambient conditions. The inset shows photographs of the institute abbreviation without (top) and with
(bottom) the semitransparent device. (a−c) Reprinted with permission from ref 292. Copyright 2017 John Wiley and Sons. (d) Absorption spectra
of PTB7-Th and IHIC pure films and PTB7-Th: IHIC blended films. (e) Visible transmission spectrum of the optimized semitransparent device
with the architecture of indium tin oxide (ITO)/zinc oxide (ZnO)/PTB7-Th:IHIC(100 nm)/MoO3/Au(1 nm)/Ag(15 nm). Inset shows the
structure of the acceptor molecule. (f) Photographs of Bo Ya Pogoda in Peking University taken with exactly the same camera settings on a sunny
day: (left) without any filter; (right) with semitransparent device based on the architecture of ITO/ZnO/PTB7-Th:IHIC(100 nm)/MoO3/Au(1
nm)/Ag(15 nm). (d−f) Reprinted with permission from ref 293. Copyright 2017 John Wiley and Sons.

device performance but also resulted in a significant variation in surprising considering the more thermally stable and less
device shelf stability in air. The efficiency of the device based on diffusive nature of the A-D-A NFA.262
TPC-PDI4, which was the NFA with the most intramolecular Shastry et al. demonstrated another avenue to improve
twisting, decreased to a level of 76.8% of the original PCE after device stability.37 By adding high mobility single-walled carbon
2 weeks. In contrast, the less twisted core offered TPPz-PDI4 nanotubes as the third component in a nonfullerene blend
and TPE-PDI4-based devices with excellent shelf stability by formed by PTB7 and a helical PDI dimer(hPDI2), the ternary
maintaining over 90% of the original PCE, which outperformed device showed reduced bimolecular recombination and
that of the fullerene-based one. This result might be caused by enhanced stability at both 1 sun and concentrated illumination.
the excessive twisting of the TPC core that could facilitate 10.4. Semitransparent Nonfullerene Solar Cell with
molecular diffusion in the blend, which is detrimental for Enhanced Stability
morphological stability.
Li et al. investigated thermal stability for nonfullerene solar In addition to stability, the readily tunable optical property of
cell devices consisting of P3HT and a DPP based NFA, NFAs made them ideal candidates for fabricating semi-
SF(DPPB) 4.36 After 3 h of thermal annealing at 150 °C, while transparent photovoltaic devices. In ref 292, the authors
the nonfullerene device showed no evidence of PCE decay, the demonstrated a semitransparent OPV device fabricated using
fullerene based counterpart displayed an efficiency below 1%. PTB7-Th:ATT-2, whose main absorption located in the range
of 600−850 nm, which allowed high transmittance of the blue
10.3. Enhanced Stability of Ternary Blend Nonfullerene photons (Figure 38a−c). The normal device with thick
Solar Cells electrode showed a high PCE of 9.58% with decent ambient
In the work mentioned in section 10.1,26 in addition to the stability, i.e., >8.5% PCE after 14 days in air without
enhanced stability obtained by replacing fullerene with IDTBR, encapsulation. By reducing the Ag electrode thickness to
interestingly, the authors found that the ternary device based on below 20 nm, the semitransparent device showed a high PCE of
a blend of P3HT:IDTBR:IDFBR (1:0.7:0.3 weight ratio) 7.74%.
demonstrated further enhanced dark and photostability A more recent publication by Wang et al. demonstrated the
compared to the binary devices. 80% initial performance was synthesis of a novel NFA, IHIC, which exhibited strong
retained after 800 h exposure to air and 85% was retained after absorption in the NIR region (∼600−900 nm) with a superior
90 h in air under illumination for the ternary device. electron mobility.293 The authors exploited the narrow bandgap
Wang et al. reported that using m-ITIC as the third of IHIC (1.38 eV) and blended it with the archetypal narrow
component could significantly enhance the thermal stability bandgap polymer, PTB7-Th (bandgap ∼1.58 eV). The
compared to the PTB7-Th:PC71BM binary blend, which is not resultant BHJ solar cell manifested not only a high PCE of
3484 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

9.77% but also a high average visible transmittance of 36% due qΔV = Eg − qVOC
to the minimal absorption of the blend film in the green-blue SQ SQ rad rad
wavelength region (Figure 38d−f). Furthermore, an initial = (Eg − qV OC ) + (qV OC − qVOC ) + (qVOC − qVOC)
stability test revealed that the unencapsulated semitransparent SQ rad,belowgap non − rad
= (Eg − qV OC ) + qΔVOC + qΔVOC
device retained ∼97% of its initial PCE after a continuous
= ΔE1 + ΔE2 + ΔE3
illumination for 130 min (AM 1.5G, 1 sun).
(2)
10.5. Summary
where q is the elementary charge, ΔV is the voltage loss, VSQOC is
Despite the limited number of publications on device stability the maximum voltage attainable according to the Shockley−
so far, substitution of fullerenes with NFAs appears to fulfill the Queisser limit (which assumes no absorption below the optical
expectation of stability improvement. However, other degrada- gap of the cell),300 Vrad
OC is the open-circuit voltage when only
tion routes still affect device longevity strongly, such as the radiative recombination is present, ΔVrad,belowgap is the voltage
OC
oxidative degradation of the donor as well as the interfacial loss of radiative recombination from sub-bandgap absorption,
layer degradation. Intriguingly, the strategy of designing ternary and ΔVnon−rad is voltage loss due to nonradiative recombina-
OC
blend nonfullerene OPVs may not be necessarily detrimental to tion.
device longevity. Future research on NFAs should not only The first term of the energy losses in eq 2 (Eg − qVSQ OC)
focus on their photovoltaic PCE but also on the device corresponds to radiative recombination originating from
longevity and other factors that affect eventual commercializa- absorption above the bandgap. This loss is unavoidable for
tion. Introducing industrial figure of merit is necessary in the any type of solar cell and is typically between 0.25 and 0.30 eV.
near future to evaluate the full potential of a new material.294 The second term (qΔVrad,belowgap
OC ) corresponds to additional
Considering that most OPVs in the literature were not radiative recombination from absorption below the bandgap.
optimized for device longevity, there is a huge potential for For inorganic and perovskite cells, this term is negligible due to
chemists and device engineers to design more stable NFAs and steep absorption edges and thus can be neglected from the
further improve device stability to essentially reduce the energy diagram shown in Figure 39.297 In contrast, this loss is
efficiency-stability-cost gap. typically much higher for OPVs due to the presence of CTS at
the donor−acceptor heterojunction as evident from the red-
11. REDUCING VOLTAGE LOSS shifted absorption spectrum.301 As discussed in section 2, OPVs
typically rely on an energy offset between the electron affinity
One of the most important factors that limit the efficiency of of the donor and acceptor materials to provide a driving force
OPVs is the relatively large voltage loss from the bandgap (Eg) for overcoming the binding energy of the excitons to form
of the light absorber to the device open-circuit voltage (VOC). charge carriers. The requirement of this energy offset (typically
For example, one of the best performing OPVs reported to >0.2 eV) to enable efficient charge generation leads to
date, PffBT4T-2OD:PC71BM,295 has a VOC of 0.77 V for an Eg significant voltage loss. For examples, this loss is as large as
of 1.65 eV, corresponding to a voltage loss of 0.88 V. In 0.67 V for P3HT:PCBM blend and 0.2 V for the state-of-the-art
contrast, state-of-the-art c-Si or perovskite cells have much PTB7:PCBM blend.
lower voltage losses, typically in the range of 0.40−0.55 In order to minimize this loss in OPVs, it is necessary to
V.86,87,296 In order to understand the reason behind the large reduce the energy difference between the singlet exciton on the
voltage loss suffered by OPV, we have to review the origins of donor and/or acceptor and the interfacial CTS without
voltage losses in a solar cell. reducing charge generation efficiency. Previous attempts to
Figure 39 illustrates the voltage loss pathways found in an achieving this goal using blends of donor polymer and fullerene
inorganic (a) and an organic (b) solar cell. Based on detailed acceptors have generally been unsuccessful, with the gain in
balance theory,87,297−299 the overall energy loss for any solar PCE due to increased VOC compromised by a large reduction in
cell (both inorganic and organic) can be attributed to the three photocurrent.71 However, recent studies have reported OPV
terms shown in eq 2. blends of novel donor polymers and nonfullerene acceptors
that exhibit highly efficient charge generation efficiencies
despite having small donor−acceptor energy offset. Examples
of these novel systems include P3TEA:SF-PDI2,87 PTB7-
Th:IEICO-4F,19 and PvBDTTAZ:O-IDTBR.88 These blends
have steep absorption edges as typically found in inorganic
cells, and, as a result, the voltage loss corresponding to CTS
absorption in these blends is much smaller than that observed
in typical high-performance OPVs.
The third term (qΔVnon−rad
OC = −kT ln (ηEL)) is due to
nonradiative recombination, where k is the Boltzmann constant,
T is the temperature and ηEL is the electroluminescence (EL)
quantum efficiency of the diode when charge carriers are
injected into the device in the dark.298,302 Inorganic solar cells
typically have ΔVnon−rad
OC in the range of 0.25 V, while this value
is typically larger for OPVs (0.38−0.44 V).303 The large
difference is due to the very low ηEL of bulk-heterojunction
Figure 39. Energy loss in inorganic and organic solar cells. Adapted blends (typically ∼10−6−10−8) compared to that typically
with permission from ref 87. Copyright 2016 Nature Publishing found in silicon or perovskite cells (typically ∼10−2−10−3). It is
Group. thus clear that increasing the ηEL of OPVs can lead to lower
3485 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

voltage losses, with 1 order of magnitude increase in ηEL recombination has been studied theoretically based on
corresponding to 60 mV reduction in ΔVnon−radOC . This is indeed calculations309−311 and experimentally with the use of
the case in the model high-efficiency, low-voltage loss OPV controlled donor and/or acceptor orientations in planar
blend, P3TEA:SF-PDI2,87 where a high ηEL of ∼10−4 is heterojunction devices.312−315 In particular, the recent study
achieved. Unlike typical OPV blends based on fullerene by Ran et al.314 shows that a face-on molecular geometry of the
acceptors, where the electroluminescence that arises from donor at the interface reduces nonradiative recombination
CTS emission is clearly red-shifted relative to the donor/ (increase VOC), while an edge-on orientation favors charge
acceptor singlet emission, the EL spectrum of the P3TEA:SF- generation (increase current).
PDI2 blend resembles that of the pure polymer (P3TEA) In summary, the large voltage loss in OPVs is due to two
device. This indicates that the EL emission of P3TEA:SF-PDI2 main factors. One is the existence of a significant offset between
is mainly from the singlet excitons in the polymer, which are the bandgap of the donor/acceptor materials and the energy of
most likely populated following electron transfer from the CTS the CT state. The other is the relatively large nonradiative
enabled by the small donor−acceptor energy offset.304 This recombination loss in OSCs, evidenced by low electro-
finding indicates that reducing the energetic offset at the luminescence quantum efficiency (ηEL) of OPV blends.
interface can in principle improve VOC by reducing both Identifying and optimizing the microscopic details that governs
ΔVrad,belowgap
OC and ΔVnon−rad
OC and is a promising way to improve these loss pathways would facilitate the development of highly
OPV efficiency provided that efficient charge generation can be efficient OPVs with small voltage losses comparable to those
maintained. typically found in inorganic devices.
Such systems inspire the OPV community to reconsider
whether the widely quoted exciton binding energy (∼0.3 eV)
12. CHALLENGES AND PERSPECTIVES
was based on a sound theoretical foundation. In general, this
value was estimated by calculating the Columbic attraction of 12.1. Perspectives on Molecular Design Strategy and
an electron and a hole assuming a close contact (e.g., 1 nm) Synthetic Complexity
between the two charges.305 However, recent studies have
shown that the positive charge (i.e., a hole) on the chain of a Thanks to the diversified material design strategy, over 100
polymer might have a large initial excitation volume NFAs have been reported for efficient OPV devices and the
approaching 20 nm3.306 Consider a polymer as a cylinder, PCE of single-junction nonfullerene OPVs has boosted from
assuming a 0.8 nm radius for the cylinder, the delocalization ∼3% to >13% within ∼5 years. In addition to concerning the
distance of the exciton on the polymer chain could be as large energetics and optical properties, due to the requirement for an
as 9 nm. Another study measured the electro-absorption signal interpenetrating bicontinuous network, the chemists need to
in a fullerene system using transient absorption spectroscopy make balanced assessment on the solubility, aggregation
and estimated the electron−hole distance to be ∼4 nm, which tendency, and molecular miscibility with the donor as well as
would significantly reduce the exciton binding energy to <0.1 the potential carrier mobility when designing new NFAs.
eV.83 Thereafter, the exciton binding energy might be smaller Developing molecules to satisfy these demands would be the
than 0.3 eV considering that the effective electron−hole pair essential goal. Whereas, at the current stage, each material class
distance is significantly larger than that assumed previously. demonstrates merits fulfilling some of these criterions but
This analysis should apply to both fullerene and nonfullerene shows limitations in the others. Future molecular design should
solar cells, and it does not explain why NFA seems to be more have balanced attentions paid on energetics (influencing charge
successful in achieving highly efficient systems with a negligible separation and VOC, see section 11), light absorption, free
driving force. Nevertheless, the fundamental mechanism of carrier generation (morphology), charge transport, and
efficient charge generation in an organic blend without synthetic flexibility.
requiring a driving force remains unclear, and the origin of 12.1.1. Molecular Design to Enhance Light Absorp-
nonradiative recombination in OPVs on a molecular level at the tion. A highly complementary absorption between the donor
interface is far from understood. and acceptor is the prerequisite of a broad and strong EQE
It is thus necessary to identify the microscopic details of OPV spectrum and is one of the most crucial reasons for the success
blends that govern voltage loss. The analysis above based on of NFAs. PDI-based NFAs, in general, exhibit absorption
detailed balance does not account for the microscopic coefficients of ∼105 M−1 cm−1 in solution and ∼104 cm−1 in a
parameters associated with OPV blends. Examples of these thin film while typical high performance IDT-based NFAs show
parameters include the degree of donor−acceptor mixing and absorption coefficients of ∼105 cm−1 in a thin film. Compared
the interfacial energetic disorder, which is expected to be more to the A-D-A type NFAs whose absorption profiles are more
significant in organic materials compared to inorganics. Several readily to tune, PDI molecules show relatively fixed absorption
studies have applied more complex and realistic models for the in the range 500−650 nm. Consequently, design strategies to
density of interfacial CTS and free carriers in OPV enhance extinction coefficients and alternating the spectral
blends.85,86,307,308 In particular, the recent work by Burke et range are strongly desired for PDI acceptors. As demonstrated
al. relates the low VOC of OPVs to a combination of their high in section 4.1.5, multiple cases have shown that the ring-fusion
degree of mixing, short CTS lifetimes, large amounts of strategy can be used to significantly increase the absorption
interfacial energetic disorder, and low dielectric constants coefficients and therefore serves as the top candidate for solving
leading to high CTS binding energies.86 By optimizing these this issue. Modifying the molecular geometry through
microscopic parameters of the interface between donor and connecting multiple PDI units together can also lead to
acceptor materials through molecular design, the performance changes in both absorptivity as well as absorption range.
of OPVs can be systematically improved through increases in Furthermore, the higher homologue of PDI such as TDIs can
VOC. Furthermore, the impact of molecular orientation at the be considered as they exhibit higher absorption coefficients and
interface on charge generation efficiency and nonradiative more red-shifted absorption spectra than PDIs.
3486 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

In contrast, the absorption profile of A-D-A type NFAs is intermolecular charge transfer while maintaining adequate
easier to tailor due to the facile synthesis. The electron solubility and miscibility are strongly needed. Historically,
donating and accepting building blocks as well as the spacer sacrificing aggregation for solubility/miscibility results in many
units can be modified separately, which creates a large number amorphous PDI molecules, which can be employed as
of combinations and makes it feasible to realize strong structural templates for modifications to occur. Approaches
absorptions even in the NIR region. This also opens up such as ring-fusion and ortho-position functionalization have
numerous opportunities for the design of semitransparent been proved effective in enhancing the overly reduced planarity
OPVs with a wide range of color options. To make further and should be further explored on more structures. Novel
improvement of the optical property, future molecular design molecular shapes such as 3D geometries, “double-decker”
could take advantage of the established methods such as tuning configurations,165 or “slip-stacked” structures127 can also be
the intramolecular push−pull character, elongating/expanding further investigated.
the conjugation, introducing halogenation to the “A” moiety, On the other hand, high performance A-D-A type NFAs
etc. typically exhibit an electron mobility in the range of 10−4 to
Besides their tunable absorption in the long-wavelength 10−3 cm2 V−1 s−1. In addition, a large fraction of IDT-based
region, the presence of a wide range of efficient large bandgap NFAs show a preferentially face-on orientation in the blend
polymers is an important reason for the rapid progress of the A- film, which is a desirable property for charge transport in OPV
D-A type NFAs. In contrast, there are less narrow-bandgap applications. Since the “D”, “A”, and spacer units can be
polymer donors with competitive performances. Therefore, to functionalized independently, the currently established method-
further utilize NFAs such as PDIs, effort should also be made ologies can still be employed to realize more combination of
on developing compatible donors, particularly those with these moieties. Extending the central conjugation length/area
absorption in the NIR region. seems to be a useful approach to improve the optical and
Regarding light absorption, building blocks such as quinoidal electronic property. Side chain engineering is another effective
thiophene derivatives and DPP deserve further exploration as method to tune intermolecular interaction. Replacing the
they can be used to construct chromophores with strong alkylphenyl side chain with an alkylthienyl one or changing the
absorptions that have direct improvement over the fullerenes. grafting site of the side chain have shown positive effects on
Further structural modification should focus on reducing the device performance. Furthermore, in addition to the well-
high-lying LUMO levels of DPP-based NFAs, which could known fluorination method, using a fused thiophene ring to
make them compatible with more donor materials. substitute the phenyl ring can also be applied to tune the
12.1.2. Molecular Design to Balance Morphology and electron withdrawing ability and electron mobility when
Charge Transport. The requirement for both a reasonably designing the “A” unit.
small domain size for efficient exciton dissociation and a Finally, considering that the PDI-based and IDT-based NFAs
bicontinuous network for long-range charge transport increases showed <3% PCE only a few years ago, new NFAs with other
the complexity of the material design. In the NFA-based active core units should not be ignored. When properly designed,
layers, the small domain size is typically achieved through a fused-ring aromatics, linear oligomers, and twisted dimers
decent miscibility between the donor and the NFA while the comprising other building blocks may reveal a great potential to
bicontinuous charge-transport network is facilitated by an compete with the current generation NFAs.
enhanced intermolecular order (or sometimes referred to as the 12.1.3. Synthetic Complexity. As the potential alternative
crystallinity in most literature). As discussed in section 4.1.4, to fullerenes, NFAs may be able to overcome the high synthetic
finding the optimal balance between miscibility and crystallinity complexity of fullerenes, which is important for the future large-
is not straightforward. Using a highly crystalline material with a scale production and commercialization. The high-performance
high OTFT or pure-film SCLC mobility may result in a largely PDI molecule SdiPBI-Se (see Table 1, molecule 4.9) was
phase-separated morphology in the BHJ blend. As a result, synthesized through a simple one-pot procedure with a high
crystallinity is sometimes exchanged for miscibility to harvest yield from a simple precursor 1-nitroperylene bisimide.135
more excitons during material design. For instance, introducing Moreover, the PDI acceptor TPH-Se (molecule 4.37) that
molecular twisting to enhance solution processability and exhibited a 9.3% PCE was also prepared through a one-pot
reduce the domain size is widely employed to construct PDI synthesis using a readily available precursor with an excellent
acceptors. However, controlling the extent of twisting is yield (87%).162 Regarding the production of fullerenes, Anctil
challenging. An excessive amount of molecular twisting may et al. have systematically investigated the energy and environ-
result in molecules that are too amorphous to form pure mental impacts through quantifying the life cycle embodied
domains with sufficient intermolecular stacking for efficient energy of both C60 and C70 fullerenes and their derivatives.316 It
long-range charge transport. In these cases, molecular design was shown that each step in the purification and modification of
approaches to exchange miscibility for crystallinity is needed. fullerenes increases both energy inputs and material losses. It
Despite the facts that techniques such as thermal annealing and was concluded that the embodied energy of all fullerenes are
the solvent additive approach can be employed to tune the notably higher than most common chemicals, which will have a
morphology during or after film formation, a proper pair of strong impact on the embodied energy of the final product.
donor and acceptor materials with appropriate molecular This is one of the most fundamental motivations to replace
miscibility and crystallinity are strongly desired. fullerenes with NFAs. While the synthetic complexity for
The electron mobilities of the PDI acceptors are typically in polymer and molecular donors has been analyzed,317,318 similar
the range of 10−5 to ∼10−4 cm2 V−1 s−1, which are able to studies on NFAs have not been reported.
enable a relatively high FF (mostly in the range of 50−65%). Despite the fact that a comprehensive analysis of all of the
However, further enhancement is essential to match the high NFAs is outside the scope of this review, here we briefly discuss
hole mobility of the state-of-the-art donor materials. To this the synthesis of one of the best-performing PDI derivatives
end, juditious molecular design methods that can enhance (TPH-Se) and the benchmark A-D-A type NFA, ITIC. Starting
3487 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

from perylene-3,4,9,10-tetracarboxylic dianhydride, 6-undeca- the macroscopic mobility mentioned above. Time-resolved
none, and 1,3,5-tribromobenzene, TPH-Se can be synthesized microwave conductivity (TRMC) is a technique for this
in eight steps with a total yield of ∼48%.162,319−321 In contrast, purpose with a nanosecond time resolution.334 TRMC has
the synthesis of ITIC from 3-bromothiophene, 1,4-dibromo- been carried out to compare the local carrier mobility and
2,5-dimethylbenzene, and 1,3-indandione requires 11 steps with charge photogeneration between fullerene and nonfullerene
a total yield of <7%.322−327 For the best-performing A-D-A type systems.43,101 For instance, TRMC studies have been
NFA, IT-4F, the fluorination of the IC group further increases performed on PSEHTT:DBFI-T and PSEHTT:BFI-P2 non-
the synthetic complexity.194 It is clear that currently rylene fullerene systems, in comparison to a PCBM-based control
derivatives are better than A-D-A type NFAs in terms of sample.224 The authors showed that the dependence of
synthetic complexity. However, previous reports of A-D-A type transient photoconductivity maxima for the PSEHTT:DBFI-T
NFAs may have paid more attention to assessing the target blend is significantly higher than that for the fullerene or BFI-
molecule instead of optimizing the synthetic steps and P2 based blend, consistent with their photovoltaic performance.
improving the yields of each step. Therefore, further study on In addition, by varying the excitation wavelength, the position
chemistry will be helpful for not only reducing the synthetic of the peaks could be used to reveal the interplay of charge
complexity of existing NFAs but also developing novel NFAs generation pathways from excitons in the donor and acceptor
with improved synthetic schemes. phase.
A good example is the recent report of the synthesis of a 12.3. Understanding Morphology of Nonfullerene Active
PDI-based NFA without column chromatography.136 Other Layers
preliminary work showing simplified material synthesis for
individual NFAs is also available.200,328,329 These results The active layer morphology of BHJ type solar cells requires a
indicate that there is a tremendous room to improve in terms bicontinuous interpenetrating network for balancing exciton
of the synthesis of NFA materials. In addition, as some NFAs harvest and charge transport. Morphology formation is a
can only achieve their best performance with specific donor complex process influenced by numerous factors including the
polymers, the synthetic complexity of the corresponding donor intrinsic material properties such as molecular weight,335−338
polymer should also be considered. For example, the donor regioregularity,335,339−344 purity, miscibility between donor and
polymer PDBT-T1, which was used together with TPH-Se, acceptors,345−352 the effect of solvents and additives,353−356 film
contains a monomer dithienobenzodithiophene with a high processing conditions, and post film formation treat-
synthetic complexity based on a previous analysis.317 Therefore, ments.281,335,357−362 Both thermodynamics and kinetics play
analysis of the active layer as a synergistic whole instead of each critical roles in determining the final morphology under which
material will be more accurate and instructive. the device is established. Accumulated understanding of
morphology is accompanied by the emergence of novel
12.2. Studying Charge Transport materials and device fabrication techniques. The morphology
Electron mobility is one of the most important parameters to of polymer:fullerene BHJs has been investigated for decades
evaluate the overall potential of an acceptor material. Unlike and is still one of the most debatable topics. The current state
PCBM, whose electron mobility is typically comparable to or of understanding of polymer:fullerene BHJ morphology
higher than the hole mobility of polymers in the blend, many involves a multiphase model where ordered crystallites,
NFAs show lower mobility than polymers, which is one of the relatively pure donor and acceptor aggregates, and molecularly
key reasons limiting the FF of the OPV device. A subtle mixed amorphous domains coexist in the blend film.363 The
modification of the chemical structure can result in a drastic delegated roles of these phases on light absorption, charge
change in the stacking behavior and thus the electron mobility generation, and charge transport are still under debate.364
of the NFA. The nanoscale network of NFAs in thin film is 12.3.1. Quantitative Morphology-Performance Corre-
dramatically different among different backbones, let alone lations. The aggregation properties of NFAs and the
between NFAs and fullerenes. miscibility between NFAs and donor materials can be largely
To date, most reported mobilities for NFAs are obtained different from fullerenes. For instance, DPP-based polymers
from fitting space-charge-limited current (SCLC) models to show inferior miscibility with ITIC and the microsized phase
single-carrier diodes. SCLC model is useful for evaluating separation make the BHJ device perform much worse than their
macroscopic, device length scale mobilities for injected carriers. fullerene-based counterpart does.365 At this point in time, most
To experimentally probe the mobility of photogenerated morphological studies on nonfullerene OPVs are specific to the
carriers, others techniques such as time-of-flight and photo- presented case. Quantitative and systematic investigation of film
induced charge extraction by linearly increasing voltage (photo- morphology with a focus on correlating them with measured
CELIV) could be employed.25,27,90,177,267,330,331 Besides photo- device parameters are strongly needed. In addition, polymer:-
generated carriers, photo-CELIV is also the only technique that NFA blend films are different from polymer:fullerene films in
can be performed on actual OPV devices.332 terms of material contrasts, which increases the complexity to
Nevertheless, these methods measure the motion of carriers extract meaningful information from some widely used imaging
at the device length scale, e.g., tens to hundreds of nanometers. techniques. This further elevated the difficulty to evaluate
Previous work has shown that the local mobility in a shorter- nonfullerene-based active layer morphology at length scales
ranged molecular length scale could be largely different from relevant for BHJ. In spite of this, insightful results have been
the mobilities in the macroscopic length scale, which can be shown in some recent articles by utilizing soft and/or hard X-
limited by grain boundaries.333 Since the local mobility is more ray scattering techniques to characterize the nanoscale
sensitive to the probability for an electron to hop from one morphology of some highly efficient nonfullerene solar cells.
NFA molecule to another, it could be used to correlate with the For example, through the use of resonant soft X-ray scattering
nanoscale order of the NFAs revealed by morphological (R-SoXS), Ye et al. were able to establish a strong correlation
characterizations, which could serve as a great supplement to between the average compositional variation, or sometime
3488 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

referred to as domain purity, at a length scale of ∼10 nm, and performance is more strongly influenced by the π−π stacking of
the FF and JSC of solar cell devices based on high-efficiency the NFA rather than that of the polymer suggests the charge
polymer:NFA blends, namely PBDB-T:IT-M and PBDB-T:IT- transport of these state-of-the-art IDT-based NFAs is rather
DM with different fabrication processes (Figure 40).366 different from both polymers and fullerenes. It was established
that the most efficient charge transport is along the backbone,
followed by the π−π stacking direction and then the lamellar
stacking direction. Due to the fact that polymers in general have
much longer conjugated backbones than the small molecular
NFAs, although still important, the impact of the (010)
coherence length for polymers on the overall hole conduction is
less than the impact of (010) coherence length for the small
molecules. These conclusions should not be limited to IDT-
based A-D-A type NFAs. As the molecular twisting and
coplanarity paradigm for the rylene-based NFAs is more
significant, it would be expected that similar studies been
carried out onto more polymer:NFA combinations.
Since IDT-based A-D-A type small molecules represent the
state-of-the-art NFAs, in a broader context, these results
indicate that at the moment, in a high-performance polymer:-
NFA blend, the nanoscale morphology and local conductivity
of NFAs are the limiting factors that govern the overall device
performance. Nevertheless, the high FFs achieved in some of
the NFA-based BHJ blends prove that these issues can be
overcome through judicious molecular design and morphology
optimization.
In addition to understanding the solid state morphology after
the solid film is formed, grasping the morphology formation
process from an in situ point of view is also critical because the
formation of the solid state BHJ usually involves the use of
solvent mixtures, solvent additive, postfilm-formation-treat-
ments (thermal annealing, solvent vapor annealing, etc.), or
temperature dependent film drying.364 Building a thorough
understanding of morphology forming processes arising from
either liquid−liquid demixing or crystallization is necessary to
develop structure−property-processing-performance relation-
Figure 40. (a) Plot of FF of the PBDB-T:IT-DM nonfullerene devices
versus the respective composition variation of the high-q peak (Δρ)
ships.
obtained from R-SoXS. Reprinted with permission from ref 366. 12.3.2. Molecular Miscibility. Despite the fact that kinetic
Copyright 2017 John Wiley and Sons. (b) Plot of (010) coherence controls during film formation can make the active layer dry at
length of SMA versus the FF for a series of polymer:IDT-based NFA a thermodynamically instable state, understanding the mixing
blends. Reprinted with permission from ref 367. Copyright 2017 John capabilities of semiconducting polymer blends is essential for
Wiley and Sons. predicting the thermodynamically stable state of the film which
is related to the morphological stability. To this end, Flory−
Hu et al. combined GIWAXS and R-SoXS to characterize a Huggins368 or Flory−Rehner369 theory have been utilized by
series of polymer:NFA blends.367 To better unravel the role of various groups to evaluate the miscibility between different
NFA in the high-performance polymer:NFA blends, the components in polymer-based blends.347,351,352,366,370−378 Most
archetypal NFA, ITIC-Th, was chosen as the single NFA and of these studies focused on the polymer and fullerene
a series of polymers were compared. It was found that, when miscibility, but the preliminary result indicates the potential
combined with different polymers, ITIC-Th demonstrated to use the Flory−Huggins interaction parameter and the
different propensity to aggregate/crystallize. Further analysis calculated spinodal demixing curve to explain certain
revealed that the photovoltaic devices’ FF and JSC increased morphology result such as average compositional variation
monotonically with increasing coherence length of the (010) (Figure 41). Nevertheless, robust theoretical models and
peak of the NFA (Figure 40), calculated from GIWAXS experimental techniques to evaluate the miscibility and assess
experiments. Intriguingly, such correlation could not be built phase separation from the thermodynamic viewpoint is needed.
with the (010) peak coherence lengths of the polymer. 12.3.3. Solvent Additive Effect. The use of solvent
These results have two main indications. First, different additives has been considered one of the most effective
polymers interact with these IDT-based NFAs differently, the methods to tune the morphology of the active layer of BHJ
results of which were clearly reflected by the different OPVs.379−381 Most of the high PCEs for polymer:fullerene
aggregation of NFAs in the film. The aggregation of the devices were realized through the implantation of solvent
NFAs may or may not affect the aggregation/crystallization of additive(s).355 High boiling point solvents such as DIO, 1,8-
the polymer, judging from the data alone, although the average octanedithiol (ODT), and CN are the most widely used
compositional variation seemed to correlate better with the additives, which have been employed in both fullerene355,379,382
NFAs’ (010) coherence lengths. Second, the fact that the device and nonfullerene-based blends (Tables 1−3).171,241
3489 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

toward a more ideal BHJ nanostructure than that the original


solvent can provide.
Compared to fullerene-based OPVs, nonfullerene donor:-
acceptor blends, in general, demonstrate less demand for
solvent additives. A possible explanation could be that the
solubility selectivity between the donor polymer and the NFA
is reduced in comparison to that between the polymer and
fullerene. For instance, Song et al. compared the effect of DIO
on film morphology and device performance in both PC71BM
and ITIC based blends with the same polymer donor, PTB7-
Th.388 It was concluded that, unlike the fullerene-based blend
where DIO increased the absorption and aggregation of
PC71BM (and thus the FF and JSC), the PTB7-Th:ITIC
blend showed slightly inferior morphological, optical, and
electrical properties after adding DIO into the main solvent
(CB).
Nevertheless, the requirement for solvent additive including
the choice of the additive and/or its volume fraction in the
solvent mixture varies among different material systems. For
example, the nonfullerene blends based on PDI molecules
demonstrate, on average, a stronger requirement for solvent
additive (Table 1) than the nonfullerene blend based on A-D-A
type NFAs (Table 2). The main reason could be that, although
different functionalities and/or molecular geometries have been
Figure 41. (a) Calculated χ parameters at room temperature from introduced, most PDI acceptors are still not able to enable a
Hansen solubility parameters and inferred χ at the annealing favorable morphology with highly crystalline, reasonable small
temperature (160 °C) in relation to spinodal and binodal curves of yet pure domains, which, consequently, requires additional
the polymer:SMA systems for a given weight ratio of 1:1 (the dashed
lines indicate the composition in the amorphous mixed domain in the efforts such as the assist of a solvent additive to address. In
limit of complete phase separation at the respective χ and the gray contrast, several high-performance A-D-A type NFAs are
horizontal line represents the critical χc ≈ 0.8). (b) Plot of capable of forming a near optimal morphology with the
composition variation of the high-q peak (Δρ) under various donor due to a balanced molecular miscibility and crystallinity,
processing conditions in relation to the room temperature and which therefore releases the requirement for solvent additives.
Hansen solubility parameters derived χ parameter of the nonfullerene It is noteworthy that this is only a general discussion, and there
OSCs. Reprinted with permission from ref 366. Copyright 2017 John are certainly cases where an A-D-A type NFA requires a solvent
Wiley and Sons. additive or a donor:PDI blend does not rely on any additive to
achieve a high device performance.
Despite the fact that solvent additives are beneficial for the
The role of a solvent additive in tuning the morphology is
realization of a high PCE, they should better be precluded for
complicated, and the explanation varies from case to case.
eventual industry-scale production of the OPV devices as they
Generally, the main functions of solvent additives are tuning
may induce device instability and irreproducibility.389 There-
the degree of phase separation and/or changing the crystallinity
of one or both components. In most fullerene-based systems, fore, future material design for the active layer should focus not
solvent additives have been considered a selective solvent that only on the optical property and energetics but also on
exhibits differential solubility to the acceptor and donor in the optimizing miscibility as well as crystallinity of the materials to
blend solution.8,10,383,384 Particularly, it is widely acknowledged achieve additive-free processing.
that DIO, the solvent additive that is less volatile than common 12.3.4. Vertical Phase Segregation. Besides morphology
organic solvents such as CB and CF, exhibits selective solubility in the bulk, the distribution of organic materials in the direction
to PC71BM, which promotes the aggregation of fullerenes and perpendicular to the substrate, also known as the vertical phase
allows the fullerene to achieve an overall higher crystallinity segregation, is of equal importance because the transport of
(more crystalline domains with comparable crystal size) when charges in the vertical direction results in charge extrac-
combined with the workhorse material PTB7.355 Akin to the tion.348,362,390−408 It is accepted by the field that relative surface
role of DIO, in the NREL certified 11.5% polymer:fullerene tension can be used to predict the wetting propensity of a
OPV system,10 the high-boiling point solvent additive 1- component in the blend at a certain interface, e.g., the
phenylnaphthalene (PN) has been reported to play multiple component with higher surface tension tends to segregate
synergistic effects including (i) enhancing the electron mobility, toward the interface with a high surface tension.390,394,409
(ii) enhancing domain purity while reducing domain size, and PCBM, in general, has a higher surface tension than lots of
(iii) promoting a face-on orientation of the polymer. polymers reported and is therefore more likely to segregate to
Alternatively, solvents such as CN can be a cosoluble the bottom (substrate). However, surface tension of novel
additive, which has been found to be able to suppress the NFAs may differ strongly from fullerenes or polymers.
aggregation of polymer(s) and prevent the formation of Therefore, depending on the relative surface tension between
undesirably large domains.385,386 In addition, solvent additives the NFA and the polymer, appropriate device architectures, e.g.,
can modify the vertical phase segregation.371,387 Nevertheless, normal vs inverted, needs be selected to avoid intense surface
the goal of a solvent additive is to alternate the morphology recombination. Alternatively, solvent additives and post-treat-
3490 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 42. (a, c) Molecular structures of electron donor and electron acceptor materials in the (a) front and (c) rear cells. (b) Device architecture of
the tandem OSCs. (d) Molecular structure of the interconnection layer, PCP-Na, in the tandem cells. (e) Transmittance spectra of ZnO/PCP-Na
ICLs with different PCP-Na thicknesses. (f) Schematic energy level diagram of each material used in this study. Adapted with permission from ref 20.
Copyright (2017) American Chemical Society.

ments such as solvent vapor annealing are needed to tune the optimal active layer thicknesses, i.e., about or less than 100 nm,
vertical phase segregation for a given device architecture. as a result of the limited electron mobility of the NFAs. The
12.4. Multijunction NFA OPVs thin active layer may not have sufficient absorption for
achieving high JSC. Therefore, fabricating homotandem device,
Multijunction OPVs have been previously explored in the where the top and bottom subcells consist of the same
polymer:fullerene BHJ solar cells with the majority effort materials sets, can enhance light trapping and make further use
focused on the series connected tandem structure,410−416 where of the photons penetrated through the front cell. On the other
the bottom and top subcells are separated by a recombination hand, reduced energy loss provides multiple nonfullerene single
medium. Apparently, the total VOC of the series connection junction devices with high VOC while maintaining decent JSC
tandem device should theoretically equal the sum of the VOCs and FF. Hence, stacking two such subcells in a series
of the two subcells. The main motivation for this device connection fashion can afford the tandem device an exception-
platform is to increase the absorption breadth by using two ally high VOC, which may allow the realization of certain
subcells with complementary absorption profiles. application requiring high-voltage operation, such as water
In general, nonfullerene tandem OPVs should inherit all splitting.418,419
advantages of the tandem structure and avoid the aforemen- Driven by the high VOC and the low active layer thickness in
tioned drawbacks inherent to fullerenes. In a recent publication the 9.5% single-junction PTFB-O:SF-PDI2 device,87 Chen et al.
by Cui et al., the authors employed nonfullerene blends fabricated homotandem devices and achieved a homotandem
consisting of materials with both large and narrow bandgaps to nonfullerene solar cell with a PCE of 10.8% and an
broaden the absorption spectrum (Figure 42). Through the use extraordinarily high VOC of >2.1 V.420 This result was obtained
of a recombination layer with extremely high transmittance that through constructing a novel all-solution processed recombi-
minimizes parasitic absorption, the authors demonstrated a nation layer that did not require harsh post-treatment and
PCE of above 13.0% for the tandem device, which showed carefully designing the two active layers’ thicknesses under the
dramatic improvement over the individually optimized top and guidance of optical modeling. The high VOC has enabled
bottom subcells (9.0% and 10.1%). Another work by Liu et successful demonstration of photovoltaic-driven water splitting
al.417 also took advantage of the complementary absorption of (Figure 43).
the subcells. In combination with the high VOC of each cell, the Zuo et al. fabricated highly efficient homotandem devices
resultant tandem device achieved a PCE of 8.48% with a high based on PTB7-Th and ITIC or 4TIC.421 The tandem device
VOC of 1.97 V. was connected in parallel, which does not require strict
Besides these known benefits, some intrinsic properties of photocurrent matching to maximize the overall JSC. The use of
the NFA based single-junction OPVs may further stimulate the homotandem also guarantees voltage matching across the
construction of the tandem device. For example, a large fraction parallel-connected subcells. As a result, high JSCs of ∼17.92 and
of high-performance nonfullerene solar cells have relatively thin 20.81 mA cm−2 were demonstrated for the ITIC and 4TIC
3491 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

nation. The requirement of harsh treatment of most


recombination layers also limits the choices of active layers.
With the increasingly emerged recombination layers with
desired properties, the tandem strategy could be considered as
another promising option to further optimize device efficiency.
12.5. Device Structure Engineering and New Interfacial
Layers
In addition to the majority effort spent on active layer
optimization such as donor/acceptor designing and morphol-
ogy tuning, another critical aspect in improving the overall
device efficiency and stability is interfacial engineering. The
importance of choosing the proper interlayer cannot be
overstated as an inferior interface between active layer and
electrodes have manifold negative impacts on charge extraction
including (i) forming a nonohmic contact with the active layer,
Figure 43. Photo of photovoltaic-driven water splitting using the high- (ii) causing an unfavorable vertical phase segregation of the
voltage homotandem nonfullerene organic solar cell in ref 420. An
electrochemical cell was connected to the tandem solar cell illuminated
active layer, (iii) undesired doping of the active medium, (iv)
under AM 1.5G (100 mW cm−2), with the platinum wire and nickel diffusion into the active layer (v) increasing series resistance,
foam electrodes immersed in a 1 M NaOH electrolyte solution. (vi) causing parasitic absorption, and (vii) reducing device
Reprinted with permission from ref 420. Copyright 2017 John Wiley stability, etc.
and Sons. Multiple interlayers and/or electrodes have been shown to
interact with fullerenes.422 The influence of such interaction can
based tandem device, respectively. The improved JSCs enabled be either positive or negative. Many interlayers were inten-
high PCEs of 10.03% and 11.08%, respectively. tionally optimized for fullerene based OPV devices. Con-
In summary, the easily tunable optical properties of NFAs sequently, after the fullerenes are replaced with novel NFAs, the
make them suitable candidates for application in a multi- known benefits from previous interlayers may not exist and new
junction device. Considering the low energy loss offered by interfacial interaction may be introduced. Exploiting a
certain donor:NFA combinations, the tandem device could compatible interlayer is essential for further boost of PCE
achieve extraordinarily high VOC while maintaining decent and stability of the NFA-based devices.
current and FF. One of the reasons limiting the development of As a typical example, the recent publication by Sun et al.423
this device platform is the lack of effective recombination layers introduced a NDI based water-soluble conjugated polymer
with high photon transmittance and efficient charge recombi- (WSCP), namely PDI-2TNDI, as the electron transport layer

Figure 44. (a) Typical J−V characteristics of PBDB-T:ITIC devices with a structure of ITO/PEDOT:PSS/PBDB-T:ITIC/WSCP/Al under
simulated AM 1.5G irradiation (100 mW cm−2). The WSCP interlayer thickness is ∼5 nm. (b) Schematic illustration of the exciton separation and
transport in BHJ and the charge transfer states at the donor/WSCP interface. (c−f) Energy level diagrams of PTB7-Th:N2200 BHJ with a (c) thin
or (d and e) thick interface layer and (f) PTB7-Th:IDT-2BR BHJ with a thick PFN-2TNDI interface layer. Reproduced from ref 423 with
permission. Copyright 2017 Royal Society of Chemistry.

3492 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Figure 45. (a) PCE of P3HT:IDT-2BR based BHJ solar cells as a function of active layer thickness. Reproduced from ref 184 with permission.
Copyright 2015 Royal Society of Chemistry. (b) PCE versus active layer thickness of the solar cells based on PTZ1:IDIC. (c) J−V characteristics of
the optimized solar cells based on PTZ1:IDIC with different active area. (b, c) Reprinted with permission from ref 426. Copyright 2017 John Wiley
and Sons. (d) Device parameters (PCE, JSC, VOC, and FF) versus the active layer (PBDB-T-SF:IT-4F) thickness. Adapted with permission from ref
19. Copyright (2017) American Chemical Society.

between the active layer and the cathode. By comparing the for large-scale production. Nevertheless, the emergence of
PFN-2TNDI with another interlayer (PFN), the authors recent high mobility NFAs have alleviated this problem. For
delineated the difference of interfacial doping between a example, in ref 184, the device performance of P3HT:IDT-2BR
fullerene-based BHJ and a nonfullerene based BHJ (Figure 44). was shown to be relatively insensitive to the active layer
A clear interfacial n-doping was observed at the fullerene/ thickness in the range of 60−220 nm (Figure 45a), which
WSCP while the doping at the NFA/WSCP interface was not further confirmed the efficient charge transport enabled by the
prominent. Therefore, in terms of reducing contact resistance high electron mobility of the acceptor. This was the first report
and forming an ohmic contact with the NFA based active layer, that nonfullerene OSCs could perform well in a relatively thick
the self-doped PFN-2TNDI outperformed PFN. Furthermore, (over 200 nm) active layer.
the authors revealed that PFN-2TNDI could serve the role of The PCE of the PTZ1:IDIC nonfullerene solar cell in ref 426
dissociating the exciton of the polymer donor at the interface. A retains a high value (>9%) at an active layer thickness of 210
pronounced PCE value of 11.1% was achieved in nonfullerene nm with a decent FF (66.8%; Figure 45b). It is worth noting
solar cells. that a 10.5% efficient solar cell at an active device area of 81
This research clearly demonstrates the importance of mm2 was demonstrated for PTZ1:IDIC (Figure 45c). The PCE
tailoring interfacial properties among different material systems. of the PBDB-T-SF:IT-4F based device in ref 19 also
To better harvest the potential of the novel nonfullerene active demonstrated a weak dependence on the active layer thickness:
medium, further understanding of the role of different a PCE >12% was obtained for active layers with thickness up to
interlayers is strongly demanded. 200 nm and a PCE >10% was achieved using an 300 nm-thick
active layer (Figure 45d). The authors also obtained an 11.1%
12.6. Thick Active Layer, Large-Area Devices,
efficient device at a device area of 1.00 cm2 in this work.
Environmentally Friendly Processing, and Roll-to-Roll
Similarly, a thick active layer and efficient nonfullerene solar
Printing
cells were achieved using NFAs including IDT-2BR,184 m-
The thickness of the active layer of an organic solar cell has a ITIC,191 IT-OM-2,427 di-PBI,133 ITIC,428 etc.
strong impact on its potential for industrial-scale production. In The active layers of most efficient OPV devices are processed
general, the thickness of the active layer of an polymer:fullerene using chlorinated solvents such as chlorobenzene, o-dichlor-
based organic solar cell is from tens to hundreds of obenzene, or chloroform, which are not environmentally
nanometers.295,424,425 In comparison, most of the nonfullerene friendly solvents. However, previous work have already
BHJ organic solar cells reported contain an active layer with shown that eco-friendlier solvents can be employed to process
thickness around or below 100 nm due to the relatively low the organic layer based on polymer:fullerene blends.10
electron mobility of the NFA, which limits photon harvesting Similarly, nonchlorinated, environmentally friendlier solvents
and could affect the overall robustness of the device, particularly have also been utilized to process nonfullerene solar cells. A few
3493 DOI: 10.1021/acs.chemrev.7b00535
Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

typical examples include the use of anisole to make a 5.43% all- blocks and molecular design strategies, along with convergent
polymer solar cell,429 the use of a toluene/diphenylether and low-cost synthetic routes have been employed in practice
mixture to fabricate a 9.73% polymer:NFA solar cell,32 the use to build over 100 NFAs with tunable spectroscopical and
of a mixture of THF and IPA to fabricate a highly efficient electrochemical properties. NFAs have been employed to
(11.34%) polymer:NFA solar cell,430 and the use of mesitylene combine with either polymer or molecular donors in a variety
as the active layer solvents to achieve a 11.1% (higher than that OPV device platforms including binary/ternary single-junction
processed from chlorobenzene) polymer:NFA solar cell with and multijunction solar cells, which demonstrated not only
long-term device stability.28 record power conversion efficiencies but also improved
One of the unique features of organic solar cells is the chemical, thermal, and photostability. At the current PCE-
capability of being roll-to-roll (R2R) printed into large area, advancing pace, efficiencies over 15% can be forecasted for
flexible devices. Although most R2R based mass production of single-junction nonfullerene OPVs in the near future. More
OPVs were based on fullerene acceptors,431−436 nonfullerene excitingly, environmentally friendly solvents, room-temperature
roll-coated devices have also been realized.382,437,438 For processing, large area, and thick-active layer have been exercised
instance, Cheng et al. demonstrated roll-coated all polymer by various groups to fabricate NFA-based organic solar cells.
solar cells in 2014437 and Liu et. fabricated polymer:NFA based These accomplishments manifest the potential for industrial
nonfullerene solar cells using a slot-die coating and flexographic production and eventual commercialization of OPVs based on
printing methods.438 Recently, Liu et al. showed polymer:NFA this material class.
(PTB7-Th:IEIC) solar cells fabricated via a roll-coating process On the other hand, morphology and device physics studies
under ambient conditions.382 The authors compared both have shown that, at the current stage, improving the nanoscale
flexible ITO and ITO-free substrates (Figure 46), which organization and charge carrier conduction of the acceptors by
designing new or optimizing known NFAs to match with the
high-performance polymers, whose emergence were based on
the knowledge gained from the past decades, is still essential to
further enhance device performance. Empirical design rules can
be gained from analysis of the known material properties and
device parameters, which can stimulate the development of new
materials and/or discover better material combinations. Never-
theless, a better understanding of the physical processes
governing device operation, such as the morphology forming
process and charge generation and recombination mechanism,
is required to guide future research.

AUTHOR INFORMATION
Corresponding Author
*E-mail: hyan@ust.hk.
ORCID
Zonglong Zhu: 0000-0002-8285-9665
Jie Zhang: 0000-0001-8536-5436
Fei Huang: 0000-0001-9665-6642
He Yan: 0000-0003-1780-8308
Author Contributions

G.Z., J.Z., P.C.Y.C., and K.J. contributed equally to this work.
Notes
Figure 46. (a) PEDOT:PSS PH1000 and ZnO coated PET flexible
substrate with an Ag grid (ITO-free substrate), (b) slot-die coating of The authors declare no competing financial interest.
the active layer of PTB7-TH:IEIC and top PEDOT:PSS layer on the Biographies
mini roll-coater in sequence, (c) flexographic printing of the top Ag
electrode, and (d) long stripes of roll-coated nonfullerene OSCs based Guangye Zhang received his Ph.D. degree in 2015 from University of
on the ITO-free substrate. Reproduced from ref 437 with permission. California, Los Angeles for his study in the field of organic
Copyright 2015 Royal Society of Chemistry. photovoltaics under the direction of Professor Benjamin Schwartz.
He is currently a research assistant professor at HKUST. Besides
exhibited PCEs up to 2.26% and 1.79%, respectively, organic and hybrid solar cells, his research interests include
comparable to the reference devices with fullerene acceptors optoelectronic techniques such as transient photovoltage, charge
under the same conditions. Noticeably, the nonfullerene extraction, and photocurrent spectral response.
devices displayed better dark shelf stability than the fullerene- Jingbo Zhao received his B.Sc. degree in Peking University in 2012,
based devices. after which he joined Professor He Yan’s group in HKUST and
obtained his Ph.D. degree in 2015. He worked as a Postdoctoral
13. CONCLUDING REMARKS Fellow in the same group and then moved to City University of Hong
The examples presented in this review clearly demonstrate that Kong as a Research Fellow in Professor Alex K.-Y. Jen’s group. His
the nonfullerene acceptor is a promising class of materials to research interests include development of functional materials for
replace fullerenes in organic solar cells. Diversified building organic and perovskite solar cells, understanding the structure-

3494 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

performance relationship of organic semiconductors and fabrication of Technology Commission for the support through projects ITC-
flexible electronic devices. CNERC14SC01 and ITS/083/15.
Philip C. Y. Chow obtained his Ph.D. degree from the University of
REFERENCES
Cambridge in 2016 for his studies of charge dynamics in organic solar
cells. He is currently a Research Assistant Professor at HKUST with (1) Office of Science; U.S. Department of Energy. Basic Research
Needs for Solar Energy Utilization; 2005.
research interests in organic optoelectronic devices.
(2) U.S. Energy Information Administration. International Energy
Kui Jiang obtained his B.S. degree in Chemistry from Wuhan Outlook 2016; www.eia.gov/forecasts/ieo/pdf/0484(2016).pdf.
University with Professor Jingui Qin in 2011. After that, he moved to (3) Espinosa, N.; Hösel, M.; Angmo, D.; Krebs, F. C. Solar Cells with
the Hong Kong University of Science and Technology. He is currently One-Day Energy Payback for the Factories of the Future. Energy
a Ph.D. candidate under the direction of Professor He (Henry) Yan. Environ. Sci. 2012, 5, 5117−5132.
(4) Krebs, F. C.; Espinosa, N.; Hösel, M.; Søndergaard, R. R.;
His current research focuses on the devices fabrication for photovoltaic
Jørgensen, M. 25th Anniversary Article: Rise to Power - OPV-Based
solar cells. Solar Parks. Adv. Mater. 2014, 26, 29−39.
Jianquan Zhang received his B.S. degree in Polymer Science and (5) Lacal-Arantegui, R. A. J.-W.; Andrei Bocin-Dumitriu, B. S.; Zubi,
Engineering from Zhejiang University, People’s Republic of China, in G.; Magagna, D.; Carlsson, J.; Moss, R.; Fortes, M.; del, M. P.;
2015. He is currently a Ph.D. student under the supervision of Lazarou, S.; Baxter, D.; Scarlat, N.; et al. Technology Map of the
Professor He Yan. His interests lie in the synthesis of novel donor European Strategic Energy Technology Plan, 2013; http://setis.ec.
europa.eu/system/files/2013TechnologyMap.pdf.
polymers and small molecular acceptors for nonfullerene organic solar (6) Mazzio, K. A.; Luscombe, C. K. The Future of Organic
cells. Photovoltaics. Chem. Soc. Rev. 2015, 44, 78−90.
Zonglong Zhu received his Ph.D. degree in Hong Kong University of (7) Yu, G.; Gao, J.; Hummelen, J. C.; Wudl, F.; Heeger, A. J. Polymer
Science and Technology in 2015 and worked as a postdoc fellow in Photovoltaic Cells: Enhanced Efficiencies via a Network of Internal
Materials Science & Engineering Department at the University of Donor-Acceptor Heterojunctions. Science (Washington, DC, U. S.)
1995, 270, 1789−1791.
Washington until 2017. Currently, he works in Prof. Henry Yan’s
(8) Heeger, A. J. 25th Anniversary Article: Bulk Heterojunction Solar
group as a visiting scholar. His current research is in electronic devices Cells: Understanding the Mechanism of Operation. Adv. Mater. 2014,
based on inorganic/organic perovskites, especially for their applica- 26, 10−28.
tions in photovoltaics and light-emitting diodes. (9) Lu, L.; Zheng, T.; Wu, Q.; Schneider, A. M.; Zhao, D.; Yu, L.
Jie Zhang received his B.S. degree in Chemistry from Northeast Recent Advances in Bulk Heterojunction Polymer Solar Cells. Chem.
Rev. 2015, 115, 12666−12731.
Normal University in 2007 and gained his Ph.D. degree in Materials
(10) Zhao, J.; Li, Y.; Yang, G.; Jiang, K.; Lin, H.; Ade, H.; Ma, W.;
Science from the South China University of Technology in 2012 under Yan, H. Efficient Organic Solar Cells Processed from Hydrocarbon
the supervision of Prof. Yong Cao and Prof. Fei Huang. His research Solvents. Nat. Energy 2016, 1, 15027.
interests are conjugated functional materials for organic/polymer (11) Jin, Y.; Chen, Z.; Xiao, M.; Peng, J.; Fan, B.; Ying, L.; Zhang, G.;
optoelectronics. Jiang, X.-F.; Yin, Q.; Liang, Z.; et al. Thick Film Polymer Solar Cells
Based on Naphtho[1,2- c:5,6- c ]bis[1,2,5]thiadiazole Conjugated
Fei Huang received his B.S. degree in Chemistry from Peking
Polymers with Efficiency over 11%. Adv. Energy Mater. 2017, 7,
University in 2000 and gained his Ph.D. degree in Materials Science 1700944.
from the South China University of Technology in 2005 under the (12) Hummelen, J. C.; Knight, B. W.; Lepeq, F.; Wudl, F.; Yao, J.;
supervision of Prof. Yong Cao. After postdoctoral work at University Wilkins, C. L. Preparation and Characterization of Fulleroid and
of Washington with Prof. Alex K.-Y. Jen, he began his academic career Methanofullerene Derivatives. J. Org. Chem. 1995, 60, 532−538.
in 2009 as a professor of South China University of Technology. His (13) Wienk, M. M.; Kroon, J. M.; Verhees, W. J. H.; Knol, J.;
main interests are in the fields of organic functional materials and Hummelen, J. C.; van Hal, P. A.; Janssen, R. A. J. Efficient
devices for optoelectronics. Methano[70]fullerene/MDMO-PPV Bulk Heterojunction Photovol-
taic Cells. Angew. Chem., Int. Ed. 2003, 42, 3371−3375.
He (Henry) Yan received his B.Sc. degree from Peking University in (14) Sariciftci, N. S.; Smilowitz, L.; Heeger, A. J.; Wudl, F.
2000 and obtained his Ph.D. degree from Northwestern University in Photoinduced Electron Transfer from a Conducting Polymer to
2004. He spent most his research career at Polyera Corporation before Buckminsterfullerene. Science (Washington, DC, U. S.) 1992, 258,
joining HKUST in 2012. He is currently an associate professor in the 1474−1476.
department of chemistry at HKUST, an associate director of HKUST (15) Liu, T.; Troisi, A. What Makes Fullerene Acceptors Special as
Energy Institute, and an adjunct professor at South China University Electron Acceptors in Organic Solar Cells and How to Replace Them.
Adv. Mater. 2013, 25, 1038−1041.
of Technology. His main research interests focus on organic (16) Ganesamoorthy, R.; Sathiyan, G.; Sakthivel, P. Review:
semiconducting materials for organic and perovskite solar cells. Fullerene Based Acceptors for Efficient Bulk Heterojunction Organic
Solar Cell Applications. Sol. Energy Mater. Sol. Cells 2017, 161, 102−
ACKNOWLEDGMENTS 148.
(17) Fraga Domínguez, I.; Distler, A.; Lüer, L. Stability of Organic
The work described in this paper was partially supported by the Solar Cells: The Influence of Nanostructured Carbon Materials. Adv.
National Basic Research Program of China (973 Program Energy Mater. 2017, 7, 1601320.
Project Numbers 2013CB834701 and 2014CB643501), the (18) Huang, S.; Zhang, G.; Knutson, N. S.; Fontana, M. T.; Huber, R.
ShenZhen Technology and Innovation Commission (Project C.; Ferreira, A. S.; Tolbert, S. H.; Schwartz, B. J.; Rubin, Y. Beyond
Number JCYJ20170413173814007), the Hong Kong Research PCBM: Methoxylated 1,4-bisbenzyl[60]fullerene Adducts for Efficient
Organic Solar Cells. J. Mater. Chem. A 2016, 4, 416−424.
Grants Council (Project Nos. T23-407/13 N, N_HKUST623/ (19) Zhao, W.; Li, S.; Yao, H.; Zhang, S.; Zhang, Y.; Yang, B.; Hou, J.
13, 16305915, 16322416, 606012, 16306117, and 16303917), Molecular Optimization Enables over 13% Efficiency in Organic Solar
HK JEBN Limited, HKUST president’s office (Project FP201), Cells. J. Am. Chem. Soc. 2017, 139, 7148−7151.
and the National Science Foundation of China (No. (20) Cui, Y.; Yao, H.; Gao, B.; Qin, Y.; Zhang, S.; Yang, B.; He, C.;
21374090). We especially thank Hong Kong Innovation and Xu, B.; Hou, J. Fine-Tuned Photoactive and Interconnection Layers

3495 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

for Achieving over 13% Efficiency in a Fullerene-Free Tandem (37) Shastry, T. A.; Hartnett, P. E.; Wasielewski, M. R.; Marks, T. J.;
Organic Solar Cell. J. Am. Chem. Soc. 2017, 139, 7302−7309. Hersam, M. C. Ternary Polymer−Perylenediimide−Carbon Nanotube
(21) Elumalai, N. K.; Uddin, A. Open Circuit Voltage of Organic Photovoltaics with High Efficiency and Stability under Super-Solar
Solar Cells: An in-Depth Review. Energy Environ. Sci. 2016, 9, 391− Irradiation. ACS Energy Lett. 2016, 1, 548−555.
410. (38) Duan, Y.; Xu, X.; Yan, H.; Wu, W.; Li, Z.; Peng, Q. Pronounced
(22) Baran, D.; Kirchartz, T.; Wheeler, S.; Dimitrov, S.; Abdelsamie, Effects of a Triazine Core on Photovoltaic Performance−Efficient
M.; Gorman, J.; Ashraf, R. S.; Holliday, S.; Wadsworth, A.; Gasparini, Organic Solar Cells Enabled by a PDI Trimer-Based Small Molecular
N.; et al. Reduced Voltage Losses Yield 10% Efficient Fullerene Free Acceptor. Adv. Mater. 2017, 29, 1605115.
Organic Solar Cells with > 1 V Open Circuit Voltages. Energy Environ. (39) Lin, Y.; Li, Y.; Zhan, X. Small Molecule Semiconductors for
Sci. 2016, 9, 3783−3793. High-Efficiency Organic Photovoltaics. Chem. Soc. Rev. 2012, 41, 4245.
(23) Li, S.; Liu, W.; Li, C.-Z.; Shi, M.; Chen, H. Efficient Organic (40) Zhan, C.; Yao, J. More than Conformational “Twisting” or
Solar Cells with Non-Fullerene Acceptors. Small 2017, 13, 1701120. “Coplanarity”: Molecular Strategies for Designing High-Efficiency
(24) Nikolis, V. C.; Benduhn, J.; Holzmueller, F.; Piersimoni, F.; Lau, Nonfullerene Organic Solar Cells. Chem. Mater. 2016, 28, 1948−1964.
M.; Zeika, O.; Neher, D.; Koerner, C.; Spoltore, D.; Vandewal, K. (41) Zhan, C.; Zhang, X.; Yao, J. New Advances in Non-Fullerene
Reducing Voltage Losses in Cascade Organic Solar Cells While Acceptor Based Organic Solar Cells. RSC Adv. 2015, 5, 93002−93026.
Maintaining High External Quantum Efficiencies. Adv. Energy Mater. (42) Liang, N.; Jiang, W.; Hou, J.; Wang, Z. New Developments in
2017, 7, 1700855.
Non-Fullerene Small Molecule Acceptors for Polymer Solar Cells.
(25) Holliday, S.; Ashraf, R. S.; Wadsworth, A.; Baran, D.; Yousaf, S.
Mater. Chem. Front. 2017, 1, 1291−1303.
A.; Nielsen, C. B.; Tan, C. H.; Dimitrov, S. D.; Shang, Z.; Gasparini,
(43) Nielsen, C. B.; Holliday, S.; Chen, H.-Y.; Cryer, S. J.;
N.; et al. High-Efficiency and Air-Stable P3HT-Based Polymer Solar
McCulloch, I. Non-Fullerene Electron Acceptors for Use in Organic
Cells with a New Non-Fullerene Acceptor. Nat. Commun. 2016, 7,
11585. Solar Cells. Acc. Chem. Res. 2015, 48, 2803−2812.
(26) Baran, D.; Ashraf, R. S.; Hanifi, D. A.; Abdelsamie, M.; (44) Kang, H.; Lee, W.; Oh, J.; Kim, T.; Lee, C.; Kim, B. J. From
Gasparini, N.; Röhr, J. A.; Holliday, S.; Wadsworth, A.; Lockett, S.; Fullerene-Polymer to All-Polymer Solar Cells: The Importance of
Neophytou, M.; et al. Reducing the Efficiency-Stability-Cost Gap of Molecular Packing, Orientation, and Morphology Control. Acc. Chem.
Organic Photovoltaics with Highly Efficient and Stable Small Molecule Res. 2016, 49, 2424−2434.
Acceptor Ternary Solar Cells. Nat. Mater. 2017, 16, 363−369. (45) Benten, H.; Mori, D.; Ohkita, H.; Ito, S. Recent Research
(27) Gasparini, N.; Salvador, M.; Strohm, S.; Heumueller, T.; Progress of Polymer Donor/polymer Acceptor Blend Solar Cells. J.
Levchuk, I.; Wadsworth, A.; Bannock, J. H.; de Mello, J. C.; Egelhaaf, Mater. Chem. A 2016, 4, 5340−5365.
H.-J.; Baran, D.; et al. Burn-in Free Nonfullerene-Based Organic Solar (46) Guo, X.; Tu, D.; Liu, X. Recent Advances in Rylene Diimide
Cells. Adv. Energy Mater. 2017, 7, 1700770. Polymer Acceptors for All-Polymer Solar Cells. J. Energy Chem. 2015,
(28) Wadsworth, A.; Ashraf, R. S.; Abdelsamie, M.; Pont, S.; Little, 24, 675−685.
M.; Moser, M.; Hamid, Z.; Neophytou, M.; Zhang, W.; Amassian, A.; (47) Facchetti, A. Polymer Donor-Polymer Acceptor (All-Polymer)
et al. Highly Efficient and Reproducible Nonfullerene Solar Cells from Solar Cells. Mater. Today 2013, 16, 123−132.
Hydrocarbon Solvents. ACS Energy Lett. 2017, 2, 1494−1500. (48) Deshmukh, K. D.; Qin, T.; Gallaher, J. K.; Liu, A. C. Y.; Gann,
(29) Zhao, W.; Qian, D.; Zhang, S.; Li, S.; Inganäs, O.; Gao, F.; Hou, E.; O’Donnell, K.; Thomsen, L.; Hodgkiss, J. M.; Watkins, S. E.;
J. Fullerene-Free Polymer Solar Cells with over 11% Efficiency and McNeill, C. R. Performance, Morphology and Photophysics of High
Excellent Thermal Stability. Adv. Mater. 2016, 28, 4734−4739. Open-Circuit Voltage, Low Band Gap All-Polymer Solar Cells. Energy
(30) Chen, S.; Cho, H. J.; Lee, J.; Yang, Y.; Zhang, Z.-G.; Li, Y.; Yang, Environ. Sci. 2015, 8, 332−342.
C. Modulating the Molecular Packing and Nanophase Blending via a (49) Jung, J. W.; Jo, J. W.; Chueh, C.-C.; Liu, F.; Jo, W. H.; Russell,
Random Terpolymerization Strategy toward 11% Efficiency Non- T. P.; Jen, A. K.-Y. Fluoro-Substituted N-Type Conjugated Polymers
fullerene Polymer Solar Cells. Adv. Energy Mater. 2017, 7, 1701125. for Additive-Free All-Polymer Bulk Heterojunction Solar Cells with
(31) Li, Q.-Y.; Xiao, J.; Tang, L.-M.; Wang, H.-C.; Chen, Z.; Yang, Z.; High Power Conversion Efficiency of 6.71%. Adv. Mater. 2015, 27,
Yip, H.-L.; Xu, Y.-X. Thermally Stable High Performance Non- 3310−3317.
Fullerene Polymer Solar Cells with Low Energy Loss by Using Ladder- (50) Lee, C.; Kang, H.; Lee, W.; Kim, T.; Kim, K. H.; Woo, H. Y.;
Type Small Molecule Acceptors. Org. Electron. 2017, 44, 217−224. Wang, C.; Kim, B. J. High-Performance All-Polymer Solar Cells via
(32) Park, G. E.; Choi, S.; Park, S. Y.; Lee, D. H.; Cho, M. J.; Choi, D. Side-Chain Engineering of the Polymer Acceptor: The Importance of
H. Eco-Friendly Solvent-Processed Fullerene-Free Polymer Solar Cells the Polymer Packing Structure and the Nanoscale Blend Morphology.
with over 9.7% Efficiency and Long-Term Performance Stability. Adv.
Adv. Mater. 2015, 27, 2466−2471.
Energy Mater. 2017, 7, 1700566.
(51) Kim, Y.; Lim, E. Development of Polymer Acceptors for
(33) Zhao, F.; Dai, S.; Wu, Y.; Zhang, Q.; Wang, J.; Jiang, L.; Ling,
Organic Photovoltaic Cells. Polymers (Basel, Switz.) 2014, 6, 382−407.
Q.; Wei, Z.; Ma, W.; You, W.; et al. Single-Junction Binary-Blend
(52) Earmme, T.; Hwang, Y. J.; Subramaniyan, S.; Jenekhe, S. A. All-
Nonfullerene Polymer Solar Cells with 12.1% Efficiency. Adv. Mater.
Polymer Bulk Heterojuction Solar Cells with 4.8% Efficiency Achieved
2017, 29, 1700144.
(34) Ko, E. Y.; Park, G. E.; Lee, J. H.; Kim, H. J.; Lee, D. H.; Ahn, H.; by Solution Processing from a Co-Solvent. Adv. Mater. 2014, 26,
Uddin, M. A.; Woo, H. Y.; Cho, M. J.; Choi, D. H. Excellent Long- 6080−6085.
Term Stability of Power Conversion Efficiency in Non-Fullerene- (53) Hwang, Y. J.; Earmme, T.; Courtright, B. A. E.; Eberle, F. N.;
Based Polymer Solar Cells Bearing Tricyanovinylene-Functionalized Jenekhe, S. A. N-Type Semiconducting Naphthalene Diimide-Perylene
N-Type Small Molecules. ACS Appl. Mater. Interfaces 2017, 9, 8838− Diimide Copolymers: Controlling Crystallinity, Blend Morphology,
8847. and Compatibility Toward High-Performance All-Polymer Solar Cells.
(35) Lin, H.; Chen, S.; Hu, H.; Zhang, L.; Ma, T.; Lai, J. Y. L.; Li, Z.; J. Am. Chem. Soc. 2015, 137, 4424−4434.
Qin, A.; Huang, X.; Tang, B.; et al. Reduced Intramolecular Twisting (54) Vezie, M. S.; Few, S.; Meager, I.; Pieridou, G.; Dörling, B.;
Improves the Performance of 3D Molecular Acceptors in Non- Ashraf, R. S.; Goñi, A. R.; Bronstein, H.; McCulloch, I.; Hayes, S. C.;
Fullerene Organic Solar Cells. Adv. Mater. 2016, 28, 8546−8551. et al. Exploring the Origin of High Optical Absorption in Conjugated
(36) Li, S.; Liu, W.; Shi, M.; Mai, J.; Lau, T.-K.; Wan, J.; Lu, X.; Li, Polymers. Nat. Mater. 2016, 15, 746−753.
C.-Z.; Chen, H. A Spirobifluorene and Diketopyrrolopyrrole Moieties (55) Bredas, J.-L. Molecular Understanding of Organic Solar Cells:
Based Non-Fullerene Acceptor for Efficient and Thermally Stable The Challenges. AIP Conf. Proc. 2012, 42, 55−58.
Polymer Solar Cells with High Open-Circuit Voltage. Energy Environ. (56) Clarke, T. M.; Durrant, J. R. Charge Photogeneration in Organic
Sci. 2016, 9, 604−610. Solar Cells. Chem. Rev. 2010, 110, 6736−6767.

3496 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

(57) Yang, Y.; Li, G. Progress in High-Efficient Solution Process (78) Albrecht, S.; Vandewal, K.; Tumbleston, J. R.; Fischer, F. S. U.;
Organic Photovoltaic Devices. Yang, Y., Li, G., Eds.; Topics in Applied Douglas, J. D.; Fréchet, J. M. J.; Ludwigs, S.; Ade, H.; Salleo, A.; Neher,
Physics; Springer: Berlin, 2015; Vol. 130. D. On the Efficiency of Charge Transfer State Splitting in
(58) Gregg, B. A. Excitonic Solar Cells. J. Phys. Chem. B 2003, 107, Polymer:Fullerene Solar Cells. Adv. Mater. 2014, 26, 2533−2539.
4688−4698. (79) Jamieson, F. C.; Domingo, E. B.; McCarthy-Ward, T.; Heeney,
(59) Deibel, C.; Dyakonov, V. Polymer-Fullerene Bulk Hetero- M.; Stingelin, N.; Durrant, J. R. Fullerenecrystallisation as a Key Driver
junction Solar Cells. Rep. Prog. Phys. 2010, 73, 096401. of Charge Separation in Polymer/fullerene Bulk Heterojunction Solar
(60) Tang, C. W. Two-Layer Organic Photovoltaic Cell. Appl. Phys. Cells. Chem. Sci. 2012, 3, 485−492.
Lett. 1986, 48, 183−185. (80) Zusan, A.; Vandewal, K.; Allendorf, B.; Hansen, N. H.; Pflaum,
(61) Mikhnenko, O. V.; Cordella, F.; Sieval, A. B.; Hummelen, J. C.; J.; Salleo, A.; Dyakonov, V.; Deibel, C. The Crucial Influence of
Blom, P. W. M.; Loi, M. A. Temperature Dependence of Exciton Fullerene Phases on Photogeneration in Organic Bulk Geterojunction
Diffusion in Conjugated Polymers. J. Phys. Chem. B 2008, 112, 11601− Solar Cells. Adv. Energy Mater. 2014, 4, 1400922.
11604. (81) Savoie, B. M.; Rao, A.; Bakulin, A. A.; Gelinas, S.; Movaghar, B.;
(62) Mikhnenko, O. V.; Blom, P. W. M.; Nguyen, T.-Q. Exciton Friend, R. H.; Marks, T. J.; Ratner, M. A. Unequal Partnership:
Diffusion in Organic Semiconductors. Energy Environ. Sci. 2015, 8, Asymmetric Roles of Polymeric Donor and Fullerene Acceptor in
1867−1888. Generating Free Charge. J. Am. Chem. Soc. 2014, 136, 2876−2884.
(63) Menke, S. M.; Holmes, R. J. Exciton Diffusion in Organic (82) Bernardo, B.; Cheyns, D.; Verreet, B.; Schaller, R. D.; Rand, B.
Photovoltaic Cells. Energy Environ. Sci. 2014, 7, 499−512. P.; Giebink, N. C. Delocalization and Dielectric Screening of Charge
(64) Halls, J. J. M.; Walsh, C. A.; Greenham, N. C.; Marseglia, E. A.; Transfer States in Organic Photovoltaic Cells. Nat. Commun. 2014, 5,
Friend, R. H.; Moratti, S. C.; Holmes, A. B. Efficient Photodiodes from 3245.
Interpenetrating Polymer Networks. Nature 1995, 376, 498−500. (83) Gelinas, S.; Rao, A.; Kumar, A.; Smith, S. L.; Chin, A. W.; Clark,
(65) Cnops, K.; Rand, B. P.; Cheyns, D.; Verreet, B.; Empl, M. A.; J.; van der Poll, T. S.; Bazan, G. C.; Friend, R. H. Ultrafast Long-Range
Heremans, P. 8.4% Efficient Fullerene-Free Organic Solar Cells Charge Separation in Organic Semiconductor Photovoltaic Diodes.
Exploiting Long-Range Exciton Energy Transfer. Nat. Commun. 2014, Science (Washington, DC, U. S.) 2014, 343, 512−516.
5, 3406. (84) Jakowetz, A. C.; Böhm, M. L.; Zhang, J.; Sadhanala, A.;
(66) Menke, S. M.; Luhman, W. A.; Holmes, R. J. Tailored Exciton Huettner, S.; Bakulin, A. A.; Rao, A.; Friend, R. H. What Controls the
Diffusion in Organic Photovoltaic Cells for Enhanced Power Rate of Ultrafast Charge Transfer and Charge Separation Efficiency in
Conversion Efficiency. Nat. Mater. 2013, 12, 152−157. Organic Photovoltaic Blends. J. Am. Chem. Soc. 2016, 138, 11672−
(67) Marcus, R. A. Electron Transfer Reactions in Chemistry. Theory 11679.
and Experiment. Rev. Mod. Phys. 1993, 65, 599−610. (85) Ndjawa, G. O. N.; Graham, K. R.; Mollinger, S.; Wu, D. M.;
(68) Bässler, H.; Köhler, A. Hot or Cold”: How Do Charge Transfer Hanifi, D.; Prasanna, R.; Rose, B. D.; Dey, S.; Yu, L.; Brédas, J. L.; et al.
States at the Donor−acceptor Interface of an Organic Solar Cell Open-Circuit Voltage in Organic Solar Cells: The Impacts of Donor
Dissociate? Phys. Chem. Chem. Phys. 2015, 17, 28451−28462. Semicrystallinity and Coexistence of Multiple Interfacial Charge-
(69) Vandewal, K.; Ma, Z.; Bergqvist, J.; Tang, Z.; Wang, E.; Transfer Bands. Adv. Energy Mater. 2017, 7, 1601995.
Henriksson, P.; Tvingstedt, K.; Andersson, M. R.; Zhang, F. (86) Burke, T. M.; Sweetnam, S.; Vandewal, K.; McGehee, M. D.
Quantification of Quantum Efficiency and Energy Losses in Low Beyond Langevin Recombination: How Equilibrium between Free
Bandgap Polymer: Fullerene Solar Cells with High Open-Circuit Carriers and Charge Transfer States Determines the Open-Circuit
Voltage. Adv. Funct. Mater. 2012, 22, 3480−3490. Voltage of Organic Solar Cells. Adv. Energy Mater. 2015, 5, 1500123.
(70) Janssen, R. A. J.; Nelson, J. Factors Limiting Device Efficiency in (87) Liu, J.; Chen, S.; Qian, D.; Gautam, B.; Yang, G.; Zhao, J.;
Organic Photovoltaics. Adv. Mater. 2013, 25, 1847−1858. Bergqvist, J.; Zhang, F.; Ma, W.; Ade, H.; et al. Fast Charge Separation
(71) Li, W.; Hendriks, K. H.; Furlan, A.; Wienk, M. M.; Janssen, R. A. in a Non-Fullerene Organic Solar Cell with a Small Driving Force. Nat.
J. High Quantum Efficiencies in Polymer Solar Cells at Energy Losses Energy 2016, 1, 16089.
below 0.6 eV. J. Am. Chem. Soc. 2015, 137, 2231−2234. (88) Chen, S.; Liu, Y.; Zhang, L.; Chow, P. C. Y.; Wang, Z.; Zhang,
(72) Jailaubekov, A. E.; Willard, A. P.; Tritsch, J. R.; Chan, W. L.; Sai, G.; Ma, W.; Yan, H. A Wide-Bandgap Donor Polymer for Highly
N.; Gearba, R.; Kaake, L. G.; Williams, K. J.; Leung, K.; Rossky, P. J.; Efficient Non-Fullerene Organic Solar Cells with a Small Voltage Loss.
et al. Hot Charge-Transfer Excitons Set the Time Limit for Charge J. Am. Chem. Soc. 2017, 139, 6298−6301.
Separation at Donor/acceptor Interfaces in Organic Photovoltaics. (89) Goris, L.; Haenen, K.; Nesládek, M.; Wagner, P.; Vanderzande,
Nat. Mater. 2013, 12, 66−73. D.; De Schepper, L.; D’haen, J.; Lutsen, L.; Manca, J. V. Absorption
(73) Bakulin, A. A.; Rao, A.; Pavelyev, V. G.; van Loosdrecht, P. H. Phenomena in Organic Thin Films for Solar Cell Applications
M.; Pshenichnikov, M. S.; Niedzialek, D.; Cornil, J.; Beljonne, D.; Investigated by Photothermal Deflection Spectroscopy. J. Mater. Sci.
Friend, R. H. The Role of Driving Energy and Delocalized States for 2005, 40, 1413−1418.
Charge Separation in Organic Semiconductors. Science (Washington, (90) Stoltzfus, D. M.; Donaghey, J. E.; Armin, A.; Shaw, P. E.; Burn,
DC, U. S.) 2012, 335, 1340−1344. P. L.; Meredith, P. Charge Generation Pathways in Organic Solar
(74) Scharber, M. Measuring Internal Quantum Efficiency to Cells: Assessing the Contribution from the Electron Acceptor. Chem.
Demonstrate Hot Exciton Dissociation. Nat. Mater. 2013, 12, 594. Rev. 2016, 116, 12920−12955.
(75) Faist, M. A.; Kirchartz, T.; Gong, W.; Ashraf, R. S.; McCulloch, (91) Deotare, P. B.; Chang, W.; Hontz, E.; Congreve, D. N.; Shi, L.;
I.; De Mello, J. C.; Ekins-Daukes, N. J.; Bradley, D. D. C.; Nelson, J. Reusswig, P. D.; Modtland, B.; Bahlke, M. E.; Lee, C. K.; Willard, A.
Competition between the Charge Transfer State and the Singlet States P.; et al. Nanoscale Transport of Charge-Transfer States in Organic
of Donor or Acceptor Limiting the Efficiency in Polymer: Fullerene Donor-Acceptor Blends. Nat. Mater. 2015, 14, 1130−1134.
Solar Cells. J. Am. Chem. Soc. 2012, 134, 685−692. (92) Koster, L. J. A.; Smits, E. C. P.; Mihailetchi, V. D.; Blom, P. W.
(76) Bakulin, A. A.; Dimitrov, S. D.; Rao, A.; Chow, P. C. Y.; Nielsen, M. Device Model for the Operation of Polymer/fullerene Bulk
C. B.; Schroeder, B. C.; McCulloch, I.; Bakker, H. J.; Durrant, J. R.; Heterojunction Solar Cells. Phys. Rev. B: Condens. Matter Mater. Phys.
Friend, R. H. Charge-Transfer State Dynamics Following Hole and 2005, 72, 85205.
Electron Transfer in Organic Photovoltaic Devices. J. Phys. Chem. Lett. (93) Chow, P. C. Y.; Bayliss, S. L.; Lakhwani, G.; Greenham, N. C.;
2013, 4, 209−215. Friend, R. H. In Situ Optical Measurement of Charge Transport
(77) Vandewal, K.; Albrecht, S.; Hoke, E. T.; Graham, K. R.; Widmer, Dynamics in Organic Photovoltaics. Nano Lett. 2015, 15, 931−935.
J.; Douglas, J. D.; Schubert, M.; Mateker, W. R.; Bloking, J. T.; (94) Proctor, C. M.; Love, J. A.; Nguyen, T. Q. Mobility Guidelines
Burkhard, G. F.; et al. Efficient Charge Generation by Relaxed Charge- for High Fill Factor Solution-Processed Small Molecule Solar Cells.
Transfer States at Organic Interfaces. Nat. Mater. 2014, 13, 63−68. Adv. Mater. 2014, 26, 5957−5961.

3497 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

(95) Rao, A.; Chow, P. C. Y.; Gélinas, S.; Schlenker, C. W.; Li, C.-Z.; Perylene-3,4-Dicarboximide Trefoils. J. Phys. Chem. Lett. 2014, 5,
Yip, H.-L.; Jen, A. K.-Y.; Ginger, D. S.; Friend, R. H. The Role of Spin 1608−1615.
in the Kinetic Control of Recombination in Organic Photovoltaics. (115) Banda, H.; Damien, D.; Nagarajan, K.; Raj, A.; Hariharan, M.;
Nature 2013, 500, 435−439. Shaijumon, M. M. Twisted Perylene Diimides with Tunable Redox
(96) Brédas, J. L.; Beljonne, D.; Coropceanu, V.; Cornil, J. Charge- Properties for Organic Sodium-Ion Batteries. Adv. Energy Mater. 2017,
Transfer and Energy-Transfer Processes in π-Conjugated Oligomers 7, 1701316.
and Polymers: A Molecular Picture. Chem. Rev. 2004, 104, 4971− (116) Liu, C.; Liu, Z.; Lemke, H. T.; Tsao, H. N.; Naber, R. C. G.; Li,
5003. Y.; Banger, K.; Müllen, K.; Nielsen, M. M.; Sirringhaus, H. High-
(97) Chow, P. C. Y.; Albert-Seifried, S.; Gélinas, S.; Friend, R. H. Performance Solution-Deposited Ambipolar Organic Transistors
Nanosecond Intersystem Crossing Times in Fullerene Acceptors: Based on Terrylene Diimides. Chem. Mater. 2010, 22, 2120−2124.
Implications for Organic Photovoltaic Diodes. Adv. Mater. 2014, 26, (117) Dittmer, I. J.; Petri, K. Photovoltaic Properties of MEH-PPV/
4851−4854. PPEI Blend Devices. Synth. Met. 1999, 102, 879−880.
(98) Chow, P. C. Y.; Gélinas, S.; Rao, A.; Friend, R. H. Quantitative (118) Dittmer, J. J.; Marseglia, E. A.; Friend, R. H. Electron Trapping
Bimolecular Recombination in Organic Photovoltaics through Triplet in Dye/polymer Blend Photovoltaic Cells. Adv. Mater. 2000, 12,
Exciton Formation. J. Am. Chem. Soc. 2014, 136, 3424−3429. 1270−1274.
(99) Menke, S. M.; Sadhanala, A.; Nikolka, M.; Ran, N. A.; Ravva, M. (119) Schmidt-Mende, L. Self-Organized Discotic Liquid Crystals for
K.; Abdel-Azeim, S.; Stern, H. L.; Wang, M.; Sirringhaus, H.; Nguyen, High-Efficiency Organic Photovoltaics. Science (Washington, DC, U. S.)
T. Q.; et al. Limits for Recombination in a Low Energy Loss Organic 2001, 293, 1119−1122.
Heterojunction. ACS Nano 2016, 10, 10736−10744. (120) Sharenko, A.; Proctor, C. M.; Van Der Poll, T. S.; Henson, Z.
(100) Bredas, J.-L. Mind the Gap! Mater. Horiz. 2014, 1, 17−19. B.; Nguyen, T. Q.; Bazan, G. C. A High-Performing Solution-
(101) Zhan, X.; Facchetti, A.; Barlow, S.; Marks, T. J.; Ratner, M. A.; Processed Small Molecule: Perylene Diimide Bulk Heterojunction
Wasielewski, M. R.; Marder, S. R. Rylene and Related Diimides for Solar Cell. Adv. Mater. 2013, 25, 4403−4406.
Organic Electronics. Adv. Mater. 2011, 23, 268−284. (121) Singh, R.; Aluicio-Sarduy, E.; Kan, Z.; Ye, T.; MacKenzie, R. C.
(102) Li, C.; Liu, M.; Pschirer, N. G.; Baumgarten, M.; Müllen, K. I.; Keivanidis, P. E. Fullerene-Free Organic Solar Cells with an
Polyphenylene-Based Materials for Organic Photovoltaics. Chem. Rev. Efficiency of 3.7% Based on a Low-Cost Geometrically Planar Perylene
2010, 110, 6817−6855. Diimide Monomer. J. Mater. Chem. A 2014, 2, 14348−14353.
(103) Nolde, F.; Pisula, W.; Müller, S.; Kohl, C.; Müllen, K. Synthesis (122) Shin, W. S.; Jeong, H.-H.; Kim, M.-K.; Jin, S.-H.; Kim, M.-R.;
and Self-Organization of Core-Extended Perylene Tetracarboxdiimides Lee, J.-K.; Lee, J. W.; Gal, Y.-S. Effects of Functional Groups at
with Branched Alkyl Substituents. Chem. Mater. 2006, 18, 3715−3725. Perylene Diimide Derivatives on Organic Photovoltaic Device
(104) Würthner, F. Perylene Bisimide Dyes as Versatile Building Application. J. Mater. Chem. 2006, 16, 384−390.
Blocks for Functional Supramolecular Architectures. Chem. Commun. (123) Mikroyannidis, J. A.; Stylianakis, M. M.; Suresh, P.; Sharma, G.
2004, No. 14, 1564−1579. D. Efficient Hybrid Bulk Heterojunction Solar Cells Based on
(105) Arnaud, A.; Belleney, J.; Boué, F.; Bouteiller, L.; Carrot, G.; Phenylenevinylene Copolymer, Perylene Bisimide and TiO2. Sol.
Wintgens, V. Aqueous Supramolecular Polymer Formed from an Energy Mater. Sol. Cells 2009, 93, 1792−1800.
Amphiphilic Perylene Derivative. Angew. Chem., Int. Ed. 2004, 43, (124) Sharma, G. D.; Balraju, P.; Mikroyannidis, J. A.; Stylianakis, M.
1718−1721. M. Bulk Heterojunction Organic Photovoltaic Devices Based on Low
(106) Zhang, X.; Rehm, S.; Safont-Sempere, M. M.; Würthner, F. Band Gap Small Molecule BTD-TNP and Perylene-Anthracene
Vesicular Perylene Dye Nanocapsules as Supramolecular Fluorescent Diimide. Sol. Energy Mater. Sol. Cells 2009, 93, 2025−2028.
pH Sensor Systems. Nat. Chem. 2009, 1, 623−629. (125) Rajaram, S.; Armstrong, P. B.; Bumjoon, J. K.; Fréchet, J. M. J.
(107) Rao, K. V.; George, S. J. Supramolecular Alternate Co- Effect of Addition of a Diblock Copolymer on Blend Morphology and
Assembly through a Non-Covalent Amphiphilic Design: Conducting Performance of poly(3-Hexylthiophene):perylene Diimide Solar Cells.
Nanotubes with a Mixed D-A Structure. Chem. - Eur. J. 2012, 18, Chem. Mater. 2009, 21, 1775−1777.
14286−14291. (126) Sharma, G. D.; Suresh, P.; Mikroyannidis, J. A.; Stylianakis, M.
(108) Gao, Y.; Li, H.; Yin, S.; Liu, G.; Cao, L.; Li, Y.; Wang, X.; Ou, M. Efficient Bulk Heterojunction Devices Based on Phenylenevinylene
Z.; Wang, X. Supramolecular Electron Donor−acceptor Complexes Small Molecule and Perylene−pyrene Bisimide. J. Mater. Chem. 2010,
Formed by Perylene Diimide Derivative and Conjugated Phenazines. 20, 561−567.
New J. Chem. 2014, 38, 5647−5653. (127) Hartnett, P. E.; Timalsina, A.; Matte, H. S. S. R.; Zhou, N.;
(109) Liu, X.; Zhang, Y.; Pang, X.; Yue, E.; Zhang, Y.; Yang, D.; Guo, X.; Zhao, W.; Facchetti, A.; Chang, R. P. H.; Hersam, M. C.;
Tang, J.; Li, J.; Che, Y.; Zhao, J. Nanocoiled Assembly of Asymmetric Wasielewski, M. R.; et al. Slip-Stacked Perylenediimides as an
Perylene Diimides: Formulation of Structural Factors. J. Phys. Chem. C Alternative Strategy for High Efficiency Nonfullerene Acceptors in
2015, 119, 6446−6452. Organic Photovoltaics. J. Am. Chem. Soc. 2014, 136, 16345−16356.
(110) Céspedes-Guirao, F. J.; García-Santamaría, S.; Fernndez-Lzaro, (128) Cai, Y.; Huo, L.; Sun, X.; Wei, D.; Tang, M.; Sun, Y. High
F.; Sastre-Santos, A.; Bolink, H. J. Efficient Electroluminescence from a Performance Organic Solar Cells Based on a Twisted Bay-Substituted
Perylenediimide Fluorophore Obtained from a Simple Solution Tetraphenyl Functionalized Perylenediimide Electron Acceptor. Adv.
Processed OLED. J. Phys. D: Appl. Phys. 2009, 42, 105106. Energy Mater. 2015, 5, 1500032.
(111) Li, L.; Guan, M.; Cao, G.; Li, Y.; Zeng, Y. Highly Efficient and (129) Rajaram, S.; Shivanna, R.; Kandappa, S. K.; Narayan, K. S.
Stable Organic Light-Emitting Diodes Employing MoO3-Doped Nonplanar Perylene Diimides as Potential Alternatives to Fullerenes in
Perylene-3, 4, 9, 10-Tetracarboxylic Dianhydride as Hole Injection Organic Solar Cells. J. Phys. Chem. Lett. 2012, 3, 2405−2408.
Layer. Appl. Phys. A: Mater. Sci. Process. 2010, 99, 251−254. (130) Zhang, X.; Lu, Z.; Ye, L.; Zhan, C.; Hou, J.; Zhang, S.; Jiang, B.;
(112) Kim, S. H.; Yang, Y. S.; Lee, J. H.; Lee, J.-I.; Chu, H. Y.; Lee, Zhao, Y.; Huang, J.; Zhang, S.; et al. A Potential Perylene Diimide
H.; Oh, J.; Do, L.-M.; Zyung, T. Organic Field-Effect Transistors Dimer-Based Acceptor Material for Highly Efficient Solution-
Using Perylene. Opt. Mater. (Amsterdam, Neth.) 2003, 21, 439−443. Processed Non-Fullerene Organic Solar Cells with 4.03% Efficiency.
(113) Weitz, R. T.; Amsharov, K.; Zschieschang, U.; Villas, E. B.; Adv. Mater. 2013, 25, 5791−5797.
Goswami, D. K.; Burghard, M.; Dosch, H.; Jansen, M.; Kern, K.; Klauk, (131) Lu, Z.; Jiang, B.; Zhang, X.; Tang, A.; Chen, L.; Zhan, C.; Yao,
H. Organic N-Channel Transistors Based on Core-Cyanated Perylene J. Perylene-Diimide Based Non-Fullerene Solar Cells with 4.34%
Carboxylic Diimide Derivatives. J. Am. Chem. Soc. 2008, 130, 4637− Efficiency through Engineering Surface Donor/acceptor Composi-
4645. tions. Chem. Mater. 2014, 26, 2907−2914.
(114) Lefler, K. M.; Kim, C. H.; Wu, Y. L.; Wasielewski, M. R. Self- (132) Jiang, W.; Ye, L.; Li, X.; Xiao, C.; Tan, F.; Zhao, W.; Hou, J.;
Assembly of Supramolecular Light-Harvesting Arrays from Symmetric Wang, Z. Bay-Linked Perylene Bisimides as Promising Non-Fullerene

3498 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Acceptors for Organic Solar Cells. Chem. Commun. 2014, 50, 1024− (150) Lee, J.; Singh, R.; Sin, D. H.; Kim, H. G.; Song, K. C.; Cho, K.
1026. A Nonfullerene Small Molecule Acceptor with 3D Interlocking
(133) Zang, Y.; Li, C. Z.; Chueh, C. C.; Williams, S. T.; Jiang, W.; Geometry Enabling Efficient Organic Solar Cells. Adv. Mater. 2016,
Wang, Z. H.; Yu, J. S.; Jen, A. K. Y. Integrated Molecular, Interfacial, 28, 69−76.
and Device Engineering towards High-Performance Non-Fullerene (151) Liu, X.; Liu, T.; Duan, C.; Wang, J.; Pang, S.; Xiong, W.; Sun,
Based Organic Solar Cells. Adv. Mater. 2014, 26, 5708−5714. Y.; Huang, F.; Cao, Y. Non-Planar Perylenediimide Acceptors with
(134) Sun, D.; Meng, D.; Cai, Y.; Fan, B.; Li, Y.; Jiang, W.; Huo, L.; Different Geometrical Linker Units for Efficient Non-Fullerene
Sun, Y.; Wang, Z. Non-Fullerene-Acceptor-Based Bulk-Heterojunction Organic Solar Cells. J. Mater. Chem. A 2017, 5, 1713−1723.
Organic Solar Cells with Efficiency over 7%. J. Am. Chem. Soc. 2015, (152) Zhang, A.; Li, C.; Yang, F.; Zhang, J.; Wang, Z.; Wei, Z.; Li, W.
137, 11156−11162. An Electron Acceptor with Porphyrin and Perylene Bisimides for
(135) Meng, D.; Sun, D.; Zhong, C.; Liu, T.; Fan, B.; Huo, L.; Li, Y.; Efficient Non-Fullerene Solar Cells. Angew. Chem., Int. Ed. 2017, 56,
Jiang, W.; Choi, H.; Kim, T.; et al. High-Performance Solution- 2694−2698.
Processed Non-Fullerene Organic Solar Cells Based on Selenophene- (153) Qian, H.; Wang, Z.; Yue, W.; Zhu, D. Exceptional Coupling of
Containing Perylene Bisimide Acceptor. J. Am. Chem. Soc. 2016, 138, Tetrachloroperylene Bisimide: Combination of Ullmann Reaction and
375−380. C-H Transformation. J. Am. Chem. Soc. 2007, 129, 10664−10665.
(136) Hendsbee, A. D.; Sun, J. P.; Law, W. K.; Yan, H.; Hill, I. G.; (154) Li, Y.; Wang, C.; Li, C.; Di Motta, S.; Negri, F.; Wang, Z.
Spasyuk, D. M.; Welch, G. C. Synthesis, Self-Assembly, and Solar Cell Synthesis and Properties of Ethylene-Annulated Di(perylene Dii-
Performance of N-Annulated Perylene Diimide Non-Fullerene mides). Org. Lett. 2012, 14, 5278−5281.
Acceptors. Chem. Mater. 2016, 28, 7098−7109. (155) Zhong, Y.; Trinh, M. T.; Chen, R.; Wang, W.; Khlyabich, P. P.;
(137) Yan, Q.; Zhou, Y.; Zheng, Y.-Q.; Pei, J.; Zhao, D. Towards Kumar, B.; Xu, Q.; Nam, C. Y.; Sfeir, M. Y.; Black, C.; et al. Efficient
Rational Design of Organic Electron Acceptors for Photovoltaics: A Organic Solar Cells with Helical Perylene Diimide Electron Acceptors.
Study Based on Perylenediimide Derivatives. Chem. Sci. 2013, 4, 4389. J. Am. Chem. Soc. 2014, 136, 15215−15221.
(138) Lin, Y.; Wang, J.; Dai, S.; Li, Y.; Zhu, D.; Zhan, X. A Twisted (156) Zhong, H.; Wu, C. H.; Li, C. Z.; Carpenter, J.; Chueh, C. C.;
Dimeric Perylene Diimide Electron Acceptor for Efficient Organic Chen, J. Y.; Ade, H.; Jen, A. K. Y. Rigidifying Nonplanar Perylene
Solar Cells. Adv. Energy Mater. 2014, 4, 1400420. Diimides by Ring Fusion Toward Geometry-Tunable Acceptors for
(139) Wang, J.; Yao, Y.; Dai, S.; Zhang, X.; Wang, W.; He, Q.; Han, High-Performance Fullerene-Free Solar Cells. Adv. Mater. 2016, 28,
L.; Lin, Y.; Zhan, X. Oligothiophene-Bridged Perylene Diimide Dimers 951−958.
for Fullerene-Free Polymer Solar Cells: Effect of Bridge Length. J. (157) Hartnett, P. E.; Matte, H. S. S. R.; Eastham, N. D.; Jackson, N.
Mater. Chem. A 2015, 3, 13000−13010. E.; Wu, Y.; Chen, L. X.; Ratner, M. A.; Chang, R. P. H.; Hersam, M.
(140) Zhao, J.; Li, Y.; Zhang, J.; Zhang, L.; Lai, J. Y. L.; Jiang, K.; Mu,
C.; Wasielewski, M. R.; et al. Ring-Fusion as a Perylenediimide Dimer
C.; Li, Z.; Chan, C. L. C.; Hunt, A.; et al. The Influence of Spacer
Design Concept for High-Performance Non-Fullerene Organic
Units on Molecular Properties and Solar Cell Performance of Non-
Photovoltaic Acceptors. Chem. Sci. 2016, 7, 3543−3555.
Fullerene Acceptors. J. Mater. Chem. A 2015, 3, 20108−20112.
(158) Settanni, G.; Zhou, J.; Suo, T.; Schöttler, S.; Landfester, K.;
(141) Park, G. E.; Kim, H. J.; Choi, S.; Lee, D. H.; Uddin, M. A.;
Schmid, F.; Mailänder, V. Protein Corona Composition of PEGylated
Woo, H. Y.; Cho, M. J.; Choi, D. H. New M- and V-Shaped Perylene
Nanoparticles Correlates Strongly with Amino Acid Composition of
Diimide Small Molecules for High-Performance Nonfullerene Polymer
Protein Surface. J. Mater. Chem. A 2016, 4, 14983−14987.
Solar Cells. Chem. Commun. 2016, 52, 8873−8876.
(159) Zhong, Y.; Kumar, B.; Oh, S.; Trinh, M. T.; Wu, Y.; Elbert, K.;
(142) Hadmojo, W. T.; Nam, S. Y.; Shin, T. J.; Yoon, S. C.; Jang, S.-
Y.; Jung, I. H. Geometrically Controlled Organic Small Molecule Li, P.; Zhu, X.; Xiao, S.; Ng, F.; et al. Helical Ribbons for Molecular
Acceptors for Efficient Fullerene-Free Organic Photovoltaic Devices. J. Electronics. J. Am. Chem. Soc. 2014, 136, 8122−8130.
Mater. Chem. A 2016, 4, 12308−12318. (160) Zhong, Y.; Trinh, M. T.; Chen, R.; Purdum, G. E.; Khlyabich,
(143) Lin, Y.; Wang, Y.; Wang, J.; Hou, J.; Li, Y.; Zhu, D.; Zhan, X. A P. P.; Sezen, M.; Oh, S.; Zhu, H.; Fowler, B.; Zhang, B.; et al.
Star-Shaped Perylene Diimide Electron Acceptor for High-Perform- Molecular Helices as Electron Acceptors in High-Performance Bulk
ance Organic Solar Cells. Adv. Mater. 2014, 26, 5137−5142. Heterojunction Solar Cells. Nat. Commun. 2015, 6, 8242.
(144) Liang, N.; Sun, K.; Zheng, Z.; Yao, H.; Gao, G.; Meng, X.; (161) Sisto, T. J.; Zhong, Y.; Zhang, B.; Trinh, M. T.; Miyata, K.;
Wang, Z.; Ma, W.; Hou, J. Perylene Diimide Trimers Based Bulk Zhong, X.; Zhu, X. Y.; Steigerwald, M. L.; Ng, F.; Nuckolls, C. Long,
Heterojunction Organic Solar Cells with Efficiency over 7%. Adv. Atomically Precise Donor-Acceptor Cove-Edge Nanoribbons as
Energy Mater. 2016, 6, 1600060. Electron Acceptors. J. Am. Chem. Soc. 2017, 139, 5648−5651.
(145) Li, S.; Liu, W.; Li, C.-Z.; Liu, F.; Zhang, Y.; Shi, M.; Chen, H.; (162) Meng, D.; Fu, H.; Xiao, C.; Meng, X.; Winands, T.; Ma, W.;
Russell, T. P. A Simple Perylene Diimide Derivative with a Highly Wei, W.; Fan, B.; Huo, L.; Doltsinis, N. L.; et al. Three-Bladed Rylene
Twisted Geometry as an Electron Acceptor for Efficient Organic Solar Propellers with Three-Dimensional Network Assembly for Organic
Cells. J. Mater. Chem. A 2016, 4, 10659−10665. Electronics. J. Am. Chem. Soc. 2016, 138, 10184−10190.
(146) Liu, Y.; Mu, C.; Jiang, K.; Zhao, J.; Li, Y.; Zhang, L.; Li, Z.; Lai, (163) Wu, Q.; Zhao, D.; Yang, J.; Sharapov, V.; Cai, Z.; Li, L.; Zhang,
J. Y. L.; Hu, H.; Ma, T.; et al. A Tetraphenylethylene Core-Based 3d N.; Neshchadin, A.; Chen, W.; Yu, L. Propeller-Shaped Acceptors for
Structure Small Molecular Acceptor Enabling Efficient Non-Fullerene High-Performance Non-Fullerene Solar Cells: Importance of the
Organic Solar Cells. Adv. Mater. 2015, 27, 1015−1020. Rigidity of Molecular Geometry. Chem. Mater. 2017, 29, 1127−1133.
(147) Liu, S. Y.; Wu, C. H.; Li, C. Z.; Liu, S. Q.; Wei, K. H.; Chen, H. (164) Wang, B.; Liu, W.; Li, H.; Mai, J.; Liu, S.; Lu, X.; Li, H.; Shi,
Z.; Jen, A. K. Y. A Tetraperylene Diimides Based 3D Nonfullerene M.; Li, C.-Z.; Chen, H. Electron Acceptors with Varied Linkages
Acceptor for Efficient Organic Photovoltaics. Adv. Sci. 2015, 2, between Perylene Diimide and Benzotrithiophene for Efficient
1500014. Fullerene-Free Solar Cells. J. Mater. Chem. A 2017, 5, 9396−9401.
(148) Liu, Y.; Lai, J. Y. L.; Chen, S.; Li, Y.; Jiang, K.; Zhao, J.; Li, Z.; (165) Zhang, J.; Li, Y.; Huang, J.; Hu, H.; Zhang, G.; Ma, T.; Chow,
Hu, H.; Ma, T.; Lin, H.; et al. Efficient Non-Fullerene Polymer Solar P. C. Y.; Ade, H.; Pan, D.; Yan, H. Ring-Fusion of Perylene Diimide
Cells Enabled by Tetrahedron-Shaped Core Based 3D-Structure Acceptor Enabling Efficient Nonfullerene Organic Solar Cells with a
Small-Molecular Electron Acceptors. J. Mater. Chem. A 2015, 3, Small Voltage Loss. J. Am. Chem. Soc. 2017, 139, 16092−16095.
13632−13636. (166) Zhao, D.; Wu, Q.; Cai, Z.; Zheng, T.; Chen, W.; Lu, J.; Yu, L.
(149) Yan, W.; Zhang, Q.; Qin, Q.; Ye, S.; Lin, Y.; Liu, Z.; Bian, Z.; Electron Acceptors Based on α-Substituted Perylene Diimide (PDI)
Chen, Y.; Huang, C. Design, Synthesis and Photophysical Properties of for Organic Solar Cells. Chem. Mater. 2016, 28, 1139−1146.
A-D-A-D-A Small Molecules for Photovoltaic Application. Dyes Pigm. (167) Wu, Q.; Zhao, D.; Schneider, A. M.; Chen, W.; Yu, L.
2015, 121, 99−108. Covalently Bound Clusters of Alpha-Substituted PDI-Rival Electron

3499 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Acceptors to Fullerene for Organic Solar Cells. J. Am. Chem. Soc. 2016, (184) Wu, Y.; Bai, H.; Wang, Z.; Cheng, P.; Zhu, S.; Wang, Y.; Ma,
138, 7248−7251. W.; Zhan, X. A Planar Electron Acceptor for Efficient Polymer Solar
(168) Fan, Y.; Ziabrev, K.; Zhang, S.; Lin, B.; Barlow, S.; Marder, S. Cells. Energy Environ. Sci. 2015, 8, 3215−3221.
R. Comparison of the Optical and Electrochemical Properties of (185) Li, Y.; Liu, X.; Wu, F.-P.; Zhou, Y.; Jiang, Z.-Q.; Song, B.; Xia,
Bi(perylene Diimide)s Linked through Ortho and Bay Positions. ACS Y.; Zhang, Z.-G.; Gao, F.; Inganäs, O.; et al. Non-Fullerene Acceptor
Omega 2017, 2, 377−385. with Low Energy Loss and High External Quantum Efficiency:
(169) Fernando, R.; Mao, Z.; Muller, E.; Ruan, F.; Sauvé, G. Tuning Towards High Performance Polymer Solar Cells. J. Mater. Chem. A
the Organic Solar Cell Performance of Acceptor 2,6- Dialkylamino- 2016, 4, 5890−5897.
naphthalene Diimides by Varying a Linker between the Imide (186) Li, Y.; Qian, D.; Zhong, L.; Lin, J. D.; Jiang, Z. Q.; Zhang, Z.
Nitrogen and a Thiophene Group. J. Phys. Chem. C 2014, 118, G.; Zhang, Z.; Li, Y.; Liao, L. S.; Zhang, F. A Fused-Ring Based
3433−3442. Electron Acceptor for Efficient Non-Fullerene Polymer Solar Cells
(170) Ahmed, E.; Ren, G.; Kim, F. S.; Hollenbeck, E. C.; Jenekhe, S. with Small HOMO Offset. Nano Energy 2016, 27, 430−438.
A. Design of New Electron Acceptor Materials for Organic (187) Li, Y.; Zhong, L.; Wu, F.-P.; Yuan, Y.; Bin, H.-J.; Jiang, Z.-Q.;
Photovoltaics: Synthesis, Electron Transport, Photophysics, and Zhang, Z.; Zhang, Z.-G.; Li, Y.; Liao, L.-S. Non-Fullerene Polymer
Photovoltaic Properties of Oligothiophene-Functionalized Naphtha- Solar Cells Based on a Selenophene-Containing Fused-Ring Acceptor
lene Diimides. Chem. Mater. 2011, 23, 4563−4577. with Photovoltaic Performance of 8.6%. Energy Environ. Sci. 2016, 9,
(171) Ren, G.; Ahmed, E.; Jenekhe, S. A. Non-Fullerene Acceptor- 3429−3435.
Based Bulk Heterojunction Polymer Solar Cells: Engineering the (188) Lin, Y.; He, Q.; Zhao, F.; Huo, L.; Mai, J.; Lu, X.; Su, C. J.; Li,
Nanomorphology via Processing Additives. Adv. Energy Mater. 2011, 1, T.; Wang, J.; Zhu, J.; et al. A Facile Planar Fused-Ring Electron
946−953. Acceptor for As-Cast Polymer Solar Cells with 8.71% Efficiency. J. Am.
(172) Mao, Z.; Le, T. P.; Vakhshouri, K.; Fernando, R.; Ruan, F.; Chem. Soc. 2016, 138, 2973−2976.
Muller, E.; Gomez, E. D.; Sauvé, G. Processing Additive Suppresses (189) Lin, Y.; Li, T.; Zhao, F.; Han, L.; Wang, Z.; Wu, Y.; He, Q.;
Phase Separation in the Active Layer of Organic Photovoltaics Based Wang, J.; Huo, L.; Sun, Y.; et al. Structure Evolution of Oligomer
on Naphthalene Diimide. Org. Electron. 2014, 15, 3384−3391. Fused-Ring Electron Acceptors toward High Efficiency of As-Cast
(173) Liu, Y.; Zhang, L.; Lee, H.; Wang, H. W.; Santala, A.; Liu, F.; Polymer Solar Cells. Adv. Energy Mater. 2016, 6, 1600854.
Diao, Y.; Briseno, A. L.; Russell, T. P. NDI-Based Small Molecule as (190) Lin, Y.; Zhao, F.; He, Q.; Huo, L.; Wu, Y.; Parker, T. C.; Ma,
Promising Nonfullerene Acceptor for Solution-Processed Organic W.; Sun, Y.; Wang, C.; Zhu, D.; et al. High-Performance Electron
Photovoltaics. Adv. Energy Mater. 2015, 5, 1500195. Acceptor with Thienyl Side Chains for Organic Photovoltaics. J. Am.
(174) Feng, J.; Liang, N.; Jiang, W.; Meng, D.; Xin, R.; Xu, B.; Zhang, Chem. Soc. 2016, 138, 4955−4961.
J.; Wei, Z.; Hou, J.; Wang, Z. Twisted Terrylene Dyes: Synthesis and (191) Yang, Y.; Zhang, Z. G.; Bin, H.; Chen, S.; Gao, L.; Xue, L.;
Application in Organic Solar Cells. Org. Chem. Front. 2017, 4, 811− Yang, C.; Li, Y. Side-Chain Isomerization on an N-Type Organic
816. Semiconductor ITIC Acceptor Makes 11.77% High Efficiency Polymer
(175) Douglas, J. D.; Chen, M. S.; Niskala, J. R.; Lee, O. P.; Yiu, A. Solar Cells. J. Am. Chem. Soc. 2016, 138, 15011−15018.
T.; Young, E. P.; Fréchet, J. M. J. Solution-Processed, Molecular (192) Liu, Y.; Zhang, Z.; Feng, S.; Li, M.; Wu, L.; Hou, R.; Xu, X.;
Photovoltaics That Exploit Hole Transfer from Non-Fullerene, N- Chen, X.; Bo, Z. Exploiting Noncovalently Conformational Locking as
Type Materials. Adv. Mater. 2014, 26, 4313−4319. a Design Strategy for High Performance Fused-Ring Electron Acceptor
(176) Bloking, J. T.; Han, X.; Higgs, A. T.; Kastrop, J. P.; Pandey, L.; Used in Polymer Solar Cells. J. Am. Chem. Soc. 2017, 139, 3356−3359.
Norton, J. E.; Risko, C.; Chen, C. E.; Brédas, J. L.; McGehee, M. D.; (193) Yao, H.; Chen, Y.; Qin, Y.; Yu, R.; Cui, Y.; Yang, B.; Li, S.;
et al. Solution-Processed Organic Solar Cells with Power Conversion Zhang, K.; Hou, J. Design and Synthesis of a Low Bandgap Small
Efficiencies of 2.5% Using Benzothiadiazole/imide-Based Acceptors. Molecule Acceptor for Efficient Polymer Solar Cells. Adv. Mater. 2016,
Chem. Mater. 2011, 23, 5484−5490. 28, 8283−8287.
(177) Winzenberg, K. N.; Kemppinen, P.; Scholes, F. H.; Collis, G. (194) Yao, H.; Cui, Y.; Yu, R.; Gao, B.; Zhang, H.; Hou, J. Design,
E.; Shu, Y.; Birendra Singh, T.; Bilic, A.; Forsyth, C. M.; Watkins, S. E. Synthesis, and Photovoltaic Characterization of a Small Molecular
Indan-1,3-Dione Electron-Acceptor Small Molecules for Solution- Acceptor with an Ultra-Narrow Band Gap. Angew. Chem., Int. Ed.
Processable Solar Cells: A Structure−property Correlation. Chem. 2017, 56, 3045−3049.
Commun. 2013, 49, 6307. (195) Li, S.; Ye, L.; Zhao, W.; Zhang, S.; Mukherjee, S.; Ade, H.;
(178) Kim, Y.; Song, C. E.; Moon, S.-J.; Lim, E. Rhodanine Dye- Hou, J. Energy-Level Modulation of Small-Molecule Electron
Based Small Molecule Acceptors for Organic Photovoltaic Cells. Chem. Acceptors to Achieve over 12% Efficiency in Polymer Solar Cells.
Commun. 2014, 50, 8235−8238. Adv. Mater. 2016, 28, 9423−9429.
(179) Holliday, S.; Ashraf, R. S.; Nielsen, C. B.; Kirkus, M.; Rohr, J. (196) Dai, S.; Zhao, F.; Zhang, Q.; Lau, T.-K.; Li, T.; Liu, K.; Ling,
A.; Tan, C. H.; Collado-Fregoso, E.; Knall, A. C.; Durrant, J. R.; Q.; Wang, C.; Lu, X.; You, W.; et al. Fused Nonacyclic Electron
Nelson, J.; et al. A Rhodanine Flanked Nonfullerene Acceptor for Acceptors for Efficient Polymer Solar Cells. J. Am. Chem. Soc. 2017,
Solution-Processed Organic Photovoltaics. J. Am. Chem. Soc. 2015, 139, 1336−1343.
137, 898−904. (197) Yao, H.; Ye, L.; Hou, J.; Jang, B.; Han, G.; Cui, Y.; Su, G. M.;
(180) Bai, H.; Wang, Y.; Cheng, P.; Wang, J.; Wu, Y.; Hou, J.; Zhan, Wang, C.; Gao, B.; Yu, R.; et al. Achieving Highly Efficient
X. An Electron Acceptor Based on Indacenodithiophene and 1,1- Nonfullerene Organic Solar Cells with Improved Intermolecular
Dicyanomethylene-3-Indanone for Fullerene-Free Organic Solar Cells. Interaction and Open-Circuit Voltage. Adv. Mater. 2017, 29, 1700254.
J. Mater. Chem. A 2015, 3, 1910−1914. (198) Xie, D.; Liu, T.; Gao, W.; Zhong, C.; Huo, L.; Luo, Z.; Wu, K.;
(181) Lin, Y.; Zhang, Z.-G.; Bai, H.; Wang, J.; Yao, Y.; Li, Y.; Zhu, D.; Xiong, W.; Liu, F.; Sun, Y.; et al. A Novel Thiophene-Fused Ending
Zhan, X. High-Performance Fullerene-Free Polymer Solar Cells with Group Enabling an Excellent Small Molecule Acceptor for High-
6.31% Efficiency. Energy Environ. Sci. 2015, 8, 610−616. Performance Fullerene-Free Polymer Solar Cells with 11.8%
(182) Lin, Y.; Wang, J.; Zhang, Z.-G.; Bai, H.; Li, Y.; Zhu, D.; Zhan, Efficiency. Sol. RRL 2017, 1, 1700044.
X. An Electron Acceptor Challenging Fullerenes for Efficient Polymer (199) Ni, W.; Li, M.; Kan, B.; Liu, F.; Wan, X.; Zhang, Q.; Zhang, H.;
Solar Cells. Adv. Mater. 2015, 27, 1170−1174. Russell, T. P.; Chen, Y. Fullerene-Free Small Molecule Organic Solar
(183) Bai, H.; Wu, Y.; Wang, Y.; Wu, Y.; Li, R.; Cheng, P.; Zhang, Cells with a High Open Circuit Voltage of 1.15 V. Chem. Commun.
M.; Wang, J.; Ma, W.; Zhan, X. Nonfullerene Acceptors Based on 2016, 52, 465−468.
Extended Fused Rings Flanked with Benzothiadiazolylmethylenema- (200) Zhang, H.; Liu, Y.; Sun, Y.; Li, M.; Ni, W.; Zhang, Q.; Wan, X.;
lononitrile for Polymer Solar Cells. J. Mater. Chem. A 2015, 3, 20758− Chen, Y. A Simple Small Molecule as the Acceptor for Fullerene-Free
20766. Organic Solar Cells. Sci. China: Chem. 2017, 60, 366−369.

3500 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

(201) Qiu, N.; Zhang, H.; Wan, X.; Li, C.; Ke, X.; Feng, H.; Kan, B.; (217) Li, H.; Earmme, T.; Subramaniyan, S.; Jenekhe, S. A.
Zhang, H.; Zhang, Q.; Lu, Y.; et al. A New Nonfullerene Electron Bis(Naphthalene Imide)diphenylanthrazolines: A New Class of
Acceptor with a Ladder Type Backbone for High-Performance Electron Acceptors for Efficient Nonfullerene Organic Solar Cells
Organic Solar Cells. Adv. Mater. 2017, 29, 1604964. and Applicable to Multiple Donor Polymers. Adv. Energy Mater. 2015,
(202) Wang, K.; Firdaus, Y.; Babics, M.; Cruciani, F.; Saleem, Q.; El 5, 1402041.
Labban, A.; Alamoudi, M. A.; Marszalek, T.; Pisula, W.; Laquai, F.; (218) Kwon, O. K.; Park, J. H.; Park, S. K.; Park, S. Y. Soluble
et al. π-Bridge-Independent 2-(Benzo[c][1,2,5]thiadiazol-4- Dicyanodistyrylbenzene-Based Non-Fullerene Electron Acceptors with
Ylmethylene)malononitrile-Substituted Nonfullerene Acceptors for Optimized Aggregation Behavior for High-Efficiency Organic Solar
Efficient Bulk Heterojunction Solar Cells. Chem. Mater. 2016, 28, Cells. Adv. Energy Mater. 2015, 5, 1400929.
2200−2208. (219) Kwon, O. K.; Uddin, M. A.; Park, J. H.; Park, S. K.; Nguyen, T.
(203) Zhang, G.; Yang, G.; Yan, H.; Kim, J. H.; Ade, H.; Wu, W.; Xu, L.; Woo, H. Y.; Park, S. Y. A High Efficiency Nonfullerene Organic
X.; Duan, Y.; Peng, Q. Efficient Nonfullerene Polymer Solar Cells Solar Cell with Optimized Crystalline Organizations. Adv. Mater. 2016,
Enabled by a Novel Wide Bandgap Small Molecular Acceptor. Adv. 28, 910−916.
Mater. 2017, 29, 1606054. (220) McAfee, S. M.; Topple, J. M.; Sun, J.-P.; Hill, I. G.; Welch, G.
(204) Kan, B.; Feng, H.; Wan, X.; Liu, F.; Ke, X.; Wang, Y.; Wang, Y.; C. The Structural Evolution of an Isoindigo-Based Non-Fullerene
Zhang, H.; Li, C.; Hou, J.; et al. Small-Molecule Acceptor Based on the Acceptor for Use in Organic Photovoltaics. RSC Adv. 2015, 5, 80098−
Heptacyclic Benzodi(cyclopentadithiophene) Unit for Highly Efficient 80109.
Nonfullerene Organic Solar Cells. J. Am. Chem. Soc. 2017, 139, 4929− (221) Chatterjee, S.; Ie, Y.; Karakawa, M.; Aso, Y. Naphtho[1,2-c:5,6-
4934. c′]bis[1,2,5]thiadiazole-Containing π-Conjugated Compound: Non-
(205) Cao, Q.; Xiong, W.; Chen, H.; Cai, G.; Wang, G.; Zheng, L.; fullerene Electron Acceptor for Organic Photovoltaics. Adv. Funct.
Sun, Y. Design, Synthesis, and Structural Characterization of the First Mater. 2016, 26, 1161−1168.
Dithienocyclopentacarbazole-Based N-Type Organic Semiconductor (222) Jinnai, S.; Ie, Y.; Karakawa, M.; Aernouts, T.; Nakajima, Y.;
and Its Application in Non-Fullerene Polymer Solar Cells. J. Mater. Mori, S.; Aso, Y. Electron-Accepting π-Conjugated Systems for
Chem. A 2017, 5, 7451−7461. Organic Photovoltaics: Influence of Structural Modification on
(206) Storm, F. E.; Olsen, S. T.; Hansen, T.; De Vico, L.; Jackson, N. Molecular Orientation at Donor-Acceptor Interfaces. Chem. Mater.
E.; Ratner, M. A.; Mikkelsen, K. V. Boron Subphthalocyanine Based 2016, 28, 1705−1713.
Molecular Triad Systems for the Capture of Solar Energy. J. Phys. (223) Li, H.; Kim, F. S.; Ren, G.; Hollenbeck, E. C.; Subramaniyan,
Chem. A 2016, 120, 7694−7703. S.; Jenekhe, S. A. Tetraazabenzodifluoranthene Diimides: Building
(207) Ebenhoch, B.; Prasetya, N. B. A.; Rotello, V. M.; Cooke, G.; Blocks for Solution-Processable N-Type Organic Semiconductors.
Samuel, I. D. W. Solution-Processed Boron Subphthalocyanine
Angew. Chem., Int. Ed. 2013, 52, 5513−5517.
Derivatives as Acceptors for Organic Bulk-Heterojunction Solar (224) Li, H.; Earmme, T.; Ren, G.; Saeki, A.; Yoshikawa, S.; Murari,
Cells. J. Mater. Chem. A 2015, 3, 7345−7352.
N. M.; Subramaniyan, S.; Crane, M. J.; Seki, S.; Jenekhe, S. A. Beyond
(208) Claessens, C. G.; González-Rodríguez, D.; Rodríguez-Morgade,
Fullerenes: Design of Nonfullerene Acceptors for Efficient Organic
M. S.; Medina, A.; Torres, T. Subphthalocyanines, Subporphyrazines,
Photovoltaics. J. Am. Chem. Soc. 2014, 136, 14589−14597.
and Subporphyrins: Singular Nonplanar Aromatic Systems. Chem. Rev.
(225) Li, H.; Hwang, Y.-J.; Courtright, B. A. E.; Eberle, F. N.;
2014, 114, 2192−2277.
Subramaniyan, S.; Jenekhe, S. A. Fine-Tuning the 3D Structure of
(209) Lin, Y.; Li, Y.; Zhan, X. A Solution-Processable Electron
Nonfullerene Electron Acceptors Toward High-Performance Polymer
Acceptor Based on Dibenzosilole and Diketopyrrolopyrrole for
Organic Solar Cells. Adv. Energy Mater. 2013, 3, 724−728. Solar Cells. Adv. Mater. 2015, 27, 3266−3272.
(210) Shi, H.; Fu, W.; Shi, M.; Ling, J.; Chen, H. A Solution- (226) Hwang, Y. J.; Li, H.; Courtright, B. A. E.; Subramaniyan, S.;
Processable Bipolar Diketopyrrolopyrrole Molecule Used as Both Jenekhe, S. A. Nonfullerene Polymer Solar Cells with 8.5% Efficiency
Electron Donor and Acceptor for Efficient Organic Solar Cells. J. Enabled by a New Highly Twisted Electron Acceptor Dimer. Adv.
Mater. Chem. A 2015, 3, 1902−1905. Mater. 2016, 28, 124−131.
(211) Yang, Y.; Zhang, G.; Yu, C.; He, C.; Wang, J.; Chen, X.; Yao, J.; (227) Mao, Z.; Senevirathna, W.; Liao, J.-Y.; Gu, J.; Kesava, S. V.;
Liu, Z.; Zhang, D. New Conjugated Molecular Scaffolds Based on Guo, C.; Gomez, E. D.; Sauvé, G. Azadipyrromethene-Based Zn(II)
[2,2]paracyclophane as Electron Acceptors for Organic Photovoltaic Complexes as Nonplanar Conjugated Electron Acceptors for Organic
Cells. Chem. Commun. 2014, 50, 9939−9942. Photovoltaics. Adv. Mater. 2014, 26, 6290−6294.
(212) Wu, X. F.; Fu, W. F.; Xu, Z.; Shi, M.; Liu, F.; Chen, H. Z.; (228) Qiu, N.; Yang, X.; Zhang, H.; Wan, X.; Li, C.; Liu, F.; Zhang,
Wan, J. H.; Russell, T. P. Spiro Linkage as an Alternative Strategy for H.; Russell, T. P.; Chen, Y. Nonfullerene Small Molecular Acceptors
Promising Nonfullerene Acceptors in Organic Solar Cells. Adv. Funct. with a Three-Dimensional (3D) Structure for Organic Solar Cells.
Mater. 2015, 25, 5954−5966. Chem. Mater. 2016, 28, 6770−6778.
(213) Jung, J. W.; Jo, W. H. Low-Bandgap Small Molecules as Non- (229) Kuzmich, A.; Padula, D.; Ma, H.; Troisi, A. Trends in the
Fullerene Electron Acceptors Composed of Benzothiadiazole and Electronic and Geometric Structure of Non-Fullerene Based Acceptors
Diketopyrrolopyrrole for All Organic Solar Cells. Chem. Mater. 2015, for Organic Solar Cells. Energy Environ. Sci. 2017, 10, 395−401.
27, 6038−6043. (230) Lopez, S. A.; Sanchez-Lengeling, B.; de Goes Soares, J.;
(214) Pho, T. V.; Toma, F. M.; Tremolet De Villers, B. J.; Wang, S.; Aspuru-Guzik, A. Design Principles and Top Non-Fullerene Acceptor
Treat, N. D.; Eisenmenger, N. D.; Su, G. M.; Coffin, R. C.; Douglas, J. Candidates for Organic Photovoltaics. Joule 2017, 1, 857.
D.; Fréchet, J. M. J.; et al. Decacyclene Triimides: Paving the Road to (231) Walker, B.; Han, X.; Kim, C.; Sellinger, A.; Nguyen, T. Q.
Universal Non-Fullerene Acceptors for Organic Photovoltaics. Adv. Solution-Processed Organic Solar Cells from Dye Molecules: An
Energy Mater. 2014, 4, 1301007. Investigation of Diketopyrrolopyrrole: Vinazene Heterojunctions. ACS
(215) Lu, R.-Q.; Zheng, Y.-Q.; Zhou, Y.-N.; Yan, X.-Y.; Lei, T.; Shi, Appl. Mater. Interfaces 2012, 4, 244−250.
K.; Zhou, Y.; Pei, J.; Zoppi, L.; Baldridge, K. K.; et al. Corannulene (232) Holcombe, T. W.; Norton, J. E.; Rivnay, J.; Woo, C. H.; Goris,
Derivatives as Non-Fullerene Acceptors in Solution-Processed Bulk L.; Piliego, C.; Griffini, G.; Sellinger, A.; Brédas, J. L.; Salleo, A.; et al.
Heterojunction Solar Cells. J. Mater. Chem. A 2014, 2, 20515−20519. Steric Control of the Donor/acceptor Interface: Implications in
(216) Chen, H.-Y.; Golder, J.; Yeh, S.-C.; Lin, C.-W.; Chen, C.-T.; Organic Photovoltaic Charge Generation. J. Am. Chem. Soc. 2011, 133,
Chen, C.-T. Diindeno[1,2-g:1′,2′-S]rubicene: All-Carbon Non-Full- 12106−12114.
erene Electron Acceptor for Efficient Bulk-Heterojunction Organic (233) Bin, H.; Yang, Y.; Zhang, Z. G.; Ye, L.; Ghasemi, M.; Chen, S.;
Solar Cells with High Open-Circuit Voltage. RSC Adv. 2015, 5, 3381− Zhang, Y.; Zhang, C.; Sun, C.; Xue, L.; et al. 9.73% Efficiency
3385. Nonfullerene All Organic Small Molecule Solar Cells with Absorption-

3501 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Complementary Donor and Acceptor. J. Am. Chem. Soc. 2017, 139, (252) Cheng, P.; Yan, C.; Wu, Y.; Wang, J.; Qin, M.; An, Q.; Cao, J.;
5085−5094. Huo, L.; Zhang, F.; Ding, L.; et al. Alloy Acceptor: Superior Alternative
(234) Qiu, B.; Xue, L.; Yang, Y.; Bin, H.; Zhang, Y.; Zhang, C.; Xiao, to PCBM toward Efficient and Stable Organic Solar Cells. Adv. Mater.
M.; Park, K.; Morrison, W.; Zhang, Z. G.; et al. All-Small-Molecule 2016, 28, 8021−8028.
Nonfullerene Organic Solar Cells with High Fill Factor and High (253) Liu, T.; Xue, X.; Huo, L.; Sun, X.; An, Q.; Zhang, F.; Russell, T.
Efficiency over 10%. Chem. Mater. 2017, 29, 7543−7553. P.; Liu, F.; Sun, Y. Highly Efficient Parallel-Like Ternary Organic Solar
(235) Yang, L.; Zhang, S.; He, C.; Zhang, J.; Yao, H.; Yang, Y.; Cells. Chem. Mater. 2017, 29, 2914−2920.
Zhang, Y.; Zhao, W.; Hou, J. New Wide Band Gap Donor for Efficient (254) Gobalasingham, N. S.; Noh, S.; Howard, J. B.; Thompson, B.
Fullerene-Free All-Small-Molecule Organic Solar Cells. J. Am. Chem. C. Influence of Surface Energy on Organic Alloy Formation in Ternary
Soc. 2017, 139, 1958−1966. Blend Solar Cells Based on Two Donor Polymers. ACS Appl. Mater.
(236) Feng, G.; Xu, Y.; Zhang, J.; Wang, Z.; Zhou, Y.; Li, Y.; Wei, Z.; Interfaces 2016, 8, 27931−27941.
Li, C.; Li, W. All-Small-Molecule Organic Solar Cells Based on an (255) Jiang, K.; Zhang, G.; Yang, G.; Zhang, J.; Li, Z.; Ma, T.; Hu, H.;
Electron Donor Incorporating Binary Electron-Deficient Units. J. Ma, W.; Yan, H. Multiple Cases of Efficient Non-Fullerene Ternary
Mater. Chem. A 2016, 4, 6056−6063. Organic Solar Cells Enabled by An Effective Morphology Control
(237) Badgujar, S.; Song, C. E.; Oh, S.; Shin, W. S.; Moon, S.-J.; Lee, Method. Adv. Energy Mater. 2017, 1701370.
J.-C.; Jung, I. H.; Lee, S. K. Highly Efficient and Thermally Stable (256) Ko, S.-J.; Lee, W.; Choi, H.; Walker, B.; Yum, S.; Kim, S.; Shin,
Fullerene-Free Organic Solar Cells Based on a Small Molecule Donor T. J.; Woo, H. Y.; Kim, J. Y. Improved Performance in Polymer Solar
and Acceptor. J. Mater. Chem. A 2016, 4, 16335−16340. Cells Using Mixed PC61BM/PC71BM Acceptors. Adv. Energy Mater.
(238) Min, J.; Kwon, O. K.; Cui, C.; Park, J.-H.; Wu, Y.; Park, S. Y.; 2015, 5, 1401687.
Li, Y.; Brabec, C. J. High Performance All-Small-Molecule Solar Cells: (257) Ye, L.; Xu, H. H.; Yu, H.; Xu, W. Y.; Li, H.; Wang, H.; Zhao,
Engineering the Nanomorphology via Processing Additives. J. Mater. N.; Xu, J. Bin. Ternary Bulk Heterojunction Photovoltaic Cells
Chem. A 2016, 4, 14234−14240. Composed of Small Molecule Donor Additive as Cascade Material. J.
(239) Lin, Y.; Wang, J.; Li, T.; Wu, Y.; Wang, C.; Han, L.; Yao, Y.; Phys. Chem. C 2014, 118, 20094−20099.
Ma, W.; Zhan, X. Efficient Fullerene-Free Organic Solar Cells Based (258) Zhang, G.; Zhang, K.; Yin, Q.; Jiang, X. F.; Wang, Z.; Xin, J.;
on Fused-Ring Oligomer Molecules. J. Mater. Chem. A 2016, 4, 1486− Ma, W.; Yan, H.; Huang, F.; Cao, Y. High-Performance Ternary
1494. Organic Solar Cell Enabled by a Thick Active Layer Containing a
(240) Kwon, O. K.; Park, J.-H.; Kim, D. W.; Park, S. K.; Park, S. Y. Liquid Crystalline Small Molecule Donor. J. Am. Chem. Soc. 2017, 139,
An All-Small-Molecule Organic Solar Cell with High Efficiency 2387−2395.
Nonfullerene Acceptor. Adv. Mater. 2015, 27, 1951−1956. (259) Zhang, M.; Wang, J.; Zhang, F.; Mi, Y.; An, Q.; Wang, W.; Ma,
(241) Chen, Y.; Zhang, X.; Zhan, C.; Yao, J. Origin of Effects of X.; Zhang, J.; Liu, X. Ternary Small Molecule Solar Cells Exhibiting
Additive Solvent on Film-Morphology in Solution-Processed Non- Power Conversion Efficiency of 10.3%. Nano Energy 2017, 39, 571−
581.
fullerene Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 6462−6471.
(260) Zhang, M.; Zhang, F.; An, Q.; Sun, Q.; Wang, W.; Ma, X.;
(242) Koppe, M.; Egelhaaf, H.-J.; Dennler, G.; Scharber, M. C.;
Zhang, J.; Tang, W. Nematic Liquid Crystal Materials as a Morphology
Brabec, C. J.; Schilinsky, P.; Hoth, C. N. Near IR Sensitization of
Regulator for Ternary Small Molecule Solar Cells with Power
Organic Bulk Heterojunction Solar Cells: Towards Optimization of
Conversion Efficiency Exceeding 10%. J. Mater. Chem. A 2017, 5,
the Spectral Response of Organic Solar Cells. Adv. Funct. Mater. 2010,
3589−3598.
20, 338−346.
(261) Wang, C.; Zhang, W.; Meng, X.; Bergqvist, J.; Liu, X.; Genene,
(243) Khlyabich, P. P.; Burkhart, B.; Thompson, B. C. Compositional
Z.; Xu, X.; Yartsev, A.; Inganäs, O.; Ma, W.; et al. Ternary Organic
Dependence of the Open-Circuit Voltage in Ternary Blend Bulk Solar Cells with Minimum Voltage Losses. Adv. Energy Mater. 2017, 7,
Heterojunction Solar Cells Based on Two Donor Polymers. J. Am. 1700390.
Chem. Soc. 2012, 134, 9074−9077. (262) Wang, C.; Xu, X.; Zhang, W.; Dkhil, S.; Meng, X.; Liu, X.;
(244) Khlyabich, P. P.; Burkhart, B.; Thompson, B. C. Efficient Margeat, O.; Yartsev, A.; Ma, W.; Ackermann, J.; et al. Ternary
Ternary Blend Bulk Heterojunction Solar Cells with Tunable Open- Organic Solar Cells with Enhanced Open Circuit Voltage. Nano Energy
Circuit Voltage. J. Am. Chem. Soc. 2011, 133, 14534−14537. 2017, 37, 24−31.
(245) Yang, L.; Zhou, H.; Price, S. C.; You, W. Parallel-like Bulk (263) Li, H.; Lu, K.; Wei, Z. Polymer/Small Molecule/Fullerene
Heterojunction Polymer Solar Cells. J. Am. Chem. Soc. 2012, 134, Based Ternary Solar Cells. Adv. Energy Mater. 2017, 7, 1602540.
5432−5435. (264) An, Q.; Zhang, F.; Zhang, J.; Tang, W.; Deng, Z.; Hu, B.
(246) Street, R. A.; Davies, D.; Khlyabich, P. P.; Burkhart, B.; Versatile Ternary Organic Solar Cells: A Critical Review. Energy
Thompson, B. C. Origin of the Tunable Open-Circuit Voltage in Environ. Sci. 2016, 9, 281−322.
Ternary Blend Bulk Heterojunction Organic Solar Cells. J. Am. Chem. (265) Lu, L.; Kelly, M. A.; You, W.; Yu, L. Status and Prospects for
Soc. 2013, 135, 986−989. Ternary Organic Photovoltaics. Nat. Photonics 2015, 9, 491−500.
(247) Mollinger, S. A.; Vandewal, K.; Salleo, A. Microstructural and (266) Lu, L.; Xu, T.; Chen, W.; Landry, E. S.; Yu, L. Ternary Blend
Electronic Origins of Open-Circuit Voltage Tuning in Organic Solar Polymer Solar Cells with Enhanced Power Conversion Efficiency. Nat.
Cells Based on Ternary Blends. Adv. Energy Mater. 2015, 5, 1501335. Photonics 2014, 8, 716−722.
(248) Wang, Z.; Zhang, Y.; Zhang, J.; Wei, Z.; Ma, W. Optimized (267) Gasparini, N.; Jiao, X.; Heumueller, T.; Baran, D.; Matt, G. J.;
“alloy-Parallel” Morphology of Ternary Organic Solar Cells. Adv. Fladischer, S.; Spiecker, E.; Ade, H.; Brabec, C. J.; Ameri, T. Designing
Energy Mater. 2016, 6, 1502456. Ternary Blend Bulk Heterojunction Solar Cells with Reduced Carrier
(249) Khlyabich, P. P.; Rudenko, A. E.; Thompson, B. C.; Loo, Y. L. Recombination and a Fill Factor of 77%. Nat. Energy 2016, 1, 16118.
Structural Origins for Tunable Open-Circuit Voltage in Ternary-Blend (268) Liu, T.; Huo, L.; Sun, X.; Fan, B.; Cai, Y.; Kim, T.; Kim, J. Y.;
Organic Solar Cells. Adv. Funct. Mater. 2015, 25, 5557−5563. Choi, H.; Sun, Y. Ternary Organic Solar Cells Based on Two Highly
(250) Kouijzer, S.; Li, W.; Wienk, M. M.; Janssen, R. A. J. Charge Efficient Polymer Donors with Enhanced Power Conversion
Transfer State Energy in Ternary Bulk-Heterojunction Polymer− Efficiency. Adv. Energy Mater. 2016, 6, 1502109.
fullerene Solar Cells. J. Photonics Energy 2015, 5, 057203. (269) Wang, Y.; Ohkita, H.; Benten, H.; Ito, S. Highly Efficient
(251) Cowan, S. R.; Leong, W. L.; Banerji, N.; Dennler, G.; Heeger, Exciton Harvesting and Charge Transport in Ternary Blend Solar
A. J. Identifying a Threshold Impurity Level for Organic Solar Cells: Cells Based on Wide- and Low-Bandgap Polymers. Phys. Chem. Chem.
Enhanced First-Order Recombination via Well-Defined PC84BM Phys. 2015, 17, 27217−27224.
Traps in Organic Bulk Heterojunction Solar Cells. Adv. Funct. Mater. (270) Ameri, T.; Khoram, P.; Heumüller, T.; Baran, D.; Machui, F.;
2011, 21, 3083−3092. Troeger, A.; Sgobba, V.; Guldi, D. M.; Halik, M.; Rathgeber, S.; et al.

3502 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Morphology Analysis of near IR Sensitized Polymer/fullerene Organic (289) Distler, A.; Sauermann, T.; Egelhaaf, H. J.; Rodman, S.; Waller,
Solar Cells by Implementing Low Bandgap Heteroanalogue C-/Si- D.; Cheon, K. S.; Lee, M.; Guldi, D. M. The Effect of PCBM
PCPDTBT. J. Mater. Chem. A 2014, 2, 19461−19472. Dimerization on the Performance of Bulk Heterojunction Solar Cells.
(271) Machui, F.; Rathgeber, S.; Li, N.; Ameri, T.; Brabec, C. J. Adv. Energy Mater. 2014, 4, 1300693.
Influence of a Ternary Donor Material on the Morphology of a (290) Inasaridze, L. N.; Shames, A. I.; Martynov, I. V.; Li, B.;
P3HT:PCBM Blend for Organic Photovoltaic Devices. J. Mater. Chem. Mumyatov, A. V.; Susarova, D. K.; Katz, E. A.; Troshin, P. A. Light-
2012, 22, 15570. Induced Generation of Free Radicals by Fullerene Derivatives: An
(272) An, Q.; Zhang, F.; Yin, X.; Sun, Q.; Zhang, M.; Zhang, J.; Tang, Important Degradation Pathway in Organic Photovoltaics? J. Mater.
W.; Deng, Z. High-Performance Alloy Model-Based Ternary Small Chem. A 2017, 5, 8044−8050.
Molecule Solar Cells. Nano Energy 2016, 30, 276−282. (291) Heumueller, T.; Mateker, W. R.; Distler, A.; Fritze, U. F.;
(273) Nian, L.; Gao, K.; Jiang, Y.; Rong, Q.; Hu, X.; Yuan, D.; Liu, F.; Cheacharoen, R.; Nguyen, W. H.; Biele, M.; Salvador, M.; von Delius,
Peng, X.; Russell, T. P.; Zhou, G. Small-Molecule Solar Cells with M.; Egelhaaf, H.-J.; et al. Morphological and Electrical Control of
Simultaneously Enhanced Short-Circuit Current and Fill Factor to Fullerene Dimerization Determines Organic Photovoltaic Stability.
Achieve 11% Efficiency. Adv. Mater. 2017, 29, 1700616. Energy Environ. Sci. 2016, 9, 247−256.
(274) Cheng, P.; Zhan, X. Versatile Third Components for Efficient (292) Liu, F.; Zhou, Z.; Zhang, C.; Zhang, J.; Hu, Q.; Vergote, T.;
and Stable Organic Solar Cells. Mater. Horiz. 2015, 2, 462−485. Liu, F.; Russell, T. P.; Zhu, X. Efficient Semitransparent Solar Cells
(275) Liu, T.; Huo, L.; Sun, X.; Fan, B.; Cai, Y.; Kim, T.; Kim, J. Y.; with High NIR Responsiveness Enabled by a Small-Bandgap Electron
Choi, H.; Sun, Y. Ternary Organic Solar Cells Based on Two Highly Acceptor. Adv. Mater. 2017, 29, 1606574.
Efficient Polymer Donors with Enhanced Power Conversion (293) Wang, W.; Yan, C.; Lau, T.-K.; Wang, J.; Liu, K.; Fan, Y.; Lu,
Efficiency. Adv. Energy Mater. 2016, 6, 10008−10015. X.; Zhan, X. Fused Hexacyclic Nonfullerene Acceptor with Strong
(276) Yu, R.; Zhang, S.; Yao, H.; Guo, B.; Li, S.; Zhang, H.; Zhang, Near-Infrared Absorption for Semitransparent Organic Solar Cells
M.; Hou, J. Two Well-Miscible Acceptors Work as One for Efficient with 9.77% Efficiency. Adv. Mater. 2017, 29, 1701308.
Fullerene-Free Organic Solar Cells. Adv. Mater. 2017, 29, 1700437. (294) Min, J.; Luponosov, Y. N.; Cui, C.; Kan, B.; Chen, H.; Wan, X.;
(277) Su, W.; Fan, Q.; Guo, X.; Meng, X.; Bi, Z.; Ma, W.; Zhang, M.; Chen, Y.; Ponomarenko, S. A.; Li, Y.; Brabec, C. J. Evaluation of
Li, Y. Two Compatible Nonfullerene Acceptors with Similar Structures Electron Donor Materials for Solution-Processed Organic Solar Cells
as Alloy for Efficient Ternary Polymer Solar Cells. Nano Energy 2017, via a Novel Figure of Merit. Adv. Energy Mater. 2017, 7, 1700465.
38, 510−517. (295) Liu, Y.; Zhao, J.; Li, Z.; Mu, C.; Ma, W.; Hu, H.; Jiang, K.; Lin,
(278) Gu, W.; Yang, L.; Chen, Y.; Wang, X.; Zhang, H.; Hou, J.; Zhu, H.; Ade, H.; Yan, H. Aggregation and Morphology Control Enables
Z.; Huang, H. Ternary Blend Polymer Solar Cells with Two Non-
Multiple Cases of High-Efficiency Polymer Solar Cells. Nat. Commun.
Fullerene Acceptors as Acceptor Alloy. Dyes Pigm. 2017, 141, 388−
2014, 5, 5293.
393.
(296) Rand, B. P.; Genoe, J.; Heremans, P.; Poortmans, J. Solar Cells
(279) Ding, G.; Yuan, J.; Jin, F.; Zhang, Y.; Han, L.; Ling, X.; Zhao,
Utilizing Small Molecular Weight Organic Semiconductors. Prog.
H.; Ma, W. High-Performance All-Polymer Nonfullerene Solar Cells
Photovoltaics 2007, 15, 659−676.
by Employing an Efficient Polymer-Small Molecule Acceptor Alloy
(297) Yao, J.; Kirchartz, T.; Vezie, M. S.; Faist, M. A.; Gong, W.; He,
Strategy. Nano Energy 2017, 36, 356−365.
Z.; Wu, H.; Troughton, J.; Watson, T.; Bryant, D.; et al. Quantifying
(280) Zhong, L.; Gao, L.; Bin, H.; Hu, Q.; Zhang, Z. G.; Liu, F.;
Russell, T. P.; Zhang, Z.; Li, Y. High Efficiency Ternary Nonfullerene Losses in Open-Circuit Voltage in Solution-Processable Solar Cells.
Polymer Solar Cells with Two Polymer Donors and an Organic Phys. Rev. Appl. 2015, 4, 14020.
Semiconductor Acceptor. Adv. Energy Mater. 2017, 7, 1602215. (298) Rau, U. Reciprocity Relation between Photovoltaic Quantum
(281) Chen, Y.; Qin, Y.; Wu, Y.; Li, C.; Yao, H.; Liang, N.; Wang, X.; Efficiency and Electroluminescent Emission of Solar Cells. Phys. Rev.
Li, W.; Ma, W.; Hou, J. From Binary to Ternary: Improving the B: Condens. Matter Mater. Phys. 2007, 76, 85303.
External Quantum Efficiency of Small-Molecule Acceptor-Based (299) Kaienburg, P.; Rau, U.; Kirchartz, T. Extracting Information
Polymer Solar Cells with a Minute Amount of Fullerene Sensitization. about the Electronic Quality of Organic Solar-Cell Absorbers from Fill
Adv. Energy Mater. 2017, 7, 1700328. Factor and Thickness. Phys. Rev. Appl. 2016, 6, 24001.
(282) Zhao, W.; Li, S.; Zhang, S.; Liu, X.; Hou, J. Ternary Polymer (300) Shockley, W.; Queisser, H. J. Detailed Balance Limit of
Solar Cells Based on Two Acceptors and One Donor for Achieving Efficiency of P-n Junction Solar Cells. J. Appl. Phys. 1961, 32, 510−
12.2% Efficiency. Adv. Mater. 2017, 29, 1604059. 519.
(283) Fan, B.; Zhong, W.; Jiang, X.-F.; Yin, Q.; Ying, L.; Huang, F.; (301) Vandewal, K.; Tvingstedt, K.; Gadisa, A.; Inganäs, O.; Manca, J.
Cao, Y. Improved Performance of Ternary Polymer Solar Cells Based V. On the Origin of the Open-Circuit Voltage of Polymer-Fullerene
on A Nonfullerene Electron Cascade Acceptor. Adv. Energy Mater. Solar Cells. Nat. Mater. 2009, 8, 904−909.
2017, 7, 1602127. (302) Vandewal, K.; Tvingstedt, K.; Gadisa, A.; Inganäs, O.; Manca, J.
(284) Chen, Y.; Ye, P.; Zhu, Z. G.; Wang, X.; Yang, L.; Xu, X.; Wu, V. Relating the Open-Circuit Voltage to Interface Molecular Properties
X.; Dong, T.; Zhang, H.; Hou, J.; et al. Achieving High-Performance of Donor:acceptor Bulk Heterojunction Solar Cells. Phys. Rev. B:
Ternary Organic Solar Cells through Tuning Acceptor Alloy. Adv. Condens. Matter Mater. Phys. 2010, 81, 125204.
Mater. 2017, 29, 1603154. (303) Benduhn, J.; Tvingstedt, K.; Piersimoni, F.; Ullbrich, S.; Fan,
(285) Cheng, P.; Zhan, X. Stability of Organic Solar Cells: Y.; Tropiano, M.; McGarry, K. A.; Zeika, O.; Riede, M. K.; Douglas, C.
Challenges and Strategies. Chem. Soc. Rev. 2016, 45, 2544−2582. J.; et al. Intrinsic Non-Radiative Voltage Losses in Fullerene-Based
(286) Soon, Y. W.; Cho, H.; Low, J.; Bronstein, H.; McCulloch, I.; Organic Solar Cells. Nat. Energy 2017, 2, 17053.
Durrant, J. R. Correlating Triplet Yield, Singlet Oxygen Generation (304) Morteani, A. C.; Sreearunothai, P.; Herz, L. M.; Friend, R. H.;
and Photochemical Stability in Polymer/fullerene Blend Films. Chem. Silva, C. Exciton Regeneration at Polymeric Semiconductor Hetero-
Commun. 2013, 49, 1291. junctions. Phys. Rev. Lett. 2004, 92, 4.
(287) Razzell-Hollis, J.; Wade, J.; Tsoi, W. C.; Soon, Y.; Durrant, J.; (305) Scharber, M. C.; Sariciftci, N. S. Efficiency of Bulk-
Kim, J.-S. Photochemical Stability of High Efficiency PTB7:PC70BM Heterojunction Organic Solar Cells. Prog. Polym. Sci. 2013, 38,
Solar Cell Blends. J. Mater. Chem. A 2014, 2, 20189−20195. 1929−1940.
(288) Heumueller, T.; Mateker, W. R.; Sachs-Quintana, I. T.; (306) Chen, K.; Barker, A. J.; Reish, M. E.; Gordon, K. C.; Hodgkiss,
Vandewal, K.; Bartelt, J. A.; Burke, T. M.; Ameri, T.; Brabec, C. J.; J. M. Broadband Ultrafast Photoluminescence Spectroscopy Resolves
McGehee, M. D. Reducing Burn-in Voltage Loss in Polymer Solar Charge Photogeneration via Delocalized Hot Excitons in polymer:-
Cells by Increasing the Polymer Crystallinity. Energy Environ. Sci. Fullerene Photovoltaic Blends. J. Am. Chem. Soc. 2013, 135, 18502−
2014, 7, 2974−2980. 18512.

3503 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

(307) Liu, X.; Ding, K.; Panda, A.; Forrest, S. R. Charge Transfer Extended Tetrathiafulvalene Analogues with Fused Thiophene Units
States in Dilute Donor-Acceptor Blend Organic Heterojunctions. ACS as π-Conjugated Spacers. J. Mater. Chem. 2003, 13, 1324−1332.
Nano 2016, 10, 7619−7626. (325) Kawabata, K.; Takeguchi, M.; Goto, H. Optical Activity of
(308) Kirchartz, T.; Taretto, K.; Rau, U. Efficiency Limits of Organic Heteroaromatic Conjugated Polymer Films Prepared by Asymmetric
Bulk Heterojunction Solar Cells. J. Phys. Chem. C 2009, 113, 17958− Electrochemical Polymerization in Cholesteric Liquid Crystals:
17966. Structural Function for Chiral Induction. Macromolecules 2013, 46,
(309) Yi, Y.; Coropceanu, V.; Brédas, J. L. Exciton-Dissociation and 2078−2091.
Charge-Recombination Processes in Pentacene/C60 Solar Cells: (326) Liu, J.; Li, K.; Liu, B. Far-Red/Near-Infrared Conjugated
Theoretical Insight into the Impact of Interface Geometry. J. Am. Polymer Nanoparticles for Long-Term In Situ Monitoring of Liver
Chem. Soc. 2009, 131, 15777−15783. Tumor Growth. Adv. Sci. 2015, 2, 1500008.
(310) Chen, X.-K.; Ravva, M. K.; Li, H.; Ryno, S. M.; Brédas, J.-L. (327) Cui, Y.; Ren, H.; Yu, J.; Wang, Z.; Qian, G. An Indanone-Based
Effect of Molecular Packing and Charge Delocalization on the Alkoxysilane Dye with Second Order Nonlinear Optical Properties.
Nonradiative Recombination of Charge-Transfer States in Organic Dyes Pigm. 2009, 81, 53−57.
Solar Cells. Adv. Energy Mater. 2016, 6, 1601325. (328) Archet, F.; Yao, D.; Chambon, S.; Abbas, M.; D'Aleo, A.;
(311) Yang, B.; Yi, Y.; Zhang, C. R.; Aziz, S. G.; Coropceanu, V.; Canard, G.; Ponce-Vargas, M.; Zaborova, E.; Le Guennic, B.; Wantz,
Brédas, J. L. Impact of Electron Delocalization on the Nature of the G.; et al. Synthesis of Bioinspired Curcuminoid Small Molecules for
Charge-Transfer States in Model pentacene/C60 Interfaces: A Density Solution-Processed Organic Solar Cells with High Open-Circuit
Functional Theory Study. J. Phys. Chem. C 2014, 118, 27648−27656. Voltage. ACS Energy Lett. 2017, 2, 1303−1307.
(312) Hörmann, U.; Lorch, C.; Hinderhofer, A.; Gerlach, A.; Gruber, (329) McAfee, S. M.; Dayneko, S. V.; Josse, P.; Blanchard, P.;
M.; Kraus, J.; Sykora, B.; Grob, S.; Linderl, T.; Wilke, A.; et al. VOC Cabanetos, C.; Welch, G. C. Simply Complex: The Efficient Synthesis
from a Morphology Point of View: The Influence of Molecular of an Intricate Molecular Acceptor for High-Performance Air-
Orientation on the Open Circuit Voltage of Organic Planar Processed and Air-Tested Fullerene-Free Organic Solar Cells. Chem.
Heterojunction Solar Cells. J. Phys. Chem. C 2014, 118, 26462−26470. Mater. 2017, 29, 1309−1314.
(313) Rand, B. P.; Cheyns, D.; Vasseur, K.; Giebink, N. C.; Mothy, (330) Firdaus, Y.; Maffei, L. P.; Cruciani, F.; Müller, M. A.; Liu, S.;
S.; Yi, Y.; Coropceanu, V.; Beljonne, D.; Cornil, J.; Brédas, J. L.; et al. Lopatin, S.; Wehbe, N.; Ndjawa, G. O. N.; Amassian, A.; Laquai, F.;
The Impact of Molecular Orientation on the Photovoltaic Properties et al. Polymer Main-Chain Substitution Effects on the Efficiency of
of a Phthalocyanine/fullerene Heterojunction. Adv. Funct. Mater. 2012, Nonfullerene BHJ Solar Cells. Adv. Energy Mater. 2017, 7, 1700834.
22, 2987−2995. (331) Min, J.; Cui, C.; Heumueller, T.; Fladischer, S.; Cheng, X.;
(314) Ran, N. A.; Roland, S.; Love, J. A.; Savikhin, V.; Takacs, C. J.; Spiecker, E.; Li, Y.; Brabec, C. J. Side-Chain Engineering for
Fu, Y. T.; Li, H.; Coropceanu, V.; Liu, X.; Brédas, J. L.; et al. Impact of Enhancing the Properties of Small Molecule Solar Cells: A Trade-off
Interfacial Molecular Orientation on Radiative Recombination and Beyond Efficiency. Adv. Energy Mater. 2016, 6, 1600515.
Charge Generation Efficiency. Nat. Commun. 2017, 8, 79. (332) Stephen, M.; Genevičius, K.; Juška, G.; Arlauskas, K.; Hiorns,
(315) Tumbleston, J. R.; Collins, B. A.; Yang, L.; Stuart, A. C.; Gann,
R. C. Charge Transport and Its Characterization Using Photo-CELIV
E.; Ma, W.; You, W.; Ade, H. The Influence of Molecular Orientation
in Bulk Heterojunction Solar Cells. Polym. Int. 2017, 66, 13−25.
on Organic Bulk Heterojunction Solar Cells. Nat. Photonics 2014, 8,
(333) Aguirre, J. C.; Arntsen, C.; Hernandez, S.; Huber, R.; Nardes,
385−391.
A. M.; Halim, M.; Kilbride, D.; Kopidakis, N.; Tolbert, S. H.; Schwartz,
(316) Anctil, A.; Babbitt, C. W.; Raffaelle, R. P.; Landi, B. J. Material
B. J.; et al. Understanding Local and Macroscopic Electron Mobilities
and Energy Intensity of Fullerene Production. Environ. Sci. Technol.
in the Fullerene Network of Conjugated Polymer-Based Solar Cells:
2011, 45, 2353−2359.
(317) Po, R.; Bianchi, G.; Carbonera, C.; Pellegrino, A. All That Time-Resolved Microwave Conductivity and Theory. Adv. Funct.
Glisters Is Not Gold”: An Analysis of the Synthetic Complexity of Mater. 2014, 24, 784−792.
Efficient Polymer Donors for Polymer Solar Cells. Macromolecules (334) Savenije, T. J.; Ferguson, A. J.; Kopidakis, N.; Rumbles, G.
2015, 48, 453−461. Revealing the Dynamics of Charge Carriers in Polymer:fullerene
(318) Po, R.; Roncali, J. Beyond Efficiency: Scalability of Molecular Blends Using Photoinduced Time-Resolved Microwave Conductivity.
Donor Materials for Organic Photovoltaics. J. Mater. Chem. C 2016, 4, J. Phys. Chem. C 2013, 117, 24085−24103.
3677−3685. (335) Zhang, G.; Huber, R. C.; Ferreira, A. S.; Boyd, S. D.;
(319) Zhang, Y. B.; Furukawa, H.; Ko, N.; Nie, W.; Park, H. J.; Luscombe, C. K.; Tolbert, S. H.; Schwartz, B. J. Crystallinity Effects in
Okajima, S.; Cordova, K. E.; Deng, H.; Kim, J.; Yaghi, O. M. Sequentially Processed and Blend-Cast Bulk-Heterojunction Polymer/
Introduction of Functionality, Selection of Topology, and Enhance- Fullerene Photovoltaics. J. Phys. Chem. C 2014, 118, 18424−18435.
ment of Gas Adsorption in Multivariate Metal-Organic Framework- (336) Paquin, F.; Yamagata, H.; Hestand, N. J.; Sakowicz, M.;
177. J. Am. Chem. Soc. 2015, 137, 2641−2650. Bérubé, N.; Côté, M.; Reynolds, L. X.; Haque, S. A.; Stingelin, N.;
(320) Hu, Y.; Chen, S.; Zhang, L.; Zhang, Y.; Yuan, Z.; Zhao, X.; Spano, F. C.; et al. Two-Dimensional Spatial Coherence of Excitons in
Chen, Y. Facile Approach to Perylenemonoimide with Short Side Semicrystalline Polymeric Semiconductors: Effect of Molecular
Chains for Nonfullerene Solar Cells. J. Org. Chem. 2017, 82, 5926− Weight. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 155202.
5931. (337) Bronstein, H. A.; Luscombe, C. K. Externally Initiated
(321) Manning, S. J.; Bogen, W.; Kelly, L. A. Synthesis, Character- Regioregular P3HT with Controlled Molecular Weight and Narrow
ization, and Photophysical Study of Fluorescent N-Substituted Polydispersity. J. Am. Chem. Soc. 2009, 131, 12894−12895.
Benzo[ghi]perylene “swallow Tail” monoimides. J. Org. Chem. 2011, (338) Scharsich, C.; Lohwasser, R. H.; Sommer, M.; Asawapirom, U.;
76, 6007−6013. Scherf, U.; Thelakkat, M.; Neher, D.; Köhler, A. Control of Aggregate
(322) Xu, Y. X.; Chueh, C. C.; Yip, H. L.; Ding, F. Z.; Li, Y. X.; Li, C. Formation in poly(3-Hexylthiophene) by Solvent, Molecular Weight,
Z.; Li, X.; Chen, W. C.; Jen, A. K. Y. Improved Charge Transport and and Synthetic Method. J. Polym. Sci., Part B: Polym. Phys. 2012, 50,
Absorption Coefficient in indacenodithieno[3,2-B]thiophene-Based 442−453.
Ladder-Type Polymer Leading to Highly Efficient Polymer Solar Cells. (339) Woo, C. H.; Thompson, B. C.; Kim, B. J.; Toney, M. F.;
Adv. Mater. 2012, 24, 6356−6361. Fréc het, J. M. J. The Influence of poly(3-Hexylthiophene)
(323) Chen, Y.-C.; Yu, C.-Y.; Fan, Y.-L.; Hung, L.-I.; Chen, C.-P.; Regioregularity on Fullerene-Composite Solar Cell Performance. J.
Ting, C. Low-Bandgap Conjugated Polymer for High Efficient Am. Chem. Soc. 2008, 130, 16324−16329.
Photovoltaic Applications. Chem. Commun. 2010, 46, 6503. (340) Mauer, R.; Kastler, M.; Laquai, F. The Impact of Polymer
(324) Leriche, P.; Raimundo, J.-M.; Turbiez, M.; Monroche, V.; Regioregularity on Charge Transport and Efficiency of P3HT:PCBM
Allain, M.; Sauvage, F.-X.; Roncali, J.; Frère, P.; Skabara, P. J. Linearly Photovoltaic Devices. Adv. Funct. Mater. 2010, 20, 2085−2092.

3504 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

(341) Adachi, T.; Brazard, J.; Ono, R. J.; Hanson, B.; Traub, M. C.; Solution Printing of All-Polymer Solar Cells. Nat. Commun. 2015, 6,
Wu, Z.; Li, Z.; Bolinger, J. C.; Ganesan, V.; Bielawski, C. W.; et al. 7955.
Regioregularity and Single Polythiophene Chain Conformation. J. (359) Li, W.; Yang, L.; Tumbleston, J. R.; Yan, L.; Ade, H.; You, W.
Phys. Chem. Lett. 2011, 2, 1400−1404. Controlling Molecular Weight of a High Efficiency Donor-Acceptor
(342) Kim, Y.; Cook, S.; Tuladhar, S. M.; Choulis, S. A.; Nelson, J.; Conjugated Polymer and Understanding Its Significant Impact on
Durrant, J. R.; Bradley, D. D. C.; Giles, M.; McCulloch, I.; Ha, C.-S.; Photovoltaic Properties. Adv. Mater. 2014, 26, 4456−4462.
et al. A Strong Regioregularity Effect in Self-Organizing Conjugated (360) Yan, H.; Wang, C.; Garcia, A.; Swaraj, S.; Gu, Z.; McNeill, C.
Polymer Films and High-Efficiency Polythiophene:fullerene Solar R.; Schuettfort, T.; Sohn, K. E.; Kramer, E. J.; Bazan, G. C.; et al.
Cells. Nat. Mater. 2006, 5, 197−203. Interfaces in Organic Devices Studied with Resonant Soft X-Ray
(343) Kohn, P.; Rong, Z.; Scherer, K. H.; Sepe, A.; Sommer, M.; Reflectivity. J. Appl. Phys. 2011, 110, 102220.
Müller-Buschbaum, P.; Friend, R. H.; Steiner, U.; Hüttner, S. (361) Collins, B. A.; Li, Z.; Tumbleston, J. R.; Gann, E.; McNeill, C.
Crystallization-Induced 10-Nm Structure Formation in P3HT/ R.; Ade, H. Absolute Measurement of Domain Composition and
PCBM Blends. Macromolecules 2013, 46, 4002−4013. Nanoscale Size Distribution Explains Performance in PTB7:PC71BM
(344) Collins, B. A.; Gann, E.; Guignard, L.; He, X.; McNeill, C. R.; Solar Cells. Adv. Energy Mater. 2013, 3, 65−74.
Ade, H. Molecular Miscibility of Polymer−Fullerene Blends. J. Phys. (362) Hawks, S. A.; Aguirre, J. C.; Schelhas, L. T.; Thompson, R. J.;
Chem. Lett. 2010, 1, 3160−3166. Huber, R. C.; Ferreira, A. S.; Zhang, G.; Herzing, A. A.; Tolbert, S. H.;
(345) Gadisa, A.; Tumbleston, J. R.; Ko, D.-H.; Aryal, M.; Lopez, R.; Schwartz, B. J. Comparing Matched Polymer:Fullerene Solar Cells
Samulski, E. T. The Role of Solvent and Morphology on Miscibility of Made by Solution-Sequential Processing and Traditional Blend
Methanofullerene and poly(3-Hexylthiophene). Thin Solid Films 2012, Casting: Nanoscale Structure and Device Performance. J. Phys.
520, 5466−5471. Chem. C 2014, 118, 17413−17425.
(346) Treat, N. D.; Brady, M. A.; Smith, G.; Toney, M. F.; Kramer, E. (363) Westacott, P.; Tumbleston, J. R.; Shoaee, S.; Fearn, S.;
J.; Hawker, C. J.; Chabinyc, M. L. Interdiffusion of PCBM and P3HT Bannock, J. H.; Gilchrist, J. B.; Heutz, S.; DeMello, J.; Heeney, M.;
Reveals Miscibility in a Photovoltaically Active Blend. Adv. Energy Ade, H.; et al. On the Role of Intermixed Phases in Organic
Mater. 2011, 1, 82−89. Photovoltaic Blends. Energy Environ. Sci. 2013, 6, 2756.
(347) Machui, F.; Brabec, C. J. Solubility, Miscibility, and the Impact (364) Richter, L. J.; Delongchamp, D. M.; Amassian, A. Morphology
on Solid-State Morphology. In Semiconducting Polymer Composites; Development in Solution-Processed Functional Organic Blend Films:
Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2013; An in Situ Viewpoint. Chem. Rev. 2017, 117, 6332−6366.
pp 1−38. (365) Jiang, X.; Xu, Y.; Wang, X.; Wu, Y.; Feng, G.; Li, C.; Ma, W.;
(348) Chen, H.; Hegde, R.; Browning, J.; Dadmun, M. D. The Li, W. Non-Fullerene Organic Solar Cells Based on Diketopyrrolo-
Miscibility and Depth Profile of PCBM in P3HT: Thermodynamic pyrrole Polymers as Electron Donors and ITIC as an Electron
Information to Improve Organic Photovoltaics. Phys. Chem. Chem. Acceptor. Phys. Chem. Chem. Phys. 2017, 19, 8069−8075.
Phys. 2012, 14, 5635−5641. (366) Ye, L.; Zhao, W.; Li, S.; Mukherjee, S.; Carpenter, J. H.;
(349) Collins, B. A.; Tumbleston, J. R.; Ade, H. Miscibility, Awartani, O.; Jiao, X.; Hou, J.; Ade, H. High-Efficiency Nonfullerene
Crystallinity, and Phase Development in P3HT/PCBM Solar Cells: Organic Solar Cells: Critical Factors That Affect Complex Multi-
Toward an Enlightened Understanding of Device Morphology and Length Scale Morphology and Device Performance. Adv. Energy Mater.
Stability. J. Phys. Chem. Lett. 2011, 2, 3135−3145. 2017, 7, 1602000.
(350) Collins, B. a.; Gann, E.; Guignard, L.; He, X.; McNeill, C. R.; (367) Hu, H.; Jiang, K.; Chow, P. C. Y.; Ye, L.; Zhang, G.; Li, Z.;
Ade, H. Molecular Miscibility of Polymer−Fullerene Blends. J. Phys. Carpenter, J. H.; Ade, H.; Yan, H. Influence of Donor Polymer on the
Chem. Lett. 2010, 1, 3160−3166. Molecular Ordering of Small Molecular Acceptors in Nonfullerene
(351) Ghasemi, M.; Ye, L.; Zhang, Q.; Yan, L.; Kim, J. H.; Awartani, Polymer Solar Cells. Adv. Energy Mater. 2017, 1701674.
O.; You, W.; Gadisa, A.; Ade, H. Panchromatic Sequentially Cast (368) Flory, P. J. Principles of Polymer Chemistry; Cornell University
Ternary Polymer Solar Cells. Adv. Mater. 2017, 29, 1604603. Press: Ithaca, NY, 1953.
(352) Li, N.; Perea, J. D.; Kassar, T.; Richter, M.; Heumueller, T.; (369) Flory, P. J. Statistical Mechanics of Chain Molecules; Wiley
Matt, G. J.; Hou, Y.; Güldal, N. S.; Chen, H.; Chen, S.; et al. Abnormal (Interscience): New York, 1969.
Strong Burn-in Degradation of Highly Efficient Polymer Solar Cells (370) Jiao, X.; Ye, L.; Ade, H. Quantitative Morphology-Performance
Caused by Spinodal Donor-Acceptor Demixing. Nat. Commun. 2017, Correlations in Organic Solar Cells: Insights from Soft X-Ray
8, 14541. Scattering. Adv. Energy Mater. 2017, 7, 1700084.
(353) Kniepert, J.; Lange, I.; Heidbrink, J.; Kurpiers, J.; Brenner, T. J. (371) Shao, M.; Keum, J. K.; Kumar, R.; Chen, J.; Browning, J. F.;
K.; Koster, L. J. A.; Neher, D. Effect of Solvent Additive on Das, S.; Chen, W.; Hou, J.; Do, C.; Littrell, K. C.; et al. Understanding
Generation, Recombination, and Extraction in PTB7:PCBM Solar How Processing Additives Tune the Nanoscale Morphology of High
Cells: A Conclusive Experimental and Numerical Simulation Study. J. Efficiency Organic Photovoltaic Blends: From Casting Solution to
Phys. Chem. C 2015, 119, 8310−8320. Spun-Cast Thin Film. Adv. Funct. Mater. 2014, 24, 6647−6657.
(354) Sharenko, A.; Gehrig, D.; Laquai, F.; Nguyen, T. Q. The Effect (372) Collins, B. A.; Tumbleston, J. R.; Ade, H. Miscibility,
of Solvent Additive on the Charge Generation and Photovoltaic Crystallinity, and Phase Development in P3HT/PCBM Solar Cells:
Performance of a Solution-Processed Small Molecule: Perylene Toward an Enlightened Understanding of Device Morphology and
Diimide Bulk Heterojunction Solar Cell. Chem. Mater. 2014, 26, Stability. J. Phys. Chem. Lett. 2011, 2, 3135−3145.
4109−4118. (373) Kozub, D. R.; Vakhshouri, K.; Orme, L. M.; Wang, C.;
(355) Liao, H. C.; Ho, C. C.; Chang, C. Y.; Jao, M. H.; Darling, S. B.; Hexemer, A.; Gomez, E. D. Polymer Crystallization of Partially
Su, W. F. Additives for Morphology Control in High-Efficiency Miscible Polythiophene/Fullerene Mixtures Controls Morphology.
Organic Solar Cells. Mater. Today 2013, 16, 326−336. Macromolecules 2011, 44, 5722−5726.
(356) Lou, S. J.; Szarko, J. M.; Xu, T.; Yu, L.; Marks, T. J.; Chen, L. (374) Chang, L.; Jacobs, I. E.; Augustine, M. P.; Moulé, A. J.
X. Effects of Additives on the Morphology of Solution Phase Correlating Dilute Solvent Interactions to Morphology and OPV
Aggregates Formed by Active Layer Components of High-Efficiency Device Performance. Org. Electron. 2013, 14, 2431−2443.
Organic Solar Cells. J. Am. Chem. Soc. 2011, 133, 20661−20663. (375) Leman, D.; Kelly, M. A.; Ness, S.; Engmann, S.; Herzing, A.;
(357) Zhou, H.; Yang, L.; You, W. Rational Design of High Snyder, C.; Ro, H. W.; Kline, R. J.; DeLongchamp, D. M.; Richter, L. J.
Performance Conjugated Polymers for Organic Solar Cells. Macro- In Situ Characterization of Polymer−Fullerene Bilayer Stability.
molecules 2012, 45, 607−632. Macromolecules 2015, 48, 383−392.
(358) Diao, Y.; Zhou, Y.; Kurosawa, T.; Shaw, L.; Wang, C.; Park, S.; (376) Dowland, S. A.; Salvador, M.; Perea, J. D.; Gasparini, N.;
Guo, Y.; Reinspach, J. A.; Gu, K.; Gu, X.; et al. Flow-Enhanced Langner, S.; Rajoelson, S.; Ramanitra, H. H.; Lindner, B. D.; Osvet, A.;

3505 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

Brabec, C. J.; et al. Suppression of Thermally Induced Fullerene (393) Ibrahem, M. A.; Wei, H.-Y.; Tsai, M.-H.; Ho, K.-C.; Shyue, J.-
Aggregation in Polyfullerene-Based Multiacceptor Organic Solar Cells. J.; Chu, C. W. Solution-Processed Zinc Oxide Nanoparticles as
ACS Appl. Mater. Interfaces 2017, 9, 10971−10982. Interlayer Materials for Inverted Organic Solar Cells. Sol. Energy Mater.
(377) Ro, H. W.; Akgun, B.; O’Connor, B. T.; Hammond, M.; Kline, Sol. Cells 2013, 108, 156−163.
R. J.; Snyder, C. R.; Satija, S. K.; Ayzner, A. L.; Toney, M. F.; Soles, C. (394) Mauger, S. A.; Chang, L.; Friedrich, S.; Rochester, C. W.;
L.; et al. Poly(3-Hexylthiophene) and [6,6]-Phenyl-C61-Butyric Acid Huang, D. M.; Wang, P.; Moulé, A. J. Self-Assembly of Selective
Methyl Ester Mixing in Organic Solar Cells. Macromolecules 2012, 45, Interfaces in Organic Photovoltaics. Adv. Funct. Mater. 2013, 23,
6587−6599. 1935−1946.
(378) Emerson, J. A.; Toolan, D. T. W.; Howse, J. R.; Furst, E. M.; (395) Kiel, J. W.; Kirby, B. J.; Majkrzak, C. F.; Maranville, B. B.;
Epps, T. H. Determination of Solvent-Polymer and Polymer-Polymer Mackay, M. E. Nanoparticle Concentration Profile in Polymer-Based
Flory-Huggins Interaction Parameters for poly(3-Hexylthiophene) via Solar Cells. Soft Matter 2010, 6, 641.
Solvent Vapor Swelling. Macromolecules 2013, 46, 6533−6540. (396) Parnell, A. J.; Dunbar, A. D. F.; Pearson, A. J.; Staniec, P. a;
(379) Perez, L. A.; Rogers, J. T.; Brady, M. A.; Sun, Y.; Welch, G. C.; Dennison, A. J. C.; Hamamatsu, H.; Skoda, M. W. A.; Lidzey, D. G.;
Schmidt, K.; Toney, M. F.; Jinnai, H.; Heeger, A. J.; Chabinyc, M. L.; Jones, R. A. L. Depletion of PCBM at the Cathode Interface in P3HT/
et al. The Role of Solvent Additive Processing in High Performance PCBM Thin Films as Quantified via Neutron Reflectivity Measure-
Small Molecule Solar Cells. Chem. Mater. 2014, 26, 6531−6541. ments. Adv. Mater. 2010, 22, 2444−2447.
(380) Coates, N. E.; Hwang, I. W.; Peet, J.; Bazan, G. C.; Moses, D.; (397) Kirschner, S. B.; Smith, N. P.; Wepasnick, K. a.; Katz, H. E.;
Heeger, A. J. 1,8-Octanedithiol as a Processing Additive for Bulk Kirby, B. J.; Borchers, J. a.; Reich, D. H. X-Ray and Neutron
Heterojunction Materials: Enhanced Photoconductive Response. Appl. Reflectivity and Electronic Properties of PCBM-Poly(bromo)styrene
Phys. Lett. 2008, 93, 072105. Blends and Bilayers with poly(3-Hexylthiophene). J. Mater. Chem.
(381) Yao, Y.; Hou, J.; Xu, Z.; Li, G.; Yang, Y. Effects of Solvent 2012, 22, 4364.
Mixtures on the Nanoscale Phase Separation in Polymer Solar Cells. (398) Lee, K. H.; Zhang, Y.; Burn, P. L.; Gentle, I. R.; James, M.;
Adv. Funct. Mater. 2008, 18, 1783−1789. Nelson, A.; Meredith, P. Correlation of Diffusion and Performance in
(382) Cheng, P.; Lin, Y.; Zawacka, N. K.; Andersen, T. R.; Liu, W.; Sequentially Processed P3HT/PCBM Heterojunction Films by Time-
Bundgaard, E.; Jørgensen, M.; Chen, H.; Krebs, F. C.; Zhan, X. Resolved Neutron Reflectometry. J. Mater. Chem. C 2013, 1, 2593−
Comparison of Additive Amount Used in Spin-Coated and Roll- 2598.
Coated Organic Solar Cells. J. Mater. Chem. A 2014, 2, 19542−19549. (399) Chen, H.; Hu, S.; Zang, H.; Hu, B.; Dadmun, M. Precise
(383) Lee, J. K.; Ma, W. L.; Brabec, C. J.; Yuen, J.; Moon, J. S.; Kim, Structural Development and Its Correlation to Function in
J. Y.; Lee, K.; Bazan, G. C.; Heeger, A. J. Processing Additives for Conjugated Polymer: Fullerene Thin Films by Controlled Solvent
Improved Efficiency from Bulk Heterojunction Solar Cells. J. Am. Annealing. Adv. Funct. Mater. 2013, 23, 1701−1710.
Chem. Soc. 2008, 130, 3619−3623. (400) Chen, D.; Nakahara, A.; Wei, D.; Nordlund, D.; Russell, T. P.
(384) Bartelt, J. A.; Douglas, J. D.; Mateker, W. R.; Labban, A. E.; P3HT/PCBM Bulk Heterojunction Organic Photovoltaics: Correlat-
Tassone, C. J.; Toney, M. F.; Fréchet, J. M. J.; Beaujuge, P. M.; ing Efficiency and Morphology. Nano Lett. 2011, 11, 561−567.
McGehee, M. D. Controlling Solution-Phase Polymer Aggregation (401) Liu, F.; Zhao, W.; Tumbleston, J. R.; Wang, C.; Gu, Y.; Wang,
with Molecular Weight and Solvent Additives to Optimize Polymer- D.; Briseno, A. L.; Ade, H.; Russell, T. P. Understanding the
Fullerene Bulk Heterojunction Solar Cells. Adv. Energy Mater. 2014, 4, Morphology of PTB7:PCBM Blends in Organic Photovoltaics. Adv.
1301733. Energy Mater. 2014, 4, 1301377.
(385) Moon, J. S.; Takacs, C. J.; Cho, S.; Coffin, R. C.; Kim, H.; (402) Welch, G. C.; Perez, L. A.; Hoven, C. V.; Zhang, Y.; Dang, X.-
Bazan, G. C.; Heeger, A. J. Effect of Processing Additive on the D.; Sharenko, A.; Toney, M. F.; Kramer, E. J.; Nguyen, T.-Q.; Bazan,
Nanomorphology of a Bulk Heterojunction Material. Nano Lett. 2010, G. C. A Modular Molecular Framework for Utility in Small-Molecule
10, 4005−4008. Solution-Processed Organic Photovoltaic Devices. J. Mater. Chem.
(386) Seo, J. H.; Nam, S. Y.; Lee, K. S.; Kim, T. D.; Cho, S. The 2011, 21, 12700.
Effect of Processing Additive on Aggregated Fullerene Derivatives in (403) Campoy-Quiles, M.; Ferenczi, T.; Agostinelli, T.; Etchegoin, P.
Bulk-Heterojunction Polymer Solar Cells. Org. Electron. 2012, 13, G.; Kim, Y.; Anthopoulos, T. D.; Stavrinou, P. N.; Bradley, D. D. C.;
570−578. Nelson, J. Morphology Evolution via Self-Organization and Lateral and
(387) Love, J. A.; Chou, S.-H.; Huang, Y.; Bazan, G. C.; Nguyen, T.- Vertical Diffusion in Polymer:fullerene Solar Cell Blends. Nat. Mater.
Q. Effects of Solvent Additive on “s-Shaped” Curves in Solution- 2008, 7, 158−164.
Processed Small Molecule Solar Cells. Beilstein J. Org. Chem. 2016, 12, (404) Campoy-Quiles, M.; Alonso, M. I.; Bradley, D. D. C.; Richter,
2543−2555. L. J. Advanced Ellipsometric Characterization of Conjugated Polymer
(388) Song, X.; Gasparini, N.; Baran, D. The Influence of Solvent Films. Adv. Funct. Mater. 2014, 24, 2116−2134.
Additive on Polymer Solar Cells Employing Fullerene and Non- (405) Madsen, M. V.; Sylvester-Hvid, K. O.; Dastmalchi, B.; Hingerl,
Fullerene Acceptors. Adv. Electron. Mater. 2017, 1700358. K.; Norrman, K.; Tromholt, T.; Manceau, M.; Angmo, D.; Krebs, F. C.
(389) Ye, L.; Jing, Y.; Guo, X.; Sun, H.; Zhang, S.; Zhang, M.; Huo, Ellipsometry as a Nondestructive Depth Profiling Tool for Roll-to-Roll
L.; Hou, J. Remove the Residual Additives toward Enhanced Efficiency Manufactured Flexible Solar Cells. J. Phys. Chem. C 2011, 115, 10817−
with Higher Reproducibility in Polymer Solar Cells. J. Phys. Chem. C 10822.
2013, 117, 14920−14928. (406) Tremolet de Villers, B.; Tassone, C. J.; Tolbert, S. H.;
(390) Orimo, A.; Masuda, K.; Honda, S.; Benten, H.; Ito, S.; Ohkita, Schwartz, B. J. Improving the Reproducibility of P3HT:PCBM Solar
H.; Tsuji, H. Surface Segregation at the Aluminum Interface of poly(3- Cells by Controlling the PCBM/Cathode Interface. J. Phys. Chem. C
Hexylthiophene)/fullerene Solar Cells. Appl. Phys. Lett. 2010, 96, 2009, 113, 18978−18982.
043305. (407) Li, G.; Chu, C.-W.; Shrotriya, V.; Huang, J.; Yang, Y. Efficient
(391) Lu, H.; Akgun, B.; Russell, T. P. Morphological Character- Inverted Polymer Solar Cells. Appl. Phys. Lett. 2006, 88, 253503.
ization of a Low-Bandgap Crystalline Polymer:PCBM Bulk Hetero- (408) Finck, B. Y.; Schwartz, B. J. Understanding the Origin of the S-
junction Solar Cells. Adv. Energy Mater. 2011, 1, 870−878. Curve in Conjugated Polymer/fullerene Photovoltaics from Drift-
(392) Xu, Z.; Chen, L.-M.; Yang, G.; Huang, C.-H.; Hou, J.; Wu, Y.; Diffusion Simulations. Appl. Phys. Lett. 2013, 103, 053306.
Li, G.; Hsu, C.-S.; Yang, Y. Vertical Phase Separation in Poly(3- (409) Germack, D. S.; Chan, C. K.; Hamadani, B. H.; Richter, L. J.;
Hexylthiophene): Fullerene Derivative Blends and Its Advantage for Fischer, D. A.; Gundlach, D. J.; DeLongchamp, D. M. Substrate-
Inverted Structure Solar Cells. Adv. Funct. Mater. 2009, 19, 1227− Dependent Interface Composition and Charge Transport in Films for
1234. Organic Photovoltaics. Appl. Phys. Lett. 2009, 94, 233303.

3506 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507
Chemical Reviews Review

(410) Hadipour, A.; de Boer, B.; Blom, P. W. M. Organic Tandem Strong Π−Π Interaction for Efficient Fullerene-Free Polymer Solar
and Multi-Junction Solar Cells. Adv. Funct. Mater. 2008, 18, 169−181. Cells. Adv. Energy Mater. 2016, 6, 1600742.
(411) Ameri, T.; Dennler, G.; Lungenschmied, C.; Brabec, C. J. (429) Li, S.; Zhang, H.; Zhao, W.; Ye, L.; Yao, H.; Yang, B.; Zhang,
Organic Tandem Solar Cells: A Review. Energy Environ. Sci. 2009, 2, S.; Hou, J. Green-Solvent-Processed All-Polymer Solar Cells
347−363. Containing a Perylene Diimide-Based Acceptor with an Efficiency
(412) Hadipour, A.; de Boer, B.; Wildeman, J.; Kooistra, F. B.; over 6.5%. Adv. Energy Mater. 2016, 6, 1501991.
Hummelen, J. C.; Turbiez, M. G. R.; Wienk, M. M.; Janssen, R. A. J.; (430) Zheng, Z.; Awartani, O. M.; Gautam, B.; Liu, D.; Qin, Y.; Li,
Blom, P. W. M. Solution-Processed Organic Tandem Solar Cells. Adv. W.; Bataller, A.; Gundogdu, K.; Ade, H.; Hou, J. Efficient Charge
Funct. Mater. 2006, 16, 1897−1903. Transfer and Fine-Tuned Energy Level Alignment in a THF-Processed
(413) Hong, Z.; Dou, L.; Li, G.; Yang, Y. Tandem Solar Cell Fullerene-Free Organic Solar Cell with 11.3% Efficiency. Adv. Mater.
Concept and Practice in Organic Solar Cells. In Topics in Applied 2017, 29, 1604241.
Physics; Yang, Y., Li, G., Eds.; Springer: Berlin, 2015; Vol. 130, pp (431) Andersen, T. R.; Larsen-Olsen, T. T.; Andreasen, B.; Böttiger,
315−346. A. P. L.; Carlé, J. E.; Helgesen, M.; Bundgaard, E.; Norrman, K.;
(414) Ameri, T.; Li, N.; Brabec, C. J. Highly Efficient Organic Andreasen, J. W.; Jørgensen, M.; et al. Aqueous Processing of Low-
Tandem Solar Cells: A Follow up Review. Energy Environ. Sci. 2013, 6, Band-Gap Polymer Solar Cells Using Roll-to-Roll Methods. ACS Nano
2390. 2011, 5, 4188−4196.
(415) Li, W.; Furlan, A.; Hendriks, K. H.; Wienk, M. M.; Janssen, R. (432) Søndergaard, R.; Hösel, M.; Angmo, D.; Larsen-Olsen, T. T.;
A. J. Efficient Tandem and Triple-Junction Polymer Solar Cells. J. Am. Krebs, F. C. Roll-to-Roll Fabrication of Polymer Solar Cells. Mater.
Chem. Soc. 2013, 135, 5529−5532. Today 2012, 15, 36−49.
(416) Yusoff, A. R. b. M.; Kim, D.; Kim, H. P.; Shneider, F. K.; da (433) Lin, Y.; Dam, H. F.; Andersen, T. R.; Bundgaard, E.; Fu, W.;
Chen, H.; Krebs, F. C.; Zhan, X. Ambient Roll-to-Roll Fabrication of
Silva, W. J.; Jang, J. A High Efficiency Solution Processed Polymer
Flexible Solar Cells Based on Small Molecules. J. Mater. Chem. C 2013,
Inverted Triple-Junction Solar Cell Exhibiting a Power Conversion
1, 8007.
Efficiency of 11.83%. Energy Environ. Sci. 2015, 8, 303−316.
(434) Helgesen, M.; Carlé, J. E.; Krebs, F. C. Slot-Die Coating of a
(417) Liu, W.; Li, S.; Huang, J.; Yang, S.; Chen, J.; Zuo, L.; Shi, M.;
High Performance Copolymer in a Readily Scalable Roll Process for
Zhan, X.; Li, C. Z.; Chen, H. Nonfullerene Tandem Organic Solar Polymer Solar Cells. Adv. Energy Mater. 2013, 3, 1664−1669.
Cells with High Open-Circuit Voltage of 1.97 V. Adv. Mater. 2016, 28, (435) Andersen, T. R.; Dam, H. F.; Hösel, M.; Helgesen, M.; Carlé, J.
9729−9734. E.; Larsen-Olsen, T. T.; Gevorgyan, S. A.; Andreasen, J. W.; Adams, J.;
(418) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Li, N.; et al. Scalable, Ambient Atmosphere Roll-to-Roll Manufacture
Mi, Q.; Santori, E. A.; Lewis, N. S. Solar Water Splitting Cells. Chem. of Encapsulated Large Area, Flexible Organic Tandem Solar Cell
Rev. 2010, 110, 6446−6473. Modules. Energy Environ. Sci. 2014, 7, 2925.
(419) Gao, Y.; Le Corre, V. M.; Gaïtis, A.; Neophytou, M.; Hamid, (436) Liu, W.; Liu, S.; Zawacka, N. K.; Andersen, T. R.; Cheng, P.;
M. A.; Takanabe, K.; Beaujuge, P. M. Homo-Tandem Polymer Solar Fu, L.; Chen, M.; Fu, W.; Bundgaard, E.; Jørgensen, M.; et al. Roll-
Cells with VOC > 1.8 V for Efficient PV-Driven Water Splitting. Adv. Coating Fabrication of Flexible Large Area Small Molecule Solar Cells
Mater. 2016, 28, 3366−3373. with Power Conversion Efficiency Exceeding 1%. J. Mater. Chem. A
(420) Chen, S.; Zhang, G.; Liu, J.; Yao, H.; Zhang, J.; Ma, T.; Li, Z.; 2014, 2, 19809−19814.
Yan, H. An All-Solution Processed Recombination Layer with Mild (437) Liu, K.; Larsen-Olsen, T. T.; Lin, Y.; Beliatis, M.; Bundgaard,
Post-Treatment Enabling Efficient Homo-Tandem Non-Fullerene E.; Jørgensen, M.; Krebs, F. C.; Zhan, X. Roll-Coating Fabrication of
Organic Solar Cells. Adv. Mater. 2017, 29, 1604231. Flexible Organic Solar Cells: Comparison of Fullerene and Fullerene-
(421) Zuo, L.; Yu, J.; Shi, X.; Lin, F.; Tang, W.; Jen, A. K. Y. High- Free Systems. J. Mater. Chem. A 2016, 4, 1044−1051.
Efficiency Nonfullerene Organic Solar Cells with a Parallel Tandem (438) Liu, W.; Shi, H.; Andersen, T. R.; Zawacka, N. K.; Cheng, P.;
Configuration. Adv. Mater. 2017, 29, 1702547. Bundgaard, E.; Shi, M.; Zhan, X.; Krebs, F. C.; Chen, H. Roll-Coating
(422) Zhang, G.; Hawks, S. A.; Ngo, C.; Schelhas, L. T.; Scholes, D. Fabrication of ITO-Free Flexible Solar Cells Based on a Non-Fullerene
T.; Kang, H.; Aguirre, J. C.; Tolbert, S. H.; Schwartz, B. J. Extensive Small Molecule Acceptor. RSC Adv. 2015, 5, 36001−36006.
Penetration of Evaporated Electrode Metals into Fullerene Films: (439) Li, S.; Yan, J.; Li, C.-Z.; Liu, F.; Shi, M.; Chen, H.; Russell, T. P.
Intercalated Metal Nanostructures and Influence on Device A Non-Fullerene Electron Acceptor Modified by Thiophene-2-
Architecture. ACS Appl. Mater. Interfaces 2015, 7, 25247−25258. Carbonitrile for Solution-Processed Organic Solar Cells. J. Mater.
(423) Sun, C.; Wu, Z.; Hu, Z.; Xiao, J.; Zhao, W.; Li, H.-W.; Li, Q.-Y.; Chem. A 2016, 4, 3777−3783.
Tsang, S.-W.; Xu, Y.-X.; Zhang, K.; et al. Interface Design for High-
Efficiency Non-Fullerene Polymer Solar Cells. Energy Environ. Sci.
2017, 10, 1784−1791.
(424) Hu, H.; Jiang, K.; Yang, G.; Liu, J.; Li, Z.; Lin, H.; Liu, Y.;
Zhao, J.; Zhang, J.; Huang, F.; et al. Terthiophene-Based D-A Polymer
with an Asymmetric Arrangement of Alkyl Chains That Enables
Efficient Polymer Solar Cells. J. Am. Chem. Soc. 2015, 137, 14149−
14157.
(425) Price, S. C.; Stuart, a. C.; Yang, L. G.; Zhou, H. X.; You, W.
Fluorine Substituted Conjugated Polymer of Medium Bandwidth Yield
7% Efficiency in Polymer-Fullerine Solar Cells. J. Am. Chem. Soc. 2011,
133, 4625−4631.
(426) Guo, B.; Li, W.; Guo, X.; Meng, X.; Ma, W.; Zhang, M.; Li, Y.
High Efficiency Nonfullerene Polymer Solar Cells with Thick Active
Layer and Large Area. Adv. Mater. 2017, 29, 1702291.
(427) Li, S.; Ye, L.; Zhao, W.; Zhang, S.; Ade, H.; Hou, J. Significant
Influence of the Methoxyl Substitution Position on Optoelectronic
Properties and Molecular Packing of Small-Molecule Electron
Acceptors for Photovoltaic Cells. Adv. Energy Mater. 2017, 7, 1700183.
(428) Yao, H.; Yu, R.; Shin, T. J.; Zhang, H.; Zhang, S.; Jang, B.;
Uddin, M. A.; Woo, H. Y.; Hou, J. A Wide Bandgap Polymer with

3507 DOI: 10.1021/acs.chemrev.7b00535


Chem. Rev. 2018, 118, 3447−3507

You might also like