Literature Review With Refrence

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 17

Dye Sensitized Solar Cell

Khatibi et al. (2019) investigated the electric current flow in first, second and third generation of
photovoltaic (PV) solar cells such as mono-crystalline silicon, amorphous silicon, and dye-
synthesized solar cells. A larger amount of current was produced by using PV array and
modules. Concentrated PV (CPV) systems absorbed more radiations through smaller absorbing
surface area. Third generation multi junction solar cell was effective to explain current-voltage
relation having higher efficiency.[1]

Ghannadi et al. (2020) explained the role of Titania (TiO2) nanorod arrays (NRAs) that
increased the efficiency of dye-sensitized solar cells (DSSCs). NRAs could enhance the
efficiency of DSSC more than two times from 1.49% to 3.37%. Titania was manufactured by
sol-electrophoretic deposition (EPD) method by using anodized aluminum oxide (AAO)
nanochannel pattern. Resistance of TiO2 was lowered from 232.2 Ω to 45.13 Ω due to presence
of one dimensional NRAs structures which improved electron transfer.[2]

Ahmad et al. (2020) described the effect of different concentrations of tungsten disulfite (WS 2)
nanosheets on polyaniline (PANi) nano-rods to find out the electronic properties. 5% WS 2 in
PANi exerted PANi 7.32% incident photo conversion efficiency (IPCE) that was 92% equal to
Platinum (Pt) counter electrode based dye synthesized solar cells (DSSC).[3]

Yao et al. (2015) prepared an electron donor N-annulated indenoperylene (NIP) having photo
chemically inactive auxiliary segments and further conjugated carried out via ethynyl with
electron-acceptor benzothiadiazolylbenzoic acid for a metal free donoracceptor dye. Tailored
indenoperylene dye achieves a high power conversion efficiency of 12.5% under an irradiation
of 100 mW cm-2 AM1.5G sunlight. New donoracceptor dye C275 having electron-donor N-
annulated indenoperylene and electron acceptor ethynyl benzothiadiazolylbenzoic acid was
prepared.[4]

Lin et al. (2020) manufactured metal-free organic dyes named as 3, 3’-dithioalkyl-2, 2’-
bithiophene (SBT) organic chromophores for dye-sensitized solar cells (DSSC). Structural
change of SBT π-linkers with different alkyl chains and conjugated thiophene units, significant
chromophore aggregation and interfacial charge recombination can be inhibited. New organic
dye with an impressive η of 9.46% has been found under 1 sun condition. It also exhibited η of
12.34% under a T5 fluorescent illumination of 2400 lux.[5]

Jiao et al. (2018) synthesized two phenothiazine-based organic dyes named as mCPPR and
pCPPR as sensitizers for dye-sensitized solar cell (DSSC) having carbazole as donor,
phenothiazine as π-bridge, and rhodanine acetic acid as acceptor. Benzene acted as the π-linker
between carbazole donor and phenothiazine π-bridge with Meta or para connection mode. Dye
with the para-connected mode, the DSSC devices exhibited higher power conversion efficiency
of 5.57% than those based on dye with meta-connection mode (4.56%).[6]

Hailu et al. (2020) used density functional theory (DFT) to observe the interaction between
complex Iodide redox couple with dyes, denoted as D1Y, D2Y and D3Y having donor-π-acceptor
(D-π-A) configuration were prepared by replacing functional groups on donor moiety. Better
adsorption and electronic properties were analyzed when addition of strong donor unit in D 1Y
and D2Ydyes having interaction with iodide electrolyte on TiO 2 surface, was carried out, rather
than the dye alone on oxide surface.[7]

Diksha et al. (2020) manufactured polymer electrolyte because of doping polyvinylpyrrolidone


(PVP) and ammonium iodide ion with an ionic liquid (IL) 1-hexyl-3-methylimidazolium iodide
using solution cast technique. Use of IL and electrical conductivity film increased the ionic
conductivity of dye-sensitized solar cells (DSSCs). Preparation of IL-based solid polymer
electrolyte and optimized PVP: NH4I + 4 wt % IL shows the highest value of conductivity, that is
1.30 ×10-3 S cm1 at room temperature.[8]

Rahman et al. (2019) utilized a hydrothermal method to describe the efficiency of dye-sensitized
mesoporous photo anode of Ti doped zinc oxide (Ti-ZnO). X-ray photoelectron spectroscopy,
transmission electron microscopy and X-ray diffraction used to sort out crystallinity,
morphology, surface area and optical and electrochemical properties of the Ti-ZnO. Light
harvesting efficiency rose up by maximizing surface area of 131.85 m 2 g-1 and controlled band
gap of Ti-ZnO.[9]

Selvanathan et al. (2020) described quasi-solid dye-sensitized solar cell (QSDDSC) using
organosoluble starch-cellulose binary polymer. Phthaloyl starch - hydroxyethyl cellulose blend
based quasi-solid electrolyte was fabricated by utilizing LiI and TPAI respectively, representing
a small and bulky cation of iodide salt. Gel composition with 12.5 wt % TPAI generated the
highest photo conversion efficiency of 3.94%.[10]

Silicon Based Solar Cells

Verner et al. (2016) introduced first a low-temperature process for semitransparent perovskite
solar cells, yielding efficiencies of up to 14.5%. Then, implemented this process to fabricate
monolithic perovskite/silicon heterojunction tandem solar cells yielding efficiencies of up to 21.2
and 19.2% for cell areas of 0.17 and 1.22 cm2.[11]

Nagamatsu et al. (2015) prevented minority carrier holes in silicon from recombining at the
cathode contact of a silicon-based photovoltaic device and employed an electron-selective
titanium dioxide (TiO2) heterojunction contact to silicon. The anode is a poly(3,4
ethylenedioxythiophene):poly(styrenesulfonate) heterojunction to silicon that produces a hole
selective contact, allowing the complete device to be manufactured at a maximum efficiency.
Silicon-based double-sided heterojunction solar cell utilizing non-silicon barrier materials was
used.[12]

Chen et al. (2015) reported high-throughput blade coating of a wide-bandgap perovskite on sub-
micrometer-scale, double-side-textured silicon to achieve efficient perovskite/ silicon monolithic
tandem solar cells. Deposition of 0.5–1 mm thick perovskite layers from solution onto textured
silicon with typical pyramid heights of 3–10 mm remains a challenge. It showed current density
of this new tandem device architecture, extending the path length of long-wavelength light
within the silicon and reducing the reflection loss to 2.3 mA/cm2 to reach above 19 mA/cm 2 and
an efficiency of 26%.[13]

Heidarzadeh et al. (2019) showed that the efficiency of silicon based ultra-thin film solar cell
depends on angle of incident light. Optical path length of incoming light inside the active layer is
increased using a distributed Bragg reflector (DBR) and a rectangular, triangular or cylindrical
shaped grating horizontally arranged and placed on top of DBR. The efficiency of the cell with
DBR and cylindrical backside reflector was better than the normal incident cases for incidence
angles between of 0° and 30°.[14]

Dreon et al. (2020) replaced the traditional (p) and (n) doped amorphous silicon selective layers
by molybdenum oxide (MoOx) as a front-side hole-selective layer. The thicknesses of the (i)a-
Si:H and MoOx layers in silicon heterojunction (SHJ) devices and their influence on the optical
and electrical performance of the cell was discussed. Based on short current density (J sc) trends, a
1 nm-thick MoOx device increased its Jsc by 0.7 mA/cm2 compared to cells featuring a 4 nm-thick
MoOx layer.[15]

Chen et al. (2020) demonstrated an industrial tunnel oxide passivated contacts (i-TOPCon)
silicon solar cell on large area n-type silicon wafers (156.75 ×156.75 cm 2). The cell contained a
boron diffused front emitter, a tunnel-SiOx/n+-poly-Si/ SiNx:H structure at the rear side, and
screen-printed electrodes on both sides. Development of i-TOPCon cells confirmed champion
efficiency of 24.58%.[16]

Liu et al. (2018) introduced energy harvesting structure that integrated a solar cell and a
triboelectric nanogenerator (TENG) device to build power generation from both sunlight and
raindrop. A mutual electrode of a poly (3, 4-ethylenedioxythiophene): poly (styrenesulfonate)
(PEDOT: PSS) film was used to reduce sun light reflection, leading to enhanced short current
density (Jsc). Imprinted-PEDOT: PSS deposited on Si achieved the higher PCE of 13.6% due to
rain drop collection. [17]

Yoshikawa et al. (2017) fabricated large-area silicon solar cells combining interdigitated back
contacts (IBC) and an amorphous silicon/crystalline silicon heterojunction. The photo conversion
efficiency is over 26% with a 180.4 cm2 designated area, which is an improvement of 2.7%
relative to the previous record efficiency of 25.6% using high-quality thin film HJ passivation
technology and low-resistance electrodes.[18]

Zhu et al. (2017) fabricated a silicon-based nanoheterostructure (p+ -Si/p-Si/ n+ -Si (and n-Si)/n-
ZnO nanowire (NW) array) photovoltaic device and demonstrate the improved device
performance through significantly enhanced light absorption by ZnO NW array and effective
charge carrier separation by the piezophototronic effect. The efficiency of the photovoltaic cell
was improved from 8.97% to 9.51% by simply applying a static compress strain.[19]

Duan et al. (2016) investigated that nanowire (NW) antireflection coating could enhance light
trapping capability, which is generally used in crystal silicon (CS) based solar cells. The convex
conical NW has the best light trapping, but the concave conical NW has the best effective light
absorption. Furthermore, if the cross section of silicon NW is changed into a square, both light
trapping and effective light absorption are enhanced, and the Eiffel Tower shaped NW arrays
have optimal effective light absorption. ffective light absorption and its enhancement factor for
silicon nanowire-based solar cell.[20]

Organic and Inorganic Solar Cells

Gao et al (2020), scientists have investigated that porphyrin is used in making low-bandgap
donor material due to its conjugated plan and strong absorption .This was also investigated that
porphyrin based material have greater efficiencies in small molecules .Scientist’s discussion
about the design and synthetic routes leads to make porphyrin multifunctional for use in OSCS
and PVSCs.[21]
Haung et al (2020), Scientists demonstrated that 3-µm-thick ultra-flexible OSCS by use of a
mixed fullerene /non-fullerene acceptors that can obtain an efficiency of 13% with 97%
efficiency in the PEC after 1000 bending cycle. Scientists further concluded that ultra-flexible
OSCS shows best mechanical behaviour under producing 89% retention in the PEC after 1000
cycles. Scientists finally concluded that high efficiency can be obtained by adding small
fullerene-based acceptor into non-fullerene system.[22]
Zhao et al (2021), scientists completely studied Ternary OSCs based on polymer/small
molecule/small and all small molecules classification .They concluded that incorporating third
component into a binary blend can highlight absorption spectra ,facilitate excitation dissociation
and charge transport .Finally scientists concluded that morphology control has great role in V oc
,Vsc and IF which tells the photoelectric performance of OSCs.[23]
Dai et al (2016) Researchers concluded that Poly (3,4ethylenedioxide thiophene): poly
(styrenesulphonate) are used in organic –silicon heterojunction solar cells with transitional
pyramida texturing surface. The structure of bottom of pyramid can be maximized by Acid
Isotropic Etching (AIE). Efficiency of 13.78% is attained by depositing on the front of device.
[24]
Duan et al (2019), Non- fullerene acceptors with optoelectronic properties have demonstrated
higher PEC over 14% in single junction and 17% in tandon OSCs by incorporating NF OSCs.
Scientists explained that fluorination could be an effective method to tune the HMO/LUMO
energy level and increase electron mobility .It was concluded that an effective intermolecular π-π
stacking is needed to obtain high carrier transporting ability in active layer.[25]
Liu et al (2017), scientists concluded that Ternary solar cells work in parallel-like device model.
High Performance .Ternary solar cells have bandgap polymer donor and two acceptor materials,
Phenyl –C 70 butyric acid methyl ester (PC BM) and non-Fullerene Acceptors. Solar cells
performance is dependent on weight ratio of ITIC-Th and Pc 70 BM form their own transport
network.Finally scientist concluded that as ITIC-Th content reaches to 50% the whole and
electron mobilities both increased.[26]
Cardinalette et al (2018), Scientists Concluded that class of organic based solar cells can become
a disruptive technology in space applications and processing possibilities. Scientists further
observed that organic –base photoelectric opens a new paradigm for solar electricity in space,
from satellites to orbitals and any space station.[27]
Singh et al. (2019) Scientists have investigated that various HTMs are high stable devices
maintain power conversion efficiency over 20%. Most efficient perovskite solar cells follow a p-
i-n type of architecture with a lot of p and n type materials for efficient contraction .Carbon
based HTM free PSC shows highest PCE above 16% with stability of 600h using CH 3NH3PbI3.
[28]
Liu et al. (2020) Scientists have investigated that a perovskite solar cells using csPbxMg-XI2Br
with a full printable framework of FTO/c -Tio 2/ Al2o3Nio/ carbon achieve a power conversion
efficiency of 108%. The phenomenon of energy loss in mesoporous perovskite solar cells can be
decreased in the case of CSPbo98 Mgo.O2I2Br.A record efficiency of 16.29% have been attained
for a device with planner structure of glass/FTo/Tio 2 by adding CaCl2 as passivator and n- type
dopant. Chemical and Electronic states in CSPb (1-X) Mg XI2Br film can be traced by
Spectroscopic characterization.[29]
Wan et al. (2020) Scientists have concluded that high quality CsPbBr 3 perovskite film can be
manufactured by two steps spin coating method in which in which CsBr methanol/H 20 mixed
solvent solution is spin coated into the lead bromide film. CsPbBr 3 exhibits CsPbBr3 a bandgap of
17.3ev which is suitable for photovoltaic application.α- CsPbBr 3 can be degraded to a non-
photoactive ϭ-phase at room temperature.The existence of H20 increase the concentration of
CsBr solution.[30]
Ho-Baillie et al (2019), Scientists have investigated that due to higher band gap and thermal
stability CsPbIXBr3 –X are suitable for tandem solar cells applications and other optoelectronic
devices .These have 90% theoretical current output limit. Device FF and out voltage can be
improved by producing positive valence and conduction BOs. 60% efficiency limit imposed on
current devices can be decreased by improving carrier life-time and improving surface
impassivation.[31]
Renaud et al. (2016) Scientists have found that integration of Molybdenum clusters as new
absorbers in all inorganic cells have been investigated via the fabrication of first cluster
sensitized solar cells. The associated emission for the molybdenum octahedral clusters inside
units correspond to broad peak form 600 to 900 nm with maximum around 730nm.Tio 2 film for
n-DSSC were obtained from a commercial Tio2 monoparticle paste deposited by screen printing
on FTO-Coated glass substrate sintered and treated with Ticl4 aqueous solution.[32]
Xiang et al. (2019) Scientists have found that dopping on the lead site has become a useful
strategy for obtaining desired film quality and material phase for stable inorganic perovskite
solar cells. When Barium is added to all inorganic CsPbI2Br perovskite and its use as an absorber
layer in Inorganic PSCs that gets a PCE of 14% under one sun illumination. The Perovskite films
were deposited from precursor solution on Mesoporous Tio 2 substrate at a spinning speed of
3000 r.p.m. The Ba-Pb mixed perovskite used to fabricate all inorganic PSCs.[33]
Wang et al. (2018) scientists have concluded that high quality and stable α-phase CsPbl 3is
obtained by solvent control growth of the precursor film in a dry environment. Cesium Lead
Iodide is good for making efficient solar cells. A 15.7% PCE of CsPbl 3 solar cells can be
obtained by SCG of absorb layer and devices can tolerate above 500 hours of continuous line
soaking .The partial change from delta phase to the Beta phase CsPbl 3 precursor materials were
diffused and precursor films were reconsidered during solvent evaporation.[34]

Fullerene Based Solar Cells

Albes et al. (2018) said that fullerene-based organic solar cells having a minute quantity of donor
gave a good amount of photocurrent supporting a large open-circuit voltage. At low
concentrations the donor was fully separated within the fullerene and no passages of holes exist
to the anode; whereas this morphology was opposite to bulk-heterojunction donor: acceptor
blends where percolation passages for both electrons and holes were present within their
respective transport phases. Two important factors controlling the process of current transfer
values of highest occupied molecular orbital (HOMO) level difference between donor and the
accepter and of recombination strength, both were found to agree the experimental and kinetic
Monte Carlo simulations. Evidence provided that the HOMO level difference between the donor
and the accepter was smaller in dilute donor systems. The reason for introducing bulk-
heterojunction was not an absolute requirement to obtain large amount of photocurrents in
organic solar cells.[35]

Cho (2017) gave an experimental data that an efficient fullerene based electron transporting
material C60,N,N,N-trimethyl-1-(2,3,4-tris(2-(2-methoxyethoxy)ethoxy)phenyl)dimetha-
naminiummonoadduct iodide salt (LT-S9079), was introduced to large area flexible polymer
solar cells (PSCs). The excellent conductivity and wettability of the LT-S9079 on the active layer
enabled the fabrication of large-area slot-die coated PSCs of conventional photovoltaic device
structure. The PSCs incorporating LT-S9079 demonstrate high current densities of 8.14mA
cm−2 and power conversion efficiencies of 1.46%. Finally he determined that slot-die coatedLT-
S9079films require relatively high thick-ness, approximately 350 nm, and near-flat surface
uniformity. In addition, the excellent wet-ting properties of LT-S9079 on P3HT:PCBM allowed
for the slot-die coating of PSCs with conventional photovoltaic device structure. He
demonstrated the viability of LT-S9079 as an ETL material in slot-die processing. His
conclusions, to the best of his knowledge, noted the highest PCE of any ordinary structured slot-
die processed PSC with fullerene based ETLs.[36]

Collado-Fregoso et al. (2019) stated that the participation of charge transfer describes the
photogeneration and rejoining of charge carriers had been a vital focus of study within the
organic photovoltaic community. In that work, they investigated the molecular factors
determining the mechanism of photocurrent generation in low-donor-content organic

Solar cells, where the activated layer was made of vacuum-deposited C60 and small amounts of
organic donor molecules. They found a pronounced decline of all photovoltaic parameters with
decreasing CT state energy. Using a combination of steady-state photocurrent measurements and
time-delayed collection field experiments, they demonstrated that the power conversion
efficiency, and specifically, the fill factor of these devices, was basically determined by the bias
dependence of photocurrent production. By joining these findings with the results from ultrafast
transient absorption spectroscopy, they demonstrated that blends with small CT energies perform
poorly because of an increased nonradiative CT state decay rate and that this decay obey an
energy-gap law. Their work challenged the common view that a large energy offset at the
heterojunction and/or the presence of fullerene clusters guarantee efficient CT dissociation and
rather indicated that charge generation benefits from high CT state energies by a slower decay to
the ground state.[37]

Elnaggar et al .(2019) A family of four different fullerene derivatives found suitable as electron
transport layer (ETL) materials for p-i-n PSCs. It was shown that all designed compounds form
high quality coatings above the perovskite films with undetectable voids or pinholes. Perovskite
solar cells (PSCs) consisting fullerene-based ETLs methano fullerenes (F1–F4) and the reference
phenyl-C61-butyric acid methyl ester (PCBM) demonstrated comparable power conversion
efficiencies, whereas their ambient stability was unique. The devices consisting reference PCBM
degraded almost completely within less than 100 h of air exposure. On the contrary, the solar
cells made with the best-performing ETL material F1 maintained more than 90% of the initial
efficiency after they spent 800 h under ambient atmosphere conditions (RH 30%–40%). The
shown impressive ambient solidness of PSCs caused by the fullerene derivative F1 applied as
ETL was attributed to its optimal molecular architecture with the side chains filling the gaps
between the fullerene spheres, thus preventing the diffusion of oxygen and moisture inside the
device. The demonstrated effect of the molecular engineering paved a way toward a substantial
improvement in the isolation features of the fullerene-based ETLs, thus paving a way to
increased operation lifetimes of perovskite photovoltaic.[38]

Li et al. (2017) stated that fullerene based ternary solar cells have gained great attention and
progress in recent years. Additives can effectively overcome disadvantages of polymer based
solar cells as providing cascade energy levels, increasing light harvesting ability and turning
morphology. Thus polymer solar cells had potential to work on and get advancement.[39]

Yang et al. (2020) described that morphology and crystallinity of a fullerene based organic solar
cells have a great impact on its efficiency of these cells. The poor compatibility of donor
accepter molecules in the bulk heterojunction (BHJs) hinders the further improvement of the
device performance and stability in organic solar cells. In this work, the conjugated polymer
PBDTTPD-COOH was introduced as a third component into BHJ films of PTB7-Th:PC71BM
and PffBT4T-2OD:PC71BM to improve the crystallinity and morphology. The crystallinity of
both donor polymers was enhanced and more face-on orientated crystals were observed in the
corresponding films, which were correlated with the enhancement of the current density of the
related solar cells. The improved BHJ morphology leads to an increased fill factor. Also, the
device stability significantly increased by the addition of the third component PBDTTPD-
COOH. The T80 lifetime value is enhanced 10 times in the doped devices as compared with the
binary solar cells in the case of the PTB7-Th: PC71 BM series.[40]

Ganesamoorthy et al. (2017) stated that the bulk hetro-junction organic solar cells (BHJ-OSCs)
blended with the coating of highly electron rich, poly-3-hexylthiophene (P3HT), donor and
extremely electron deficient, soluble C60derivative, [6,6]-phenyl-C61-butyric acid methyl ester
(PC61BM) as an acceptor between the high work function positive electrode, ITO and low work
function metal negative electrode Al, showed exciting results than the bilayer counterpart in the
last two decades, thus the scientific circle and solar cell manufacturing groups are focusing on
this treasured field. Fullerene derivative based BHJ-OSCs showed amazing efficiencies due to
the high donor-acceptor (D-A) interface, high open-circuit voltage (V oc), easy processing, and
thermal stability. So the design of these cells is very critical for its performance. Numerous
papers were published on the PC61/71BM and various functionally modified fullerene acceptors
in the past several years with higher efficiency, higher LUMO level, and better solubility. The
symmetrical structure of PC61BM showed high aggregation properties, insufficient absorption in
the visible region, but unsymmetrical PC71BM had a good optical absorption in the visible
region, hence the best solar cells. From the comparison, it was clear that the modification in the
PC61/71BM core with aryl group, alkyl chain length, and end group modification didn’t show
prominent difference in the power conversion efficiency PCE with the P3HT polymer donor in
BHJ-OSCs device structure. Fullerenes multi-adducts performed better than the PC61/71BM
because of development in the V oc. In the multi-functionalization, mainly bis-functionalization
showed better efficient than mono-adduct. C60 and C70 based hetro bis-adducts namely, C60
(CH2)(Ind), and OQMF70 showed the best PCE of 5.90%, and 6.61% respectively, which was
comparable with the indene based bis-adducts.[41]

Non Fullerene Based Solar Cells

Yu et al. (2018) reported a method of applying volatilizable SAs to improve photovoltaic


efficiency of NF OSCs based on A–D–A-type acceptors. By taking the molecular structure of the
NF acceptor IT-4F, we designed and synthesized a volatilizable solid additive SA-1, which can
be well mixed with IT-4F and volatilized by TA. The intermolecular π–π interaction and charge
transport properties of IT-4F can be greatly enhanced since the more condensed and organized
molecular arrangement after the SA-1 volatilized from the active layer, contributing to the
enhancement of Jsc and FF of the optimal OSCs. Moreover, the device designed by using SA-1
have better device stability and reproducibility than the most widely used solvent additive, 1,8-
diiodooctane (DIO). They also showed that some of SA –x, which have similar volatility to that
of SA-1, can also be effective for achieving high performance devices whereas SA-1 can be a
general additive in varied active layers for enhancing photovoltaic performance of NF OSCs.
Their results suggested that using volatilizable SAs was an alternative and potential method of
designing OSCs more efficient in future industrialization and that designing specific additives for
different active materials will be a good direction in the photovoltaicfield.[42]

Chao et al. (2018) stated that large photon energy loss (Eloss) in the conversion from photons to
electrons is still an important factor in limiting the performance of polymer solar cells (PSCs).To
obtain the decreased Eloss and expanded spectral response simultaneously the reasonable design
of wide band gap conjugated polymer and low-lying HOMO level is still a great challenge in
nonfullerene PSCs. In that work, four D–A typewide band-gap conjugated polymers by
introducing chlorine atoms on the conjugated thienyl-side chains of the BDT units, namely,
PBT1Cl-Bz, PBT2Cl as-Bz, PBT2Cls-Bz and PBT4Cl-Bz, were formed and characterized.
Because of C-Cl large dipole and the strong non covalent interaction of Cl⋯S and Cl⋯π, the
tetrachloro-substituted PBT4Cl-Bz shows a much lower HOMO level (-5.64 eV) than other three
polymers. So, a high open-circuit voltage (Voc) of 0.96 V was formed while keeping a decent
Jsc of 16.04 mA cm−2 when matched with nonfullerene acceptor IT-4F, and a quite small Eloss
of 0.54 eV was achieved which was close to the values of some inorganic and hybrid solar cells.
[43]

Li et al. (2018) reported that they found a PBDB-T-rich top layer formed when the PBDB-T:IT-
M film casted on PEDOT:PSS, indicating a vertical component distribution that would hinder
electron transport to the cathode in a conventional device. This PBDB-T-rich top layer remained
upon low-temperature annealing at 80 C, but disappeared when the annealing temperature is
increased, resulting in an optimum annealing temperature of 160C for conventional devices as
the removal of this polymer-rich layer facilitates electron transport toward top cathode layer.[44]

Lin et al. (2018) stated that a series of new electron acceptors containing a truxene core with
intense optical absorption were made and used for non-fullerene organic solar cells. Due to the
weak electron-donating characteristic of truxene core and thereby weak intramolecular charge
transfer interaction from electron-donating core to electron-with-drawing end groups, the
resulting new acceptors showed relatively wide optical band-gap and high-lying lowest
unoccupied molecular orbitals (LUMOs), which finally lead to complementary light spectra with
narrow band-gap donor polymers and high open circuit voltage (Voc) in solar cells. Particularly,
Tr(Hex)6-3BR, a star-shaped planar acceptor, produced the highest power conversion efficiency
of 2.1% with a high Voc of 1.02 V when blended with PTB7-Th.[45]

Awartani et al. (2018) investigated two similar materials differentiated by minor side-chain
differences, PBDB-O and PBDB-T, which when blended with ITIC yield efficiencies of 4.35%
and 11.21%, respectively. To elucidate this large difference in performance between these
chemically alike material systems, and what factors played a role in dictating these varying
efficiencies, we sketched out the various photophysical characteristics between the two materials
and their blends. They observed that PBDB-T: ITIC has longer polaron lifetimes (>5000 ps) and
reduced geminate recombination. RSoXS results showed that a larger mean-square composition
variation at small length scales on the order of typical diffusion lengths in the PBDB-T:ITIC is
likely to improve exciton dissociation into free charges by lowering geminate recombination.
Lastly and most importantly, the significantly improved as well as more balanced electron and
hole mobility in PBDB-T: ITIC.[46]

(Jung et al., 2018) stated that in contrast to widely investigated fullerene-based organic ETMs for
PSCs, only few types of non-fullerene-based electron transporting materials (ETMs) have been
developed. In this concept article, we describe a representative design concept for non-fullerene
ETMs for use in normal (n-i-p) and inverted (p-i-n) type PSCs. Based on a discussion on
ETM requirements for PSCs, we highlight recently developed non-fullerene polymeric and small
molecular-based ETMs for PSCs. In addition, they presented various other approaches for
enhancing PSC device performance such as using additional dopants and introducing amine
terminal groups in ETMs, and ETM-based interlayers.[47]

Doumon et al. (2019) in more than hundreds of NFAs with applications in organic photovoltaic
(OPVs) have been synthesized, enabling a notable current record efficiency of above 15%.
Hence, there was a shift in interest toward the use of NFAs in OPVs. However, there had been
negligible work on the stability of these new materials in devices. More importantly, there was
very little comparative work on the photostability of FA versus NFA solar cells to ascertain the
pros and cons of these two systems. They found that in spite of the polymer, the cell structure, or
the initial efficiency, the [70]PCBM devices were more photostable than the ITIC ones. This
observation, however, opposed the assumption that NFA solar cells were more photochemically
stable. These results suggest that complementary absorption should not take precedence in the
design rules for the synthesis of new molecules and there is still work left to be done to achieve
stable and efficient OSCs.[48]

Karuthedath et al. (2021) in bulk heterojunction (BHJ) organic solar cells (OSCs) both the
electron affinity (EA) and ionization energy (IE) offsets at the donor–acceptor interface should
equally control exciton dissociation. Here, we demonstrate that in low-bandgap non-fullerene
acceptor (NFA) BHJs ultrafast donor-to-acceptor energy transfer precedes hole transfer from the
acceptor to the donor and thus renders the EA offset virtually unimportant. Moreover, sizeable
bulk IE offsets of about 0.5 eV are needed for efficient charge transfer and high internal quantum
efficiencies, since energy level bending at the donor–NFA interface caused by the acceptors’
quadrupole moments prevents efficient exciton-to-charge-transfer state conversion at low IE
offsets. The same bending, however, is the origin of the barrier-less charge transfer state to free
charge conversion. Our results provide a compre-hensive picture of the photophysics of NFA-
based blends, and show that sizeable bulk IE offsets are essential to design efficient BHJ OSCs
based on low-bandgap NFAs.[49]

Chang et al. (2019) stated that two types of all-small-molecule ternary solar cells comprising of
two small-molecule donors and one acceptor (fullerene/non-fullerene) were developed.
Interestingly, both these devices had a common component: a carefully designed medium band
gap small molecule, which had appropriate energy levels and displays shows compatibility with
the host donor. In the fullerene system, the charge-relaying role of the additive donor is
confirmed by the improved charge transportation and suppressed charge recombination. While in
the non-fullerene system, the mixed face-on and edge-on orientation of the ternary film induced
by the additive donor dominates the promotion of charge transportation. Accordingly, both
ternary devices gave higher short-circuit current density, fill factor, and power conversion
efficiencies of over 10% compared to binary ones. That work offered a promising guideline on
the development of high-performance all-small-molecule ternary solar cells by incorporating a
miscible small-molecule donor.[50]

.
Muhammad Rizwan Dye Sensitized and Silicon based Solar Cells

Muhammad Aqib Ayaz Organic and Inorganic Solar Cells

Mehboob Ahmed Fullerene and Non Fullerene Based Solar Cells

1. Khatibi, A., F. Razi Astaraei, and M.H. Ahmadi, Generation and combination of the solar cells: A
current model review. Energy Science & Engineering, 2019. 7(2): p. 305-322.
2. Ghannadi, S., et al., Sol-electrophoretic deposition of TiO2 nanoparticle/nanorod array for
photoanode of dye-sensitized solar cell. Materials Chemistry and Physics, 2021. 258: p. 123893.
3. Ahmad, M.S., et al., Effect of WS2 nano-sheets on the catalytic activity of polyaniline nano-rods
based counter electrode for dye sensitized solar cell. Physica E: Low-dimensional Systems and
Nanostructures, 2021. 126: p. 114466.
4. Yao, Z., et al., Donor/acceptor indenoperylene dye for highly efficient organic dye-sensitized solar
cells. Journal of the American Chemical Society, 2015. 137(11): p. 3799-3802.
5. Lin, F.-S., et al., Metal-free efficient dye-sensitized solar cells based on thioalkylated bithiophenyl
organic dyes. Journal of Materials Chemistry C, 2020. 8(43): p. 15322-15330.
6. Jiao, Y., et al., Effects of meta or para connected organic dyes for dye-sensitized solar cell. Dyes
and Pigments, 2018. 158: p. 165-174.
7. Hailu, Y.M., M.T. Nguyen, and J.-C. Jiang, Theoretical study on the interaction of iodide
electrolyte/organic dye with the TiO 2 surface in dye-sensitized solar cells. Physical Chemistry
Chemical Physics, 2020. 22(45): p. 26410-26418.
8. Singh, D., et al., Polyvinylpyrrolidone with ammonium iodide and 1-hexyl-3-methylimidazolium
iodide ionic liquid-doped solid polymer electrolyte for efficient dye sensitized solar cell. High
Performance Polymers, 2020. 32(2): p. 130-134.
9. Rahman, M.U., et al., Efficient dye-sensitized solar cells composed of nanostructural ZnO doped
with Ti. Catalysts, 2019. 9(3): p. 273.
10. Selvanathan, V., et al., Organosoluble Starch-Cellulose Binary Polymer Blend as a Quasi-Solid
Electrolyte in a Dye-Sensitized Solar Cell. Polymers, 2020. 12(3): p. 516.
11. Werner, J., et al., Efficient monolithic perovskite/silicon tandem solar cell with cell area> 1 cm2.
The journal of physical chemistry letters, 2016. 7(1): p. 161-166.
12. Nagamatsu, K.A., et al., Titanium dioxide/silicon hole-blocking selective contact to enable
double-heterojunction crystalline silicon-based solar cell. Applied Physics Letters, 2015. 106(12):
p. 123906.
13. Chen, B., et al., Blade-coated perovskites on textured silicon for 26%-efficient monolithic
perovskite/silicon tandem solar cells. Joule, 2020. 4(4): p. 850-864.
14. Heidarzadeh, H. and A. Tavousi, Performance enhancement methods of an ultra-thin silicon solar
cell using different shapes of back grating and angle of incidence light. Materials Science and
Engineering: B, 2019. 240: p. 1-6.
15. Dréon, J., et al., 23.5%-efficient silicon heterojunction silicon solar cell using molybdenum oxide
as hole-selective contact. Nano Energy, 2020. 70: p. 104495.
16. Chen, D., et al., 24.58% total area efficiency of screen-printed, large area industrial silicon solar
cells with the tunnel oxide passivated contacts (i-TOPCon) design. Solar Energy Materials and
Solar Cells, 2020. 206: p. 110258.
17. Liu, Y., et al., Integrating a silicon solar cell with a triboelectric nanogenerator via a mutual
electrode for harvesting energy from sunlight and raindrops. ACS nano, 2018. 12(3): p. 2893-
2899.
18. Yoshikawa, K., et al., Silicon heterojunction solar cell with interdigitated back contacts for a
photoconversion efficiency over 26%. Nature energy, 2017. 2(5): p. 1-8.
19. Zhu, L., et al., Enhancing the efficiency of silicon-based solar cells by the piezo-phototronic effect.
ACS nano, 2017. 11(2): p. 1894-1900.
20. Duan, Z., et al., Effective light absorption and its enhancement factor for silicon nanowire-based
solar cell. Applied optics, 2016. 55(1): p. 117-121.
21. Gao, K., et al., Low‐bandgap porphyrins for highly efficient organic solar cells: materials,
morphology, and applications. Advanced Materials, 2020. 32(32): p. 1906129.
22. Huang, W., et al., Efficient and mechanically robust ultraflexible organic solar cells based on
mixed acceptors. Joule, 2020. 4(1): p. 128-141.
23. Zhao, C., et al., Recent advances, challenges and prospects in ternary organic solar cells.
Nanoscale, 2021. 13(4): p. 2181-2208.
24. Dai, X., et al., Improving performance of organic-silicon heterojunction solar cells based on
textured surface via acid processing. ACS applied materials & interfaces, 2016. 8(23): p. 14572-
14577.
25. Duan, L., et al., Progress in non-fullerene acceptor based organic solar cells. Solar Energy
Materials and Solar Cells, 2019. 193: p. 22-65.
26. Liu, T., et al., Highly efficient parallel-like ternary organic solar cells. Chemistry of Materials,
2017. 29(7): p. 2914-2920.
27. Cardinaletti, I., et al., Organic and perovskite solar cells for space applications. Solar Energy
Materials and Solar Cells, 2018. 182: p. 121-127.
28. Singh, R., et al., Review of current progress in inorganic hole-transport materials for perovskite
solar cells. Applied Materials Today, 2019. 14: p. 175-200.
29. Liu, S., et al., Stable and efficient full-printable solar cells using inorganic metal oxide framework
and inorganic perovskites. Applied Materials Today, 2020. 20: p. 100644.
30. Wan, X., et al., Efficient and stable planar all-inorganic perovskite solar cells based on high-
quality CsPbBr3 films with controllable morphology. Journal of Energy Chemistry, 2020. 46: p. 8-
15.
31. Ho-Baillie, A., et al., Untapped potentials of inorganic metal halide perovskite solar cells. Joule,
2019. 3(4): p. 938-955.
32. Renaud, A., et al., Inorganic molybdenum clusters as light‐harvester in all inorganic solar cells: a
proof of concept. ChemistrySelect, 2016. 1(10): p. 2284-2289.
33. Xiang, W., et al., Ba-induced phase segregation and band gap reduction in mixed-halide
inorganic perovskite solar cells. Nature communications, 2019. 10(1): p. 1-8.
34. Wang, P., et al., Solvent-controlled growth of inorganic perovskite films in dry environment for
efficient and stable solar cells. Nature communications, 2018. 9(1): p. 1-7.
35. Albes, T., et al., Origin of photocurrent in fullerene-based solar cells. The Journal of Physical
Chemistry C, 2018. 122(27): p. 15140-15148.
36. Cho, N., Fullerene based electron transporting layers for large area polymer solar cells.
Molecular Crystals and Liquid Crystals, 2017. 655(1): p. 159-165.
37. Collado-Fregoso, E., et al., Energy-gap law for photocurrent generation in fullerene-based
organic solar cells: the case of low-donor-content blends. Journal of the American Chemical
Society, 2019. 141(6): p. 2329-2341.
38. Elnaggar, M., et al., Molecular Engineering of the Fullerene‐Based Electron Transport Layer
Materials for Improving Ambient Stability of Perovskite Solar Cells. Solar RRL, 2019. 3(9): p.
1900223.
39. Li, H., K. Lu, and Z. Wei, Polymer/small molecule/fullerene based ternary solar cells. Advanced
Energy Materials, 2017. 7(17): p. 1602540.
40. Yang, D., et al., Tailoring Morphology Compatibility and Device Stability by Adding PBDTTPD-
COOH as Third Component to Fullerene-Based Polymer Solar Cells. ACS Applied Energy Materials,
2020. 3(3): p. 2604-2613.
41. Ganesamoorthy, R., G. Sathiyan, and P. Sakthivel, Fullerene based acceptors for efficient bulk
heterojunction organic solar cell applications. Solar Energy Materials and Solar Cells, 2017. 161:
p. 102-148.
42. Yu, R., et al., Design and application of volatilizable solid additives in non-fullerene organic solar
cells. Nature communications, 2018. 9(1): p. 1-9.
43. Chao, P., et al., Multichloro-substitution strategy: facing low photon energy loss in nonfullerene
solar cells. ACS Applied Energy Materials, 2018. 1(11): p. 6549-6559.
44. Li, W., et al., Correlating Three‐dimensional Morphology With Function in PBDB‐T: IT‐M Non‐
Fullerene Organic Solar Cells. Solar RRL, 2018. 2(9): p. 1800114.
45. Lin, K., et al., Star-shaped electron acceptors containing a truxene core for non-fullerene solar
cells. Organic Electronics, 2018. 52: p. 42-50.
46. Awartani, O.M., et al., Polymer non-fullerene solar cells of vastly different efficiencies for minor
side-chain modification: impact of charge transfer, carrier lifetime, morphology and mobility.
Journal of Materials Chemistry A, 2018. 6(26): p. 12484-12492.
47. Jung, S.K., et al., Non‐Fullerene Organic Electron‐Transporting Materials for Perovskite Solar
Cells. ChemSusChem, 2018. 11(22): p. 3882-3892.
48. Doumon, N.Y., et al., Photostability of fullerene and non-fullerene polymer solar cells: the role of
the acceptor. ACS applied materials & interfaces, 2019. 11(8): p. 8310-8318.
49. Karuthedath, S., et al., Intrinsic efficiency limits in low-bandgap non-fullerene acceptor organic
solar cells. Nature Materials, 2021. 20(3): p. 378-384.
50. Chang, Y., et al., Constructing High‐Performance All‐Small‐Molecule Ternary Solar Cells with the
Same Third Component but Different Mechanisms for Fullerene and Non‐fullerene Systems.
Advanced Energy Materials, 2019. 9(16): p. 1900190.

You might also like