Next Article in Journal
Tribo-Mechanical Characterization of Carbon Fiber-Reinforced Cyanate Ester Resins Modified With Fillers
Previous Article in Journal
Bending Behavior of Lightweight Wood-Based Sandwich Beams with Auxetic Cellular Core
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase Behavior of Amorphous/Semicrystalline Conjugated Polymer Blends

1
Institute of Engineering Materials and Biomaterials, Faculty of Mechanical Engineering, Silesian University of Technology, 44-100 Gliwice, Poland
2
School of Materials Science and Engineering, Jimma Institute of Technology, Jimma University, Post Office Box 378 Jimma, Ethiopia
3
Centre of Polymer and Carbon Materials, Polish Academy of Sciences, M. Curie-Skłodowska 34 Street, 41-819 Zabrze, Poland
4
School of Chemical Engineering, Jimma Institute of Technology, Jimma University, Post Office Box 378 Jimma, Ethiopia
*
Authors to whom correspondence should be addressed.
Polymers 2020, 12(8), 1726; https://doi.org/10.3390/polym12081726
Submission received: 22 June 2020 / Revised: 24 July 2020 / Accepted: 29 July 2020 / Published: 31 July 2020
(This article belongs to the Section Polymer Physics and Theory)

Abstract

:
We report the phase behavior of amorphous/semicrystalline conjugated polymer blends composed of low bandgap poly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta [2,1-b;3,4-b′]dithiophene) -alt-4,7(2,1,3-benzothiadiazole)] (PCPDTBT) and poly{(N,N′-bis(2-octyldodecyl)naphthalene -1,4,5,8-bis(dicarboximide)-2,6-diyl)-alt-5,5′-(2,2′-bithiophene)} (P(NDI2OD-T2)). As usual in polymer blends, these two polymers are immiscible because ΔSm ≈ 0 and ΔHm > 0, leading to ΔGm > 0, in which ΔSm, ΔHm, and ΔGm are the entropy, enthalpy, and Gibbs free energy of mixing, respectively. Specifically, the Flory–Huggins interaction parameter (χ) for the PCPDTBT /P(NDI2OD-T2) blend was estimated to be 1.26 at 298.15 K, indicating that the blend was immiscible. When thermally analyzed, the melting and crystallization point depression was observed with increasing PCPDTBT amounts in the blends. In the same vein, the X-ray diffraction (XRD) patterns showed that the π-π interactions in P(NDI2OD-T2) lamellae were diminished if PCPDTBT was incorporated into the blends. Finally, the correlation of the solid-liquid phase transition and structural information for the blend system may provide insight for understanding other amorphous/semicrystalline conjugated polymers used as active layers in all-polymer solar cells, although the specific morphology of a film is largely affected by nonequilibrium kinetics.

Graphical Abstract

1. Introduction

Conjugated polymer blends composed of polymer donor (PD) and polymer acceptor (PA) have been used as an active material for all-polymer solar cells (all-PSCs), leading to a high power conversion efficiency (PCE) of ~10–14.4% [1,2,3,4,5,6,7,8]. Here, when PD and PA are mixed together, the goal of this mixing is not to make a miscible blend unlike the versatile commercial blends, e.g., poly(vinyl chloride) (PVC)-butadiene/acrylonitrile copolymer (NBR) and polystyrene-poly(2,6-dimehtyl -1,4-phenylene oxide) (PPO) blends [9], but an immiscible (or partially miscible) blend for the separation of Frenkel excitons at the PD/PA heterojunction [10,11]. In general, the ideal size of phase-separated domains is ~10–20 nm because the exciton has a limited diffusion length (LD = ~10 nm) due to low dielectric constants of organic macromolecules [12,13,14]. However, if two homopolymers are mixed, the polymer blends may undergo macrophase separation because of the increased Gibbs free energy of mixing [15]. On the other hand, when PD and PA are linearly linked by covalent bonds, these block copolymers may undergo microphase separation into ordered structures, such as cubic sphere, hexagonal cylinder, bicontinuous gyroid, and lamellae [16,17]. Hence, considering the necessity of nanoscale phase separation (much smaller than micro-/macro-scale domains) for organic photovoltaics (OPV), we need to understand the polymer–polymer thermodynamic behavior in the PD/PA blend films.
If PD and PA are amorphous, there are two types of phase separation in liquid–liquid (L–L) phase transition: nucleation and growth (NG), and spinodal decomposition (SD) [18,19]. Here, NG proceeds in the metastable region, whereas SD takes place in the unstable region without any energy barrier. However, if PD and/or PA are semicrystalline, the liquid–solid (L–S) phase transition (crystallization) also takes place with L–L phase transition [20,21,22]. Indeed, in many PA/PD blend systems, they contain a stereoregular (or regioregular) polymer, leading to both L–L and L–S phase transition, directly affecting the morphologies of PD/PA blend films. Importantly, the pre-formed aggregation through L–S phase transition in OPV blend solutions may play a significant role in generating nanoscale phase domains instead of macrophase separation [20].
In previous studies, we reported the order-disorder phase equilibria of regioregular poly(3-hexylthiophene-2,5-dyil) (r-reg P3HT) solution [20]. Then, we also studied the phase diagrams of low bandgap poly[2,6-(4,4-bis(2-ethylhexyl)-4H–cyclopenta[2,1-b;3,4-b′]dithiophene) -alt-4,7(2,1,3-benzothiadiazole)] (PCPDTBT) solution as a function of solvent, polymer, and chain length [21]. Recently, we investigated the phase behavior of crystalline/crystalline conjugated polymer blends, composed of r-reg P3HT and poly{(N,N′-bis(2-octyldodecyl)naphthalene-1,4,5,8 -bis(dicarboximide)-2,6-diyl)-alt-5,5′-(2,2′-bithiophene)} (P(NDI2OD-T2)). Then, we also explained the upper critical solution temperature (UCST) behavior of each polymer solution as a function of solvent species, for which the Flory–Huggins lattice theory was employed [22].
In this study, we further extended our works to an amorphous/crystalline conjugated polymer blend, for which PCPDTBT and P(NDI2OD-T2) were chosen as a model system. Here, PCPDTBT is a p-type polymer acting as PD [23,24,25,26]. Interestingly, if PCPDTBT with bulky branched alkyl side chains (ethyl hexyl groups) has a relatively low number average molecular weight ( M n ≈ 3.2 kg/mol), it shows amorphous or marginally crystallizable behavior, indicating that the chain ends (just like impurities) may disturb polymer crystallization process [21]. However, PCPDTBT with high M n ≈ 35 kg/mol could be highly crystallized when this polymer was processed with exposure to chlorobenzene (CB) vapors and/or with some solvent additives such as 1,8-diiodooctane (DIO), 1,8-octanedithiol (ODT), and 1,8-dichlorooctane (DCO) [23,24,25,26]. Note that in such a case, we may not rule out the effect of substrate on polymer’s crystallization. On the other hand, P(NDI2OD-T2) is a highly crystallizable n-type polymer, acting as PA in all-PSCs or n-channels, exhibiting very high electron mobility ( μ n ≈ 0.85 cm2/Vs) in the top-gate bottom-contact (TGBC) organic thin-film transistor (OTFT) [27,28,29,30,31,32,33]. Here, it is notable that P(NDI2OD-T2)’s glass transition temperature (Tg) was reported to be −70 °C and −40 °C, based on dynamic mechanical analysis (DMA) and differential scanning calorimetry (DSC), respectively [34,35]. Hence, inspired by the important properties of these two polymers [21,23,24,25,26,27,28,29,30,31,32,33,34,35], we further studied the thermodynamic behavior of PCPDTBT/P(NDI2OD-T2) blends based on the thermal and structural analyses. Specifically, the melting phenomena of crystalline P(NDI2OD-T2) with added amorphous PCPDTBT material was investigated using DSC and confirmed by X-ray diffraction (XRD), in which π-stacking was observed to be dependent on the amount of amorphous PCPDTBT in the blend film. Finally, the morphologies of the pure polymer and 1:1 blend films were investigated using the tapping-mode atomic force microscopy (AFM).

2. Materials and Methods

2.1. Materials

P(NDI2OD-T2) [Mn = 32.1 kg/mol, Mw = 90.0 kg/mol, polydispersity index (PDI) = 2.8, and molecular formula = (C62H88N2O4S2)n] was purchased from 1-Material, Inc. (Quebec, Canada). PCPDTBT [Mn = 4.50 kg/mol, Mw = 13.5 kg/mol, PDI = 3.0, and molecular formula = (C31H38N2S3)n], chlorobenzene (CB), and chloroform (CF) were provided from Sigma-Aldrich, Inc. (Taufkirchen, Germany). All these materials were used as received without further purification.

2.2. Film Processing

Two conjugated polymers, PCPDTBT and P(NDI2OD-T2), were dissolved in the mixed solvents, CB:CF = 1:1 (wt. ratio) according to literature reports [36,37,38]. The total concentration of the blend system was 10 mg/mL. After the solvent dissolved completely, the samples were spin-coated on a glass substrate with the dimensions of 1.5 × 1.5 cm2 for thin film deposition. Prior to spin-coating, the glass substrate was sequentially cleaned in deionized water, chloroform, and isopropanol for 5 min, respectively, and then dried under a nitrogen atmosphere. The condition of spin-coating was 2000 rpm for 15 s in the air, leading to the film thickness of about 140 nm. The annealing condition for a film was 180 °C for 15 min.

2.3. Thermal Property Characterization

Thermal analysis was performed using DSC (2920-DSC, TA Instruments, Champaign, IL, USA) to characterize the transition temperature of polymers at the scan rate of 10 °C/min under N2. Thermal decomposition of polymers was monitored using thermogravimetric analysis (TGA; a METTLER TOLEDO STARe System, Warsaw, Poland), in which samples were heated from 50 to 600 °C using a conventional heating ramp with a scan rate of 10 °C/min under N2.

2.4. Film Characterization

X-ray diffraction (XRD) measurements (PANalytical Inc., Malvern, United Kingdom; PX 3040 PR; Cu Kα radiation; λ = 1.5418 Å) were performed to examine the structure and ordering of a film on glass substrate at room temperature. According to the XRD’s instrumental set-up condition, the diffraction angle (2θ) ranged from 10° to 100°. The morphologies of the polymer films were characterized by the tapping-mode AFM (XE-100 Park Systems, Mannheim, Germany). Here, the data were analyzed using the software Park Systems XEI.

3. Results

Figure 1 shows the chemical structures of (a) P(NDI2OD-T2) and (b) PCPDTBT. In the previous works, we reported that the solubility parameter (δ) of P(NDI2OD-T2) is 7.99 [22], whereas that of PCPDTBT (Mn = 3.2 kg/mol and PDI = 2.2) is 10.70 [21]. Note that in this work, PCPDTBT has Mn = 4.5 kg/mol and PDI = 3.0, indicating a very minor difference between the two PCPDTBT samples. Hence, based on the former δ data, the Flory–Huggins interaction parameter ( χ i j ) [22,39,40] for the PCPDTBT/P(NDI2OD-T2) polymer blends could be estimated using Equation (1),
χ i j = V ^ s R T ( δ i δ j ) 2
where V ^ s (= lattice site volume) and δ i   o r   j were the molar volume of a solvent [ V ^ s = (112.56 g/mol)/(1.11 g/cm3) = 101.41 cm3/mol for CB] and the solubility parameter [subscript i or j = 1, 2; PCPDTBT = 1 and P(NDI2OD-T2) = 2]. R and T were the gas constant and temperature (K). Then χ 12 was estimated to be 374.82   K / T , indicating that χ 12 = 1.26 at T = 298.15 K and χ 12 = 0.5 at T = 749.64 K. Hence, considering the critical value of interaction parameter [41], ( χ 12 ) c r i t = 1 2 ( 1 r 1 + 1 r 2 ) 2 ≈ 0.022 and χ 12 > ( χ 12 ) c r i t , we conclude that PCPDTBT and P(NDI2OD-T2) were immiscible as usual for many polymer–polymer mixtures. Here, we used r 1 = ( M n , 1 ρ 1 ) ( M W s ρ s ) = ( 4500 1 ) ( 112.56 1.11 ) 44 and r 2 = ( M n , 2 ρ 2 ) ( M W s ρ s ) = ( 32100 1.1 ) ( 112.56 1.11 ) 288 , in which r 1   and   r 2 were the relative molar volumes of component 1 and 2, M n , 1   and   M n , 2 were the number average molecular weights of component 1 and 2, M W s was the molecular weight of a solvent (CB), and ρ 1 ,   ρ 2 ,   and   ρ s were the densities of component 1, 2, and solvent, respectively. At this moment, keep in mind that the Flory–Huggins theory should be understood qualitatively, not quantitatively.
As demonstrated in the aforementioned example, the PCPDTBT/P(NDI2OD-T2) mixture was an immiscible blend, which could be understood based on Δ G m = Δ H m T Δ S m where Δ G m was the Gibbs free energy of mixing, Δ H m was the enthalpy of mixing, and Δ S m was the entropy of mixing. When mixing long chain macromolecules, Δ S m 0 and Δ H m > 0 leading to Δ G m > 0 . Note that the stability condition for a single phase is ( 2 Δ G m ϕ i 2 ) T , P > 0 where ϕ i is volume fraction of component i and P is pressure. Hence, most polymer–polymer blends including PCPDTBT and P(NDI2OD-T2) are immiscible but phase-separated. Usually, the same structural units are more attractive to each other than different ones, leading to aggregation and clusters, i.e., the phase separation of blends.
Figure 2 exhibits TGA data, indicating that PCPDTBT and P(NDI2OD-T2) generally decompose at around ~400–500 °C in organic molecules. Figure 3 shows the DSC thermograms for the PCPDTBT/P(NDI2OD-T2) blends. As shown in Figure 3a, P(NDI2OD-T2) was a semicrystalline polymer with Tm = 324 °C (1st heating) or 318 °C (2nd heating) and Tc = 294 °C (1st cooling). On the other hand, PCPDTBT did not show any melting point, indicating it was a typical amorphous polymer [42]; see Figure 3b–d. About the glass transition temperature (Tg), P(NDI2OD-T2) showed it at −44 °C in the DSC thermogram, which agreed with Gu et al.’s report (Tg = −40 °C) within experimental uncertainties [35]. Although in the previous work [21], PCPDTBT showed Tg ≈ 112 °C, in this work we could not observe it clearly. That is because, as shown in Figure 3d, some artifacts were observed for all the samples around 100 °C from the DSC instrumental conditions. Note that Tg is a second order transition, but in many conjugated polymer systems, it is hard to determine Tg due to a weak signal in DSC thermal curves. It may be from the rigid or semirigid backbone structure of conjugated polymers, leading to relatively small free volume (and heat capacity) changes around the second-order transition when compared to flexible coil polymers.
Note that, in this work, our research interest lies primarily in the S-L phase equilibria (the 1st order transition) of conjugated polymer blends composed of the semicrystalline P(NDI2OD-T2) and the amorphous PCPDTBT polymers because it is relatively more clear than other 2nd order transition. As shown in Figure 3a, the 1st heating curve, the melting point is depressed from 323.35 °C at 100% P(NDI2OD-T2) to 317.83 °C at 80% to 312.79 °C at 50%, which is a trend observed in the 1st cooling and 2nd heating curves, as well. Here, note that in Figure 3b, at PCPDTBT:P(NDI2OD-T2) = 50:50, the melting peak looks very broad, indicating that, in a blend system, a semicrystalline component should be required at least at 50% for observing a clear and sharp Tm peak. Then, the results were summarized in Figure 4, displaying both the Tm and Tc depression, in which the average difference (ΔT = Tm − Tc) between Tm and Tc is 30.0 ± 1.6 °C in the first-cycling thermal curve. The origin of this depression comes from the new equilibrium between crystalline lamellar and amorphous (liquid) chains of P(NDI2OD-T2) when a diluent PCPDTBT was introduced into the semicrystalline P(NDI2OD-T2) system for forming polymer blends. Specifically, PCPDTBT has one order lower molecule weight, Mn = 4.5 kg/mol compared to P(NDI2OD-T2) with Mn = 32.1 kg/mol, indicating that many available end groups in PCPDTBT may serve as impurities in blend systems. However, a liquid state PCPDTBT chain molecule itself may act as impurities in semicrystalline P(NDI2OD-T2) blends, depressing the melting point of the polymer blend systems.
Figure 5 shows the XRD patterns of pure P(NDI2OD-T2), mixed P(NDI2OD-T2)/PCPDTBT blends, and pure PCPDTBT, respectively, in the range of 2θ = 10–100°. At 2θ = ~26°, some peaks related with π-π stacking are observed. Here, the crystallite size (t) could be estimated based on Scherrer’s equation, t = 0.9 λ / ( B   cos   θ ) [43,44], where λ is the X-ray wavelength (= 0.154 nm) and B is the full width at half maximum (FWHM) at the diffraction angle θ. The results are summarized in Table 1, indicating that the crystallite size decreases from 2.5 nm at 100% P(NDI2OD-T2) to 2.0 nm at 50%. In the case of PCPDTBT, its nominal crystallite size is very small ~ 1.4 nm, indicating that it is an amorphous material because the length scale is around the unit cell. Importantly, if we recall the DSC results in Figure 3, when the crystallite size (related with π-stacking) is larger than 2 nm, we may observe the melting transition in DSC. It is notable also that, in polymer science, the typical unit cell, chain-folded lamella and spherulite have dimensions of ~0.2–2 nm, ~10–50 nm thick and several microns wide, and ~100–1000 μm, respectively [42]. In the case of conjugated polymer, thin-film samples for OPV, we usually observe up to a lamella scale because a film thickness is ~100–200 nm. However, there are some exceptions reporting a three-dimensional spherulite structure [26,45,46,47]. At this moment, for interested readers, the detailed XRD data about P(NDI2OD-T2) films are available in the literatures [28,29,30,31,32,33], in which edge-on and face-on morphologies were also reported.
Based on the crystallite size analysis in Table 1, as well as the melting point depression in Figure 4, we suggest the scheme displayed in Figure 6. When we add the amorphous chain molecule PCPDTBT into the semicrystalline P(NDI2OD-T2) sample, the crystalline region of P(NDI2OD-T2) may be destroyed, which could be confirmed through the decrease in XRD peaks’ intensity at 2θ = ~26°, i.e., π-π stacking, as well as the diminished melting peak in DSC thermograms in Figure 3. However, it is notable that the aforementioned phenomena do not necessarily mean the miscibility of the two polymers, i.e., they are still in phase-separate state. They are immiscible because of the thermodynamic reason, Δ G m > 0 . However, the above observation may affect the nano- and micro-structure of blend films, which should be important in all-PSCs because of the limited diffusion length of Frenkel excitons in the organic/polymer semiconductors with low dielectric constants.
Figure 7 shows the AFM images of pure P(NDI2OD-T2), P(NDI2OD-T2)/PCPDTBT blend, and pure PCPDTBT. As shown in Figure 7a, P(NDI2OD-T2) film displays rod-shaped crystalline lamellae with a few μm. However, when P(NDI2OD-T2) was mixed with PCPDTBT with a 1:1 wt. ratio, the morphology changed dramatically. Here, elongated crystal domains were destroyed and diminished into a granular shape. Finally, PCPDTBT exhibits somewhat homogenous images when compared with P(NDI2OD-T2). This smoothness is expected because PCPDTBT is amorphous (disordered liquid or glass) based on both the DSC and XRD results.

4. Conclusions and Future Works

Two low bandgap conjugated polymers, semicrystalline P(NDI2OD-T2) and amorphous PCPDTBT, were mixed together for a study on phase behavior. Here, P(NDI2OD-T2) is a high-performance n-type semiconductor whereas PCPDTBT is a p-type one when applied to optoelectronic devices. In this work, although two polymers were immiscible with χ 12 = 1.26 at T = 298.15 K, the blend exhibited the melting and crystallization temperature depression phenomena because PCPDTBT acts as a diluent or impurity for the P(NDI2OD-T2) lamellae. In this process, PCPDTBT molecules make it difficult for the P(NDI2OD-T2) chains to form lamellae, leading them to destroy the regular π-π stacking in the polymer lamellae, which was confirmed through XRD and AFM data. Specifically, when the size of crystallite-related π-stacking was larger than 2 nm, we observed the melting points in DSC. Finally, when we apply conjugated polymer blends to electronic and optoelectronic devices, it is very important to understand a phase-separation mechanism leading to a specific morphology. Further study will be designed for understanding other photovoltaic polymer blends with amorphous/amorphous, semicrystalline/amorphous, and semicrystalline/semicrystalline structures.

Author Contributions

Conceptualization, T.T. and J.Y.K.; Investigation and data curation, G.M.F., P.J., and U.S.; Writing—original draft, G.M.F.; Writing—review and editing, T.T. and J.Y.K.; Supervision, T.T. and J.Y.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by JIT Mega Project, the Ministry of Science and Higher Education in Ethiopia, and the Erasmus Plus and IEMB SUT Poland. The APC was funded by SUT.

Acknowledgments

We would like to thank Marta Musiol and Tymoteusz Jung for measuring the TGA and XRD data, respectively.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

B full width at half maximum
d distance between crystallographic planes (d-spacing)
L D diffusion length
M n number average molecular weight
M n , 1   or   2 number average molecular weight of component 1 or 2
M w weight average molecular weight
M W s molecular weight of a solvent
P pressure
P A polymer acceptor
P D polymer donor
R the gas constant
r 1   or   2 relative molar volume of component 1 or 2
T temperature
T c crystallization temperature
T g glass transition temperature
T m melting temperature
t crystallite size
V ^ s molar volume of a solvent
Δ G m Gibbs free energy of mixing
Δ H m enthalpy of mixing
Δ S m entropy of mixing
δ solubility parameter
δ i   or   j solubility parameter of component i or j
θ diffraction angle
λ X-ray wavelength (= 0.154 nm)
μ n electron mobility
ϕ i volume fraction of component i
ρ 1   or   2   or   s density of component 1 or 2 or solvent
χ Flory–Huggins interaction parameter
χ i j Flory–Huggins interaction parameter between component i and j
( χ 12 ) c r i t Flory–Huggins interaction parameter between component 1 and 2 at critical point
AFMatomic force microscopy
all-PSCsall-polymer solar cells
CBchlorobenzene
CFchloroform
DCO1,8-dichlorooctane
DIO1,8-diiodooctane
DMAdynamic mechanical analysis
DSCdifferential scanning calorimetery
FWHMfull width at half maximum
L–Lliquid–liquid
L–Sliquid–solid
NBRbutadiene/acrylonitrile copolymer
NGnucleation and growth
ODT1,8-octanedithiol
OPVorganic photovoltaics
OTFTorganic thin-film transistor
P(NDI2OD-T2)poly{(N,N′-bis(2-octyldodecyl)naphthalene -1,4,5,8-bis(dicarboximide)-2,6-diyl)-alt-5,5′-(2,2′-bithiophene)}
PCEpower conversion efficiency
PCPDTBTpoly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta [2,1-b;3,4-b′]dithiophene) -alt-4,7(2,1,3-benzothiadiazole)]
PDIpolydispersity index
PPOpolystyrene-poly(2,6-dimehtyl -1,4-phenylene oxide)
PVCpoly(vinyl chloride)
r-reg P3HTregioregular poly(3-hexylthiophene-2,5-dyil)
SDspinodal decomposition
TGAthermogravimetric analysis
TGBCtop-gate bottom-contact
UCSTupper critical solution temperature
XRDX-ray diffraction

References

  1. Jia, T.; Zhang, J.; Zhong, W.; Liang, Y.; Zhng, K.; Dong, S.; Ying, L.; Liu, F.; Wang, X.; Huang, F.; et al. 14.4% efficiency all-polymer solar cell with broad absorption and low energy loss enabled by a novel polymer acceptor. Nano Energy 2020, 72, 104718. [Google Scholar] [CrossRef]
  2. Zhao, R.; Wang, N.; Yu, Y.; Liu, J. Organoboron Polymer for 10% Efficiency All-Polymer Solar Cells. Chem. Mater. 2020, 32, 1308–1314. [Google Scholar] [CrossRef]
  3. Zhu, L.; Zhong, W.; Qiu, C.; Lyu, B.; Zhou, Z.; Zhang, M.; Song, J.; Xu, J.; Wang, J.; Ali, J.; et al. Aggregation-Induced Multilength Scaled Morphology Enabling 11.76% Efficiency in All-Polymer Solar Cells Using Printing Fabrication. Adv. Mater. 2019, 31, 1902899. [Google Scholar] [CrossRef]
  4. Yao, H.; Bai, F.; Hu, H.; Arunagiri, L.; Zhang, J.; Chen, Y.; Yu, H.; Chen, S.; Liu, T.; Lai, J.Y.L.; et al. Efficient All-Polymer Solar Cells based on a New Polymer Acceptor Achieving 10.3% Power Conversion Efficiency. ACS Energy Lett. 2019, 4, 417–422. [Google Scholar] [CrossRef]
  5. Wu, J.; Meng, Y.; Guo, X.; Zhu, L.; Liu, F.; Zhang, M. All-polymer solar cells based on a novel narrow-bandgap polymer acceptor with power conversion efficiency over 10%. J. Mater. Chem. A 2019, 7, 16190–16196. [Google Scholar] [CrossRef]
  6. Liu, X.; Li, X.; Zheng, N.; Gu, C.; Wang, L.; Fang, J.; Yang, C. Insight into the Efficiency and Stability of All-Polymer Solar Cells Based on Two 2D-Conjugated Polymer Donors: Achieving High Fill Factor of 78%. ACS Appl. Mater. Interfaces 2019, 11, 43433–43440. [Google Scholar] [CrossRef]
  7. Genene, Z.; Mammo, W.; Wang, E.; Andersson, M.R. Recent Advances in N-Type Polymers for All-Polymer Solar Cells. Adv. Mater. 2019, 31, 1807275. [Google Scholar] [CrossRef]
  8. Li, Z.; Ying, L.; Zhu, P.; Zhong, W.; Li, N.; Liu, F.; Huang, F.; Cao, Y. A generic green solvent concept boosting the power conversion efficiency of all-polymer solar cells to 11%. Energy Environ. Sci. 2019, 12, 157–163. [Google Scholar] [CrossRef]
  9. Olabisi, O.; Robeson, L.M.; Shaw, M.T. Polymer-Polymer Miscibility; Academic Press: New York, NY, USA, 1979; pp. 13–15. [Google Scholar]
  10. Pope, M.; Swenberg, C. Electronic Processes in Organic Crystals and Polymers, 2nd ed.; Oxford University Press: Oxford, UK, 1999; pp. 1–191. [Google Scholar]
  11. Gregg, B.A. Excitonic Solar Cells. J. Phys. Chem. B 2003, 107, 4688–4698. [Google Scholar] [CrossRef]
  12. Gregg, B.A.; Hanna, M.C. Comparing organic to inorganic photovoltaic cells: Theory, experiment, and simulation. J. Appl. Phys. 2003, 93, 3605–3614. [Google Scholar] [CrossRef]
  13. Mikhnenko, O.V.; Blom, P.W.M.; Nguyen, T.-Q. Exciton diffusion in organic semiconductors. Energy Environ. Sci. 2015, 8, 1867–1888. [Google Scholar] [CrossRef]
  14. Tamai, Y.; Ohkita, H.; Benten, H.; Ito, S. Exciton Diffusion in Conjugated Polymers: From Fundamental Understanding to Improvement in Photovoltaic Conversion Efficiency. J. Phys. Chem. Lett. 2015, 6, 3417–3428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Bates, F.S. Polymer-Polymer Phase Behavior. Science 1991, 251, 898–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Leibler, L. Theory of Microphase Separation in Block Copolymers. Macromolecules 1980, 13, 1602–1617. [Google Scholar] [CrossRef]
  17. Bates, F.S. Block Copolymer Thermodynamics: Theory and Experiment. Annu. Rev. Phys. Chem. 1990, 41, 525–557. [Google Scholar] [CrossRef]
  18. Cahn, J.W. Phase Separation by Spinodal Decomposition in Isotropic Systems. J. Chem. Phys. 1965, 42, 93–99. [Google Scholar] [CrossRef]
  19. McNeil, C.R.; Greenham, N.C. Conjugated-Polymer Blends for Optoelectronics. Adv. Mater. 2009, 21, 3840–3850. [Google Scholar] [CrossRef]
  20. Kim, J.Y. Order−Disorder Phase Equilibria of Regioregular Poly(3-hexylthiophene-2,5-diyl) Solution. Macromolecules 2018, 51, 9026–9034. [Google Scholar] [CrossRef]
  21. Kim, J.Y. Phase Diagrams of Binary Low Bandgap Conjugated Polymer Solutions and Blends. Macromolecules 2019, 52, 4317–4328. [Google Scholar] [CrossRef]
  22. Fanta, G.M.; Jarka, P.; Szeluga, U.; Tański, T.; Kim, J.Y. Phase Diagrams of n-Type Low Bandgap Naphthalenediimide-Bithiophene Copolymer Solutions and Blends. Polymers 2019, 11, 1474. [Google Scholar] [CrossRef] [Green Version]
  23. Mühlbacher, M.; Scharber, M.; Morana, M.; Zhu, Z.; Waller, D.; Gaudiana, R.; Brabec, C. High Photovoltaic Performance of a Low-Bandgap Polymer. Adv. Mater. 2006, 18, 2884–2889. [Google Scholar] [CrossRef]
  24. Morana, M.; Wegscheider, M.; Bonanni, A.; Kopidakis, N.; Shaheen, S.; Scharber, M.; Zhu, Z.; Waller, D.; Gaudiana, R.; Brabec, C. Bipolar Charge Transport in PCPDTBT-PCBM Bulk-Heterojunctions for Photovoltaic Applications. Adv. Funct. Mater. 2008, 18, 1757–1766. [Google Scholar] [CrossRef]
  25. Gu, Y.; Wang, C.; Russell, T.P. Multi-Length-Scale Morphologies in PCPDTBT/PCBM Bulk-Heterojunction Solar Cells. Adv. Energy Mater. 2012, 2, 683–690. [Google Scholar] [CrossRef]
  26. Fischer, F.S.U.; Trefz, D.; Back, J.; Kayunkid, N.; Tornow, B.; Albrecht, S.; Yager, K.G.; Singh, G.; Karim, A.; Neher, D.; et al. Highly Crystalline Films of PCPDTBT with Branched Side Chains by Solvent Vapor Crystallization: Influence on Opto-Electronic Properties. Adv. Mater. 2015, 27, 1223–1228. [Google Scholar] [CrossRef] [PubMed]
  27. Yan, H.; Chen, Z.; Zheng, Y.; Newman, C.; Quinn, J.R.; Dötz, F.; Kastler, M.; Facchetti, A. A high-mobility electron-transporting polymer for printed transistors. Nature 2009, 457, 679–686. [Google Scholar] [CrossRef] [PubMed]
  28. Rivnay, J.; Toney, M.F.; Zheng, Y.; Kauvar, I.V.; Chen, Z.; Wagner, V.; Facchetti, A.; Salleo, A. Unconventional Face-On Texture and Exceptional In-Plane Order of a High Mobility n-Type Polymer. Adv. Mater. 2010, 22, 4359–4363. [Google Scholar] [CrossRef] [Green Version]
  29. Brinkmann, M.; Gonthier, E.; Bogen, S.; Tremel, K.; Ludwigs, S.; Hufnagel, M.; Sommer, M. Segregated versus Mixed Interchain Stacking in Highly Oriented Films of Naphthalene Diimide Bithiophene Copolymers. ACS Nano 2012, 6, 10319–10326. [Google Scholar] [CrossRef]
  30. Takacs, C.J.; Treat, N.D.; Krämer, S.; Chen, Z.; Facchetti, A.; Chabinyc, M.L.; Heeger, A.J. Remarkable Order of a High-Performance Polymer. Nano Lett. 2013, 13, 2522–2527. [Google Scholar] [CrossRef]
  31. Zhou, N.; Lin, H.; Lou, S.J.; Yu, X.; Guo, P.; Manley, E.F.; Loser, S.; Hartnett, P.; Huang, H.; Wasielewski, M.R.; et al. Morphology-Performance Relationships in High-Efficiency All-Polymer Solar Cells. Adv. Energy Mater. 2014, 4, 1300785. [Google Scholar] [CrossRef]
  32. Zhou, K.; Zhang, R.; Liu, J.; Li, M.; Yu, X.; Xing, R.; Han, Y. Donor/Acceptor Molecular Orientation-Dependent Photovoltaic Performance in All-Polymer Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 25352–25361. [Google Scholar] [CrossRef]
  33. Zhang, R.; Yang, H.; Zhou, K.; Zhang, J.; Yu, X.; Liu, J.; Han, Y. Molecular Orientation and Phase Separation by Controlling Chain Segment and Molecule Movement in P3HT/N2200 Blends. Macromolecules 2016, 49, 6987–6996. [Google Scholar] [CrossRef]
  34. Schuettfort, T.; Huettner, S.; Lilliu, S.; Macdonald, J.E.; Thomsen, L.; McNeill, C.R. Surface and Bulk Structural Characterization of a High-Mobility Electron-Transporting Polymer. Macromolecules 2011, 44, 1530–1539. [Google Scholar] [CrossRef]
  35. Gu, X.; Yan, H.; Kurosawa, T.; Schroeder, B.C.; Gu, K.L.; Zhou, Y.; To, J.W.F.; Oosterhout, S.D.; Savikhin, V.; Molina-Lopez, F.; et al. Comparison of the Morphology Development of Polymer-Fullerene and Polymer-Polymer Solar Cells during Solution-Shearing Blade Coating. Adv. Energy Mater. 2016, 6, 1601225. [Google Scholar] [CrossRef]
  36. Yang, Q.; Wang, J.; Zhang, X.; Zhang, J.; Fu, Y.; Xie, Z. Constructing vertical phase separation of polymer blends via mixed solvents to enhance their photovoltaic performance. Sci. China Chem. 2015, 58, 309–316. [Google Scholar] [CrossRef]
  37. Van Franeker, J.J.; Turbiez, M.; Li, W.; Wienk, M.M.; Janssen, R.A.J. A real-time study of the benefits of co-solvents in polymer solar cell processing. Nat. Commun. 2015, 6, 6229. [Google Scholar] [CrossRef] [PubMed]
  38. Kadem, B.Y.; Al-hashimi, M.K.; Hassan, A.K. The Effect of Solution Processing on the Power Conversion Efficiency of P3HT-based Organic Solar Cells. Energy Procedia 2014, 50, 237–245. [Google Scholar] [CrossRef] [Green Version]
  39. Flory, P.J. Principles of Polymer Chemistry; Cornell University Press: New York, NY, USA, 1953. [Google Scholar]
  40. Hiemenz, P.C.; Lodge, T.P. Polymer Chemistry, 2nd ed.; CRC Press: New York, NY, USA, 2007. [Google Scholar]
  41. Lipatov, Y.S.; Nesterov, A.E. Thermodynamics of Polymer Blends; Technomic Publishing Co. Inc.: Lancaster, PA, USA, 1997. [Google Scholar]
  42. Kim, J.Y.; Cho, H.; Noh, S.; Lee, Y.; Nam, Y.M.; Lee, C.; Jo, W.H. Charge transport in amorphous low bandgap conjugated polymer/fullerene films. J. Appl. Phys. 2012, 111, 043710. [Google Scholar] [CrossRef]
  43. Kim, J.Y.; Frisbie, C.D. Correlation of Phase Behavior and Charge Transport in Conjugated Polymer/Fullerene Blends. J. Phys. Chem. C 2008, 112, 17726–17736. [Google Scholar] [CrossRef]
  44. Cullity, B.D.; Stock, S.R. Elements of X-ray Diffraction; Prentice Hall: Upper Saddle River, NJ, USA, 2001. [Google Scholar]
  45. Dörling, B.; Sánchez-Díaz, A.; Arteaga, O.; Veciana, A.; Alonso, M.I.; Campoy-Quiles, M. Controlled Pinning of Conjugated Polymer Spherulites and its Application in Detection. Adv. Opt. Mater. 2017, 5, 1700276. [Google Scholar] [CrossRef]
  46. Ou, C.; Cheetham, N.J.; Weng, J.; Yu, M.; Lin, J.; Wang, X.; Sun, C.; Cabanillas-Gonzalez, J.; Xie, L.; Bai, L.; et al. Hierarchical Uniform Supramolecular Conjugated Spherulites with Suppression of Defect Emission. iScience 2019, 16, 399–409. [Google Scholar] [CrossRef] [Green Version]
  47. Armas, J.A.; Reynolds, K.J.; Marsh, Z.M.; Fernández-Blázquez, J.P.; Ayala, D.; Cronin, A.D.; Del Aguila, J.; Fideldy, R.; Abdou, J.P.; Bilger, D.W.; et al. Supramolecular Assembly of Oriented Spherulitic Crystals of Conjugated Polymers Surrounding Carbon Nanotube Fibers. Macromol. Rapid Commun. 2019, 40, 1900098. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of (a) P(NDI2OD-T2) (poly{(N,N′-bis(2-octyldodecyl)naphthalene -1,4,5,8-bis(dicarboximide)-2,6-diyl)-alt-5,5′-(2,2′-bithiophene)}) and (b) PCPDTBT (poly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta [2,1-b;3,4-b′]dithiophene) -alt-4,7(2,1,3-benzothiadiazole)]), respectively.
Figure 1. Chemical structures of (a) P(NDI2OD-T2) (poly{(N,N′-bis(2-octyldodecyl)naphthalene -1,4,5,8-bis(dicarboximide)-2,6-diyl)-alt-5,5′-(2,2′-bithiophene)}) and (b) PCPDTBT (poly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta [2,1-b;3,4-b′]dithiophene) -alt-4,7(2,1,3-benzothiadiazole)]), respectively.
Polymers 12 01726 g001
Figure 2. TGA curves for P(NDI2OD-T2), P(NDI2OD-T2)/PCPDTBT blend, and PCPDTBT.
Figure 2. TGA curves for P(NDI2OD-T2), P(NDI2OD-T2)/PCPDTBT blend, and PCPDTBT.
Polymers 12 01726 g002
Figure 3. DSC curves. (a) Heating-cooling-heating curves for P(NDI2OD-T2). (b) 1st heating, (c) 1st cooling, and (d) 2nd heating curves for P(NDI2OD-T2):PCPDTBT = 100:0, 80:20, 50:50 and 0:100 (wt. ratio), respectively.
Figure 3. DSC curves. (a) Heating-cooling-heating curves for P(NDI2OD-T2). (b) 1st heating, (c) 1st cooling, and (d) 2nd heating curves for P(NDI2OD-T2):PCPDTBT = 100:0, 80:20, 50:50 and 0:100 (wt. ratio), respectively.
Polymers 12 01726 g003
Figure 4. Melting and crystallization point depression of P(NDI2OD-T2)/PCPDTBT blends.
Figure 4. Melting and crystallization point depression of P(NDI2OD-T2)/PCPDTBT blends.
Polymers 12 01726 g004
Figure 5. XRD patterns of P(NDI2OD-T2):PCPDTBT blends with 100:0, 80:20, 50:50 and 0:100 (wt. ratio).
Figure 5. XRD patterns of P(NDI2OD-T2):PCPDTBT blends with 100:0, 80:20, 50:50 and 0:100 (wt. ratio).
Polymers 12 01726 g005
Figure 6. Schematic explanation of when semicrystalline P(NDI2OD-T2) and amorphous PCPDTBT polymers are mixed together through solution processing. Here, the regular π-π stacking could be destroyed by adding PCPDTBT into a crystalline P(NDI2OD-T2) lamellae.
Figure 6. Schematic explanation of when semicrystalline P(NDI2OD-T2) and amorphous PCPDTBT polymers are mixed together through solution processing. Here, the regular π-π stacking could be destroyed by adding PCPDTBT into a crystalline P(NDI2OD-T2) lamellae.
Polymers 12 01726 g006
Figure 7. Tapping-mode AFM images of (a) pristine P(NDI2OD-T2), (b) annealed P(NDI2OD-T2), (c) pristine P(NDI2OD-T2):PCPDTBT = 1:1 blend, (d) annealed P(NDI2OD-T2):PCPDTBT = 1:1 blend, (e) pristine PCPDTBT, and (f) annealed PCPDTBT films, respectively.
Figure 7. Tapping-mode AFM images of (a) pristine P(NDI2OD-T2), (b) annealed P(NDI2OD-T2), (c) pristine P(NDI2OD-T2):PCPDTBT = 1:1 blend, (d) annealed P(NDI2OD-T2):PCPDTBT = 1:1 blend, (e) pristine PCPDTBT, and (f) annealed PCPDTBT films, respectively.
Polymers 12 01726 g007
Table 1. Crystallite size (t) and d-spacing of P(NDI2OD-T2):PCPDTBT blends with 100:0, 80:20, 50:50, and 0:100 (wt. ratio) at the diffraction angle θ, when X-ray has wavelength (λ) of 0.154 nm. Herein, d-spacing between lattice planes is calculated based on Bragg’s law, λ = 2 d   sin   θ .
Table 1. Crystallite size (t) and d-spacing of P(NDI2OD-T2):PCPDTBT blends with 100:0, 80:20, 50:50, and 0:100 (wt. ratio) at the diffraction angle θ, when X-ray has wavelength (λ) of 0.154 nm. Herein, d-spacing between lattice planes is calculated based on Bragg’s law, λ = 2 d   sin   θ .
P(NDI2OD-T2):PCPDTBT2θ (°)θ (°)B (Radians)t (nm)d-Spacing (nm)
100:026.1613.080.056672.50.34
80:2026.3113.050.063192.30.34
50:5026.2613.130.066062.00.34
0:10025.4712.740.099801.40.35

Share and Cite

MDPI and ACS Style

Fanta, G.M.; Jarka, P.; Szeluga, U.; Tański, T.; Kim, J.Y. Phase Behavior of Amorphous/Semicrystalline Conjugated Polymer Blends. Polymers 2020, 12, 1726. https://doi.org/10.3390/polym12081726

AMA Style

Fanta GM, Jarka P, Szeluga U, Tański T, Kim JY. Phase Behavior of Amorphous/Semicrystalline Conjugated Polymer Blends. Polymers. 2020; 12(8):1726. https://doi.org/10.3390/polym12081726

Chicago/Turabian Style

Fanta, Gada Muleta, Pawel Jarka, Urszula Szeluga, Tomasz Tański, and Jung Yong Kim. 2020. "Phase Behavior of Amorphous/Semicrystalline Conjugated Polymer Blends" Polymers 12, no. 8: 1726. https://doi.org/10.3390/polym12081726

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop