Donor-Acceptor Sequence Engineering in π–Conjugated Polymers for

May 22, 2019 - Orcid Connecting Research and Researchers · Portico digital preservation service · ACS Publications. 1155 Sixteenth Street N.W.; Washin...
0 downloads 0 Views 1MB Size
Subscriber access provided by Bethel University

C: Energy Conversion and Storage; Energy and Charge Transport

Donor-Acceptor Sequence Engineering in #–Conjugated Polymers for Enhanced Hole Transport and Photocurrent Longhua Li, Baoxin Ge, Bolin Han, and Weidong Shi J. Phys. Chem. C, Just Accepted Manuscript • Publication Date (Web): 22 May 2019 Downloaded from http://pubs.acs.org on May 24, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Donor-Acceptor Sequence Engineering in π–Conjugated Polymers for Enhanced Hole Transport and Photocurrent Longhua Li*, Baoxin Ge, Bolin Han, Weidong Shi* School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, P. R. China

ABSTRACT Sequence-dependent properties of synthetic polymers are of great concern. To customizing the electronic and optical properties of copolymers, it is highly desirable to realize the quantitative sequence-property relationship. The sequence-dependent carrier transport and optical properties of the donor(D)–acceptor(A) PTB7 copolymer was studied through density functional theory (DFT) calculations. The exact enumeration structures in lexicographical order with periodic boundary conditions were constructed by a high-throughput approach. The absorption coefficient, effective mass, deformation potential, and hole mobility exhibits a strong response to the sequence-patterns. Significantly, the bi/multi-block pattern [DnAn] shows much better hole mobility and short-circuit current than the well-known [(DA)n] pattern. For example, hole mobility of [DnAn] is 12%–46% larger than the [(DA)n]; and the calculated short-circuit current of [DnAn] is 1.5–2.5 times of the [(DA)n]. Our work uncovered the reason for this enhancement: the intrinsic band splittings of the additional D–D/A–A interactions could induce strong absorption coefficient and electron delocalization. The concept of additional D–D/A–A combinations may be generally applicable and could provide an additional degree of freedom for the design of donor-acceptor polymers with tunable properties.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 18

INTRODUCTION The physical steps in a solar cell can be simplified to the absorption of photons, exciton dissociation, and charge transport to electrodes. Thus, an organic polymer for a solar cell should have a high optical absorption coefficient, an efficient exciton separation, and large carrier mobility with minimized recombination. The morphology, backbone conformation, dopant, and side-chain were commonly known to play an important role in manipulating carrier mobility of polymers.1,

2

Recently, there is also increasing interest in precisely controlling monomers

sequence toward tuning polymers properties.3-5 The sequence engineering would expand the library of polymers to meet requirements in fields like energy, medicine, and data storage.6, 7 To date, However, only several kinds of patterns are realized in AnBm polymers due to the challenge of synthesis, for instance, alternating sequence (e.g., AB…AB), bi-/multi-block (AnBn…AnBn, n≥2), gradient distribution (AnB1An-1B2…A1Bn+1), tapered type (AmB1AnB2An1…BnA1Bm)

and position insertion (e.g., AmBAn).8-10 Indeed, the attractive differences in

properties were found within such a limited list of sequence architectures,11-15 suggesting that the fine-tuning sequence would offer an exciting opportunity for property design. Unfortunately, the available number of sequences in a copolymer chain can be varied by length. For instance, a copolymer (AnBm, n≥1, m≥1) that contains n A and m B monomers could generate permutations up to (n + m)! / (n! × m!). Though this is not exactly the case in practice because many sequence structures may be identical if applied proper symmetries, it remains true that the number of patterns would increase with chain length. Such increasing of sequence-patterns with length scales causes unprecedented difficulty in enumerating all possible sequence-solutions. While the effect of monomer ratios have commonly been addressed in previous studies,16,

17

the full

permutation of sequences attracts very little attention.18 Hutchison group19 has employed SIMILES20 notation system and Open Babel21 to construct sequence structures of oligomers for photovoltaic application. They have found sequence-dependent frontier orbital energy levels in tetramers and hexamers. However, the interrelation between sequence and carrier mobility and optical absorption in the donor-acceptor copolymers is still vague or unknown. In this work, the photovoltaic polymer that consists of benzodithiophene (BDT) and fluorothieno[3,4-b]thiophene (FTT) was studied. The BDT and FTT are the typical donor and acceptor motifs that have been broadly used in donor-acceptor (D-A) conjugated copolymers in the last decade.22 The most studied binary blend may be the poly[[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-

ACS Paragon Plus Environment

Page 3 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

b:4,5-b′]dithiophene-2,6-diyl][3-fluoro-2-[(2-ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl]] (PTB7)

and

poly[4,8-bis(5-(2-ethylhexyl)thiophen-2-yl)benzo[1,2-b:4,5-b’]dithiophene-co-3-

fluorothieno[3,4-b]thiophene-2-carboxylate](PTB7-Th), because they exhibit superior solar energy power conversion efficiency (PCE)

of about 8–12%.23-26 The sequence-pattern

dependence of charge transport and optical properties was systematically investigated by density functional theory (DFT) calculations. All available sequence-patterns in polymers BDTnFTTn (n ≤ 6, about 80 structures) were generated by our lexicographical search method. We have set up an integrated picture of the sequence-property relationship. Our results indicate that the property of traditional D–A polymer can be further improved by sequence manipulating. For example, [DnAn] patterns show an enhanced photocurrent density of 0.75–1.1 mA/cm2, which is at least 1.6 times of the traditional [(DA)n] pattern. In addition, [D2A2], [(D2A2)2], [D3A3], and [D3A3DA] patterns exhibit higher hole mobility than the corresponding [(DA)n] pattern. We propose that the sequence-pattern may become an additional degree of freedom in customizing the D–A polymer.

METHODS AND COMPUTATIONAL DETAILS A FORTRAN code GenLS (generate lexicographical structures) was developed by implementing the lexicographical algorithm and geometry operations. Assuming there are n BDT and FTT in the periodic unit cell, the problem becomes to discover full permutations of the sequence from [DnAn] to [AnDn]. The utilization of periodic boundary conditions has been widely realized for the investigation of D-A copolymers.27-30 Mathematically, the unique ordered set with n nodes can be enumerated lexicographically.31 However, this full permutation is lack of physical meanings, i.e., it may still involve a large number of identical structures from the view of physics. We, therefore, included symmetry operations in our code. Take a general sequence [ABCDE] (in periodic boundary conditions, PBC) for instance, which should be equal to [EDCBA], [EABCD], [DEABC], [CDEAB] and [BCDEA]. In order to verify such operations, we performed calculations on the BDT2FTT2 (D2A2) polymer, which has six sequences: [DDAA], [DADA], [DAAD], [ADDA], [ADAD] and [AADD]. As expected, the total energy and frontier orbital (HOMO and LUMO) energy level of [DDAA], [DAAD], [ADDA] and [AADD] are almost same to each other; [DADA] and [ADAD] are also same (see in Figure S1, SI). Such symmetry operations, therefore, help to eliminate those sequence-patterns that defined in the same structured way. This code not only can be used to generate full permutations structures

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

between two moieties, but also can be used to generate general polymers in periodic or nonperiodic forms, such as (N1)n1(N2)n2(N3)n3(N4)n4… with different kinds of components (N1, N2…) and numbers (n1, n2…). And it can also be extended to generate vertical/lateral stacking structures of two-dimensional (2D) materials. There are two advantages of our developed framework: (1) it can enumerate all sequences in order without duplication or omission; (2) this code also offers a high-throughput solution for processing large amounts of structures that required in DFT calculations. The generated structures were fully optimized by the density functional theory (DFT) linear combination of atomic orbitals (LCAO) method as implemented in Atomistix ToolKit (ATK) packages.32 A double-ζ polarized (DZP) basis set was used. Core electrons were described by the Troullier−Martins norm-conserving pseudopotential with the Kleinman−Bylander nonlocal projector.33, 34 The exchange-correlation energy was treated with the Perdew−Burke−Ernzerhof (PBE) version of the generalized gradient approximation (GGA).35 A 1 × 1 × 3 Monkhorst−Pack (MP)36 k-mesh was adopted. The optimization performed until the residual forces on atoms were less than 0.01 eV/Å. After the structure optimization, the ground state was calculated with a 1 × 1 × 4 Gamma centered k-mesh. The optical absorption coefficient was evaluated through:

  2 c

12   22  1 2

(1)

Where ω and c are the wave frequency and the speed of light, respectively. The real (ε1) and imaginary (ε2) part of dielectric constant was derived from the Kubo-Greenwood formula, using a denser k-mesh of 1 × 1 × 16. We further quantify the photocurrent density by the flux of absorbed photons (Jabs), assuming every photon is converted to a unit of charge carrier. Jabs was calculated by integrating the following function from the optical gap to 280 nm (the standard maximal frequency at the reference air mass 1.5 spectra, AM1.5G):37

J abs  e  A( ) J ref ( )d 

(2)

Where A(ω) is expressed as A( )  1  exp( L) and the layer thickness L = 1 nm is using. Jref (ω) is the incident photon flux (photons/cm2·s·eV), which is obtained from the reference AM1.5G spectra.38 By multiplying the elementary charge e (1.60217662 × 10-19 coulombs), the absorbed photon flux becomes to the theoretical maximum possible photocurrent density (units of mA/cm2), that is, Jabs is the upper limit contribution to the solar cell short-circuit current.

ACS Paragon Plus Environment

Page 4 of 18

Page 5 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Charge mobility (μ) was explored by the deformation potential (DP) theory,39 assuming electron–phonon coupling is weak compared with the intermolecular electronic couplings in this system and only the lattice scattering is included. In one-dimensional (1D) systems, μ can be simply using the effective mass approximation:40

1D 

eh2C

 2 kBT 

1/2

m*me

3/2

 EDP 

2

 2 E  C   2  a0    dE EDP  edge d

(3–5)

Where ħ, kB, T, me, a0, m*, C, EDP are Plank constant, Boltzmann constant, temperature, the mass of an electron, equilibrium length of the 1D system, the effective mass of hole or electron, elastic coefficient, and deformation potential, respectively. The elastic coefficient C can be calculated by a quadratic relation between total energies (E) and strains (ε, ±3%), and the deformation potential EDP for hole/electron was evaluated by a linear fit of the highest occupied molecular orbital (HOMO)/lowest unoccupied molecular orbital (LUMO) with strains from –3% to 3%. The DP model has been widely applied in π–conjugated molecular/organometallic systems.41, 42 Thus, this model may also be reasonable for the copolymers studied in this paper. RESULTS AND DISCUSSION The moieties of the donor (BDT) and acceptor (FTT), as well as the sequence-patterns for DnAn (n = 1–5), are presented in Figure 1. Patterns of D6A6 are too many to be illustrated here, which is shown in Table S1 (SI). From the GenLS code, we found that the permutation number increases markedly with the number of donor (acceptor) in the backbone (Figure S2, SI). For example, 1, 2, 3, 8, 16, 50, 133, 440, 1387 and 4752 effective permutations are obtained in BDTnFTTn (DnAn, n = 1–10). The huge number of structures is beyond our computational power when n > 6. Therefore, lexicographical structures from [DnAn] to [(DA)n] (n = 1–6) were explored to demonstrate the effect of D–A sequence on the electronic and optical properties of conjugated polymers.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. (a) Structures of the donor (BDT, D) and acceptor (FTT, A), the side chains are simplified to H; (b) backbone lexicographical patterns in periodic boundary conditions (PBC) for conjugated polymers BDTnFTTn; blue and orange blocks represent donor and acceptor, respectively. The total of 50 sequence-patterns of D6A6 is shown in Table S1. Sequence-pattern dependent electronic and optical properties The sequence-pattern dependence of the electronic structure is presented in Figure 2a. The [(DA)n] have similar HOMO and LUMO level (horizontal dash line), because they are the supercell of [DA]. For the [DnAn] pattern, the HOMO tends to increase with the number of donor/acceptor, but is not linearly dependent on n, wherein, the HOMO level of [D2A2] ([D4A4]) is similar to that of [D3A3] ([D5A5]). On the side of the LUMO, it decreases fast from [DA] (–3.6 eV) to [D2A2] (–3.9 eV) and then converges to about –4.0 eV in [DnAn] (n ≥ 3). In each copolymer, the traditional [(DA)n] pattern shows the lowest HOMO energy and the highest LUMO energy among all patterns; while the [DnAn] pattern often shows the highest HOMO and the lowest LUMO. By projecting the density of states (DOS) into each monomer, we found that the degenerate states in the conventional [(DA)n] sequence (Figure 2b) are split-off in [DnAn] (Figure 2c) and other sequences (Figure S3). This is due to the fact that additional D–D and A–A couplings are formed in [DnAn] and other patterns compared with the traditional [(DA)n] sequence. Such band splitting induced small band gap and multiple electronic states around the Fermi level are critical for the optical dipole transitions, which we shall discuss later. In order to understand the effect of sequence on frontier orbit energy, the HOMO and LUMO of [DnAn] are plotted as an example in Figure 2d. Generally, strong anti-bonding means high HOMO level while the enhancement of bonding strength means redshift of LUMO. According to Figure 2d, [DnAn] (n≥2) patterns exhibit stronger π‒π anti-bonding states as compared to [(DA)n],

ACS Paragon Plus Environment

Page 6 of 18

Page 7 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

which are contributed not only from BDT-TT but also from TT-TT and BDT-BDT. Therefore, [DnAn] (n≥2) shows higher HOMO than the corresponding [(DA)n]. It is also noticed that the localization of HOMO density increases with the number of nodes, especially notable in BDT units. Such hole charge localization should reduce hole mobility that would be displayed later. On the side of LUMO, the σ–σ bonding strength in the [DnAn] is stronger than that in the [(DA)n] due to the σ states of the [DnAn] are not only comprised of BDT-TT but also of TT-TT and BDTBDT. Moreover, in [DnAn] (n≥3), the LUMO is similar to each other and is predominated by σ bonding states among 3 TT and 1 BDT units. Thus, the LUMO of the [(DA)n] is higher than the [DnAn], and [DnAn] (n≥3) shows a similar LUMO level.

Figure 2. (a) HOMO and LUMO energy of polymers BDTnFTTn; the horizontal dash line is the value of [(DA)n] pattern and the vertical dash line represents the [DnAn] pattern; (b) projected density of states (PDOS) of the traditional [DADADADA] sequence and (c) multi-block 𝑆 ” means bands splitting. (d) HOMO [DDDDAAAA] sequence; the Fermi level is set to zero; “□

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 18

and LUMO for [DnAn] patterns; the isosurface value is set to 0.01 e*Bohr3; the cell boundary is not shown for clarity. The absorption coefficient is critical to the absorption of photons. Here, we take the optical absorption of D4A4 polymer as an example to illustrate the change of absorption spectra with DA pattern. It should be noticed that our GGA obtained optical gap is about 1.1 eV (Figure 2) for the normal [(DA)n] pattern, which is much smaller than the G0W043 and experimental44 value of 1.65 eV. In order to stay consistent with the previous studies, we assume that the same error is introduced by GGA functional in all patterns and a rigid shift of 0.55 eV is applied in Figure 2. Although our assumption is simple, the energy difference between them is independent of the shift. As shown in Figure 3a, p1–p8 represent the eight patterns of D4A4 polymer: [D4A4], [D3A3DA], [D3A2DA2], [D2A3D2A], [D2A2(DA)2], [D2ADA2DA], [(D2A2)2] and [(DA)4]. We found that, as compared to the traditional [(DA)4] pattern (the black line), (1) the absorption coefficient of all other patterns are enhanced in the violet, yellow and orange regions; (2) however, the absorption coefficient is significantly reduced in the blue and red regions; (3) in the green region, only the coefficients of p4 ([D2A3D2A]) and p5 ([D2A2(DA)2]) are enhanced; (4) all other patterns exhibit a much stronger optical absorption coefficient in the near-infrared region (NIR) than the traditional [(DA)4], because of the relative small band gap. In principle, the efficiency of the generation of electron-hole pairs is mainly determined by the optical absorption coefficient and absorption intensity in the visible and near-infrared regions. The absorption j

intensity is studied by integrating the area under the absorption line (αω), I    d  . The i

change of absorption intensity is evaluated relative to the [(DA)n] pattern. In the visible region (Figure S4), the [DnAn] pattern in each polymer, except the D2A2 polymer, shows the largest absorption intensity. The improvement of the absorption intensity is even more significant in NIR, see in Figure 3b. The absorption intensity of all other patterns is at least 1.5 times of the intensity of the [(DA)n] pattern; and the [DnAn] pattern in each polymer are shown the largest absorption intensity that is about 4.6–5.6 times of the [(DA)n] pattern. Such a significant enhancement of absorption intensity, especially in NIR, might result in a positive benefit for the generation of electron-hole pairs.

ACS Paragon Plus Environment

Page 9 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

In order to quantify the conversion of photons to the electric current (photocurrent density), the Jabc is calculated in Figure 3c. The [(DA)n] generate photocurrent around 0.45 mA/cm2; the photocurrent of [DnAn] patterns increase with monomer number n, from 0.75 to 1.08 mA/cm2; and other patterns generate photocurrents in the 0.6–1.0 mA/cm2 range. It is straightforward to see that all other patterns display stronger photocurrents than the traditional [(DA)n] pattern, and [DnAn] patterns present the largest photocurrent density among all patterns. The photocurrent of 1 nm thickness DnAn polymer is much smaller than that of transition metal dichalcogenides (2– 4.5 mA/cm2).37 However, we notice that there is a tendency of increasing photocurrent with increasing n, i.e., molecular weight in a unit cell. It is then expected that the photocurrent may be improved further by increasing the molecular weight (backbone length) of [DnAn] pattern. We, further, take the traditional alternating pattern [(DA)4] and multi-block [D4A4] as an example to interpret the optical property in D–D/A–A combinations. As the band structures are shown in Figure 3d, the band splittings of multi-block [D4A4] are significant, especially, the splitting between the highest (lowest) valence (conduction) band (VB1/CB1) and the second highest (lowest) valence (conduction) band (VB2/CB2). The band splittings of VB1-VB2 and CB1-CB2 reduce the band gap between VB1 and CB1, and this contributes to the absorption coefficient in NIR. Extra band splittings could result in additional subbands in the same energy range, which creates more inter-band transitions. In order to know the contribution of each bandband transition to the optical absorption coefficient, the band-resolved optical absorption is investigated by increasing VB and CB step by step. The contribution of band-pairs is clearly shown in Figure 3e and 3f. For the traditional pattern [(DA)4], the absorption spectrum is relatively simple, only three optical absorption bands are formed by the transitions between VB(1–4) and CB(1–4). The first absorption band (P1) are essentially due to the transitions of VB1–CB1, because P1 band does not change with the increase of conduction bands; the second and third absorption bands, P2 and P3, are mainly originated from VB2–CB2 and VB3(VB4)– CB3 transitions, respectively. For [D4A4], the effect of band splitting is pronounced; at least 10 optical absorption bands are associated with the eight electronic bands. The absorption bands of P1–P4 are dominated by the inter-band transitions from VB1 to CB(1–4), respectively; and the P5–P8 bands are originated from VB2 to CB(1–4), respectively; the strong peak P9 is mainly due to VB3–CB1 transitions; and the weak P10 is derived from VB4–CB2. Our results suggest that the additional D–D and A–A couplings induced band splitting is critical for the electronic and

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

optical properties of conjugated polymers. We believe that by further optimizing the concentration and chemical species of D–D and A–A combinations, such polymers may express greatly enhanced electronic and optical properties compared to the traditional onefold D–A couplings.

Figure 3. (a) Optical absorption of various D-A sequences in D4A4 polymer; p1–p8 represents patterns: [D4A4], [D3A3DA], [D3A2DA2], [D2A3D2A], [D2A2(DA)2], [D2ADA2DA], [(D2A2)2], [(DA)4]; the color bar from left to right represents the visible region (violet, blue, green, yellow, orange and red) and near-infrared region (NIR); a blue shift of 0.55 eV was applied for all patterns; (b) the change of the absorption intensity (as compared to the [(DA)n] pattern) in the NIR from 750 to 1500 nm; (c) the dependence of photocurrent density on D-A sequences; (d) band structures of [(DA)4] and [D4A4]; the first four valence bands (VB) and conduction bands (CB) are marked as VBn and CBn; the main inter-band transitions are presented as Pn (n = 1–10); (e) and (f) band-resolved optical absorption for [(DA)4] and [D4A4], respectively; the symbol “Pn:Vm-Ct” means that the absorption peak Pn mainly origins from the transitions between the valence band m and conduction band t.

ACS Paragon Plus Environment

Page 10 of 18

Page 11 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Sequence-pattern dependent hole transport Charge transport is a key factor in optimizing optoelectronics performance. Here, we focus on the hole mobility (μh) of studied sequences. The elastic coefficient C, which is the result of strain induced total energy shift, seems varied with D–A patterns (Figure 4a). However, the variation is slight, from the minimal 53.5 eV/Å to the maximal 55.5 eV/Å, that is, the magnitude of change in C is much less than 3.6%. Thus, the coefficient C may be regarded as a constant. The average C is 54.5 eV/Å, and the μ is simplified as:

1D 

eh2C

 2 kBT 

1/2

* 3/2

me3/2 m

 EDP 

2



431.88 m*

3/2

 EDP 

2

(6)

The difference between the result of the Eq. (3) and Eq. (6) is given in Figure 4d. The negligible difference between the Eq. (3) and Eq. (6) indicates that our assumption is reasonable. Our results suggest that the elastic coefficient is not observable dependent on the D–A sequence. A large μh could be obtained if a small mh* and EDP was achieved. The mh* is varied significantly with the D–A sequence as shown in Figure 4b. The mh* of [(DA)n] patterns (the horizontal dashed line) should be similar (0.25m0 – 0.27m0) in the studied polymers. However, the mh* of [DnAn] (n ≥ 2) patterns (the vertical dashed line) increase exponentially with n (see the insert in Figure 4b). This may be due to the Dn–An couplings (see HOMO in Figure 2d, the hole distributions become more localized with increasing n) decrease with n. The significant influence of hole localization on the reduction of hole mobility has also been discussed in other conjugated copolymers.45 It is interesting to see that mh* of all other patterns are smaller than that of the regular [(DA)n] pattern when n ≤ 4. In particularly, the [DDAA], [DDDAAA], [DDAADDAA], [DDDAADDAAA] and [DDAADDAADDAA] sequences display the smallest mh* in D2A2 (0.162m0), D3A3 (0.153m0), D4A4 (0.152m0), D5A5 (0.176m0) and D6A6 (0.164m0) copolymers, respectively. Thereby, our calculations indicate that the well-defined pattern, [(D2A2)n] (n≥1 for even D and A) or [(D3A3)(D2A2)n] (n≥0 for odd D and A), has the smallest mh* for a given copolymer with 1:1 D-A ratio. The effect of D–A sequence on DP of the hole is plotted in Figure 4c. DP can be obtained from the strain induced front molecular orbital (FMO) energy shift, which is expressed by E(FMO)=E(DP)*ε+E0, where ε is the strain on the backbone and E0 is initial energy. A high μh requires a small DP of the hole, which means that HOMO should be as insensitive as possible to the variation of the strain. All other patterns show much larger E(DP) than the [(DA)n]. Thereby,

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the HOMO of the conventional pattern [(DA)n] is much rigid as compared to other patterns, indicating again the electronic states of D–D/A–A combinations are more delocalized compared to the onefold D–A couplings. Dependence of μh on D–A sequence is shown in Figure 4d. Because all other patterns show much larger E(DP) than the [(DA)n], most patterns display smaller μh than the traditional [(DA)n] polymer, ~195 cm2V-1s-1. According to the Eq. (6), the μh can be larger than that of [(DA)n] only if mh* and E(DP) satisfied the following condition:

mh*

3/2

2 * E (DP)  mh* ([D A])

3/2

2 * E (DP) ([D A])  2.16

Only [D2A2] (2.03), [D3A3] (1.55), [(D2A2)2] (1.59) and [D3A3DA] (1.98) patterns exhibit higher μh than the [(DA)n] pattern, which is mainly due to they have relative small mh*. Our DP results provide a solid support for the previous assumption18 that D–D combinations may improve charge mobility. It should be noticed that the calculated hole mobility of PTB7 ([DA] pattern) is approximately 5–6 orders of magnitude larger than the experimental value (~0.2×10-4 – 5.9×10-4 cm2V-1s-1).46 This overestimation may be ascribed to the defect-free and rigid backbones, i.e., ideal one-dimensional wires, used in the calculations. Besides, only the intrachain mobility along the backbone was evaluated in our calculations. In fact, the consequence of the “trap-free” space charge limited current (SCLC) measurement47 is not only included the intra-chain charge transport but also the inter-chain (mostly van der Waals interactions) charge transport. Intra-chain mobility is usually several orders of magnitude higher than the inter-chain mobility.48, 49 Therefore, the calculated values can be regarded as the upper limit of intra-chain mobility. The overestimations of the deformation potential model have also been reported previously in π-conjugated systems.50 Nevertheless, the deformation potential model can describe qualitative differences in the mobility of various backbone patterns.

ACS Paragon Plus Environment

Page 12 of 18

Page 13 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 4. The effect of D–A sequence on (a) elastic coefficient; (b) effective hole mass; (c) deformation potential of hole; and (d) hole mobility of DnAn (n=1–6) copolymers; bar (–) represent hole mobility calculated with elastic coefficient of 54.5 eV/Å.

CONCLUSIONS In summary, we use density functional theory based electronic structure calculations to explore systematically the effect of donor–acceptor (D–A) sequence on the electronic and optical properties of conjugated polymer PTB7. A sequence search method is developed for the purpose of generating lexicographical D–A sequences. The DnAn (D: BDT; A: FTT; n ≤ 6) polymers were taken as the example to demonstrate that the electronic and optical properties fine-tuning can be achieved through programming the backbone patterns. Our results show that the welldefined pattern, [(D2A2)n] (n≥1 for even D and A) or [(D3A3)(D2A2)n] (n≥0 for odd D and A), has the smallest effective hole mass among all patterns. The [D2A2], [D3A3], [(D2A2)2] and [D3A3DA]

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

patterns exhibit higher hole mobility than the traditional [(DA)n] pattern. Although effective mass, deformation potential, and hole mobility are pattern-dependent, it is interesting to find that the elastic coefficient can be treated as a constant. Moreover, the photocurrents of [DnAn] patterns linearly increase with n from 0.75 to 1.08 mA/cm2, which are significantly larger than that of the traditional [(DA)n] pattern, 0.45 mA/cm2. Based on the above results, we found that the multi-block pattern [D3A3] and bi-block [(D2A2)2] may be the optimized structures to achieve high hole mobility and photocurrent density in PTB7. The enhancement of electronic and optical properties is found positively related to the subbands of D–D and A–A combinations. Note that even though the calculations are performed for PTB7, our results can be used to understand the trends in the electronic and optical properties of other DA polymers. The band splittings of D– D/A–A couplings indicate that the electronic and optical properties fine-tuning may be achieved through programming sequence and the ratio between D–A and D–D/A–A combinations. ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Including: total energy and HOMO-LUMO level of D2A2; the effective permutation number of DnAn; PDOS of sequence-patterns in D4A4 polymer; the absorption intensity in the visible region; and the table of sequences of DnAn for n=1–6. CONFLICT OF INTEREST The authors declare no conflict of interest. Corresponding author [email protected]; [email protected] ACKNOWLEDGEMENTS All calculations were performed on National Supercomputing Center in Tianjin and HPC of Jiangsu University. This work was supported by National Natural Science Foundation of China (No. 11604125); Natural Science Foundation of Jiangsu Province (No. BK20160522); China Postdoctoral Science Foundation (No. 2017M611705).

ACS Paragon Plus Environment

Page 14 of 18

Page 15 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

REFERENCES (1) Higgins, A.; Mohapatra, S. K.; Barlow, S.; Marder, S. R.; Kahn, A. Dopant Controlled Trap-Filling and Conductivity Enhancement in an Electron-Transport Polymer. Appl. Phys. Lett. 2015, 106, 163301. (2) Botiz, I.; Schaller, R. D.; Verduzco, R.; Darling, S. B. Optoelectronic Properties and Charge Transfer in Donor–Acceptor All-Conjugated Diblock Copolymers. J. Phys. Chem. C 2011, 115, 9260-9266. (3) Lutz, J.-F.; Ouchi, M.; Liu, D. R.; Sawamoto, M. Sequence-Controlled Polymers. Science 2013, 341, 1238149. (4) Lutz, J.-F.; Lehn, J.-M.; Meijer, E.; Matyjaszewski, K. From Precision Polymers to Complex Materials and Systems. Nature Rev. Mater. 2016, 1, 16024. (5) Chen, Z.; Lichtor, P. A.; Berliner, A. P.; Chen, J. C.; Liu, D. R. Evolution of Sequence-Defined Highly Functionalized Nucleic Acid Polymers. Nat. Chem. 2018, 10, 420-427. (6) Lutz, J.-F. Coding Macromolecules: Inputting Information in Polymers Using Monomer-Based Alphabets. Macromolecules 2015, 48, 4759-4767. (7) Zydziak, N.; Konrad, W.; Feist, F.; Afonin, S.; Weidner, S.; Barner-Kowollik, C. Coding and Decoding Libraries of Sequence-Defined Functional Copolymers Synthesized Via Photoligation. Nat. Commun. 2016, 7, 13672. (8) Lin, F.; Wang, M.; Pan, Y.; Tang, T.; Cui, D.; Liu, B. Sequence and Regularity Controlled Coordination Copolymerization of Butadiene and Styrene: Strategy and Mechanism. Macromolecules 2017, 50, 849856. (9) Huang, Z.; Noble, B. B.; Corrigan, N.; Chu, Y.; Satoh, K.; Thomas, D. S.; Hawker, C. J.; Moad, G.; Kamigaito, M.; Coote, M. L.; Boyer, C.; Xu, J. Discrete and Stereospecific Oligomers Prepared by Sequential and Alternating Single Unit Monomer Insertion. J. Am. Chem. Soc. 2018, 140, 13392-13406. (10) Gody, G.; Zetterlund, P. B.; Perrier, S.; Harrisson, S. The Limits of Precision Monomer Placement in Chain Growth Polymerization. Nat. Commun. 2016, 7, 10514. (11) Zhang, S.; Hutchison, G. R.; Meyer, T. Y. Sequence Effects in Conjugated Donor–Acceptor Trimers and Polymers. Macromol. Rapid Commun. 2016, 37, 882-887. (12) Zhang, S.; Bauer, N. E.; Kanal, I. Y.; You, W.; Hutchison, G. R.; Meyer, T. Y. Sequence Effects in Donor–Acceptor Oligomeric Semiconductors Comprising Benzothiadiazole and Phenylenevinylene Monomers. Macromolecules 2017, 50, 151-161. (13) Jin, C.; Wei, Z.; Yu, Y.; Sui, M.; Leng, X.; Li, Y. Copolymerization of Ethylene Brassylate with ΔValerolactone Towards Isodimorphic Random Copolyesters with Continuously Tunable Mechanical Properties. Eur. Polym. J. 2018, 102, 90-100. (14) Rosales, A. M.; Segalman, R. A.; Zuckermann, R. N. Polypeptoids: A Model System to Study the Effect of Monomer Sequence on Polymer Properties and Self-Assembly. Soft Matter 2013, 9, 8400-8414. (15) Zhang, W.; Lu, X.; Mao, J.; Hsu, C.-H.; Mu, G.; Huang, M.; Guo, Q.; Liu, H.; Wesdemiotis, C.; Li, T.; Zhang, W.-B.; Li, Y.; Cheng, S. Z. D. Sequence-Mandated, Distinct Assembly of Giant Molecules. Angew. Chem. 2017, 129, 15210-15215. (16) Boschetto, G.; Xue, H.-T.; Dziedzic, J.; Krompiec, M.; Skylaris, C.-K. Effect of Polymerization Statistics on the Electronic Properties of Copolymers for Organic Photovoltaics. J. Phys. Chem. C 2017, 121, 25292538. (17) Hung, Y.-C.; Chao, C.-Y.; Dai, C.-A.; Su, W.-F.; Lin, S.-T. Band Gap Engineering Via Controlling Donor– Acceptor Compositions in Conjugated Copolymers. J. Phys. Chem. B 2013, 117, 690-696. (18) Kanal, I. Y.; Owens, S. G.; Bechtel, J. S.; Hutchison, G. R. Efficient Computational Screening of Organic Polymer Photovoltaics. J. Phys. Chem. Lett. 2013, 4, 1613-1623. (19) Kanal, I. Y.; Bechtel, J. S.; Hutchison, G. R., Sequence Matters: Determining the Sequence Effect of Electronic Structure Properties in Π-Conjugated Polymers. In Sequence-Controlled Polymers: Synthesis, Self-Assembly, and Properties, American Chemical Society: 2014; Vol. 1170, pp 379-393.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(20) Weininger, D. Smiles, a Chemical Language and Information System. 1. Introduction to Methodology and Encoding Rules. J. Chem. Inf. Comput. Sci. 1988, 28, 31-36. (21) O'Boyle, N. M.; Banck, M.; James, C. A.; Morley, C.; Vandermeersch, T.; Hutchison, G. R. Open Babel: An Open Chemical Toolbox. J. Cheminformatics 2011, 3, 33. (22) Holliday, S.; Li, Y.; Luscombe, C. K. Recent Advances in High Performance Donor-Acceptor Polymers for Organic Photovoltaics. Prog. Polym. Sci. 2017, 70, 34-51. (23) Sun, J.; Zhang, Z.; Yin, X.; Zhou, J.; Yang, L.; Geng, R.; Zhang, F.; Zhu, R.; Yu, J.; Tang, W. High Performance Non-Fullerene Polymer Solar Cells Based on Ptb7-Th as the Electron Donor with 10.42% Efficiency. J. Mater. Chem. A 2018, 6, 2549-2554. (24) Xiao, Z.; Jia, X.; Li, D.; Wang, S.; Geng, X.; Liu, F.; Chen, J.; Yang, S.; Russell, T. P.; Ding, L. 26 Ma Cm−2 Jsc from Organic Solar Cells with a Low-Bandgap Nonfullerene Acceptor. Sci. Bull. 2017, 62, 1494-1496. (25) Lee, J. H.; Kim, K. M.; Jang, W.; Ahn, S.; Kim, Y. Y.; Park, O. O.; Wang, D. H. Vacuum-Process-Based Dry Transfer of Active Layer with Solvent Additive for Efficient Organic Photovoltaic Devices. J. Mater. Chem. C 2017, 5, 1106-1112. (26) Xie, X.; Liu, G.; Cheng, G.; Liu, Z.; Lee, E.-C. Improving Performance of Organic Solar Cells by Supplying Additional Acceptors to Surface of Bulk-Heterojunction Layers. J. Mater. Chem. C 2018, 6, 2793-2800. (27) Sun, Y.; Zhang, C.; Dai, B.; Lin, B.; Yang, H.; Zhang, X.; Guo, L.; Liu, Y. Side Chain Engineering and Conjugation Enhancement of Benzodithiophene and Phenanthrenequnioxaline Based Conjugated Polymers for Photovoltaic Devices. J. Poly. Sci. Part A: Poly. Chem. 2015, 53, 1915-1926. (28) Niskanen, M.; Hukka, T. I. Modeling of Photoactive Conjugated Donor–Acceptor Copolymers: The Effect of the Exact Hf Exchange in Dft Functionals on Geometries and Gap Energies of Oligomer and Periodic Models. Phys. Chem. Chem. Phys. 2014, 16, 13294-13305. (29) Pappenfus, T. M.; Schmidt, J. A.; Koehn, R. E.; Alia, J. D. Pbc-Dft Applied to Donor−Acceptor Copolymers in Organic Solar Cells: Comparisons between Theoretical Methods and Experimental Data. Macromolecules 2011, 44, 2354-2357. (30) Liu, X.; He, R.; Shen, W.; Li, M. Theoretical Design of Donor-Acceptor Conjugated Copolymers Based on Furo-, Thieno-, and Selenopheno[3,4-C] Thiophene-4,6-Dione and Benzodithiophene Units for Organic Solar Cells. J. Mol. Model. 2013, 19, 4283-4291. (31) Ord-Smith, R. J. Algorithms: Algorithm 323:Generation of Permutations in Lexicographic Order. Commun. ACM 1968, 11, 117. (32) Piqueras, M. C.; Crespo, R.; Tomás, F. Modulation of the Electronic Properties of Conjugated Polymers through Design of Polymer Backbone. J. Mol. Struc-THEOCHEM 1995, 330, 181-185. (33) Kleinman, L.; Bylander, D. M. Efficacious Form for Model Pseudopotentials. Phys. Rev. Lett. 1982, 48, 1425-1428. (34) Troullier, N.; Martins, J. L. Efficient Pseudopotentials for Plane-Wave Calculations. Phys. Rev. B 1991, 43, 1993-2006. (35) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865-3868. (36) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188-5192. (37) Bernardi, M.; Palummo, M.; Grossman, J. C. Extraordinary Sunlight Absorption and One Nanometer Thick Photovoltaics Using Two-Dimensional Monolayer Materials. Nano Lett. 2013, 13, 3664-3670. (38) Gueymard, C. A.; Myers, D.; Emery, K. Proposed Reference Irradiance Spectra for Solar Energy Systems Testing. Sol. Energy 2002, 73, 443-467. (39) Bardeen, J.; Shockley, W. Deformation Potentials and Mobilities in Non-Polar Crystals. Phys. Rev. 1950, 80, 72-80.

ACS Paragon Plus Environment

Page 16 of 18

Page 17 of 18 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(40) Beleznay, F. B.; Bogár, F.; Ladik, J. Charge Carrier Mobility in Quasi-One-Dimensional Systems: Application to a Guanine Stack. J. Chem. Phys. 2003, 119, 5690-5695. (41) Kocherzhenko, A. A.; Patwardhan, S.; Grozema, F. C.; Anderson, H. L.; Siebbeles, L. D. A. Mechanism of Charge Transport Along Zinc Porphyrin-Based Molecular Wires. J. Am. Chem. Soc. 2009, 131, 55225529. (42) Xi, J.; Long, M.; Tang, L.; Wang, D.; Shuai, Z. First-Principles Prediction of Charge Mobility in Carbon and Organic Nanomaterials. Nanoscale 2012, 4, 4348-4369. (43) Li, L.-H.; Kontsevoi, O. Y.; Freeman, A. J. Electronic and Optical Excitations of the Ptb7 Crystal: FirstPrinciples Gw-Bse Calculations. Phys. Rev. B 2014, 90, 195203. (44) Wang, H. X.; Yu, X. F.; Yi, C.; Ren, H.; Liu, C.; Yang, Y. L.; Xiao, S.; Zheng, J.; Karim, A.; Cheng, S. Z. D.; Gong, X. Fine-Tuning of Fluorinated Thieno[3,4-B]Thiophene Copolymer for Efficient Polymer Solar Cells. J. Phys. Chem. C 2013, 117, 4358-4363. (45) Hoffmann, S. T.; Jaiser, F.; Hayer, A.; Bässler, H.; Unger, T.; Athanasopoulos, S.; Neher, D.; Köhler, A. How Do Disorder, Reorganization, and Localization Influence the Hole Mobility in Conjugated Copolymers? J. Am. Chem. Soc. 2013, 135, 1772-1782. (46) To, C. H.; Ng, A.; Dong, Q.; Djurišić, A. B.; Zapien, J. A.; Chan, W. K.; Surya, C. Effect of Ptb7 Properties on the Performance of Ptb7:Pc71bm Solar Cells. ACS Appl. Mater. Interfaces 2015, 7, 1319813207. (47) Blakesley, J. C.; Castro, F. A.; Kylberg, W.; Dibb, G. F. A.; Arantes, C.; Valaski, R.; Cremona, M.; Kim, J. S.; Kim, J.-S. Towards Reliable Charge-Mobility Benchmark Measurements for Organic Semiconductors. Org. Electron. 2014, 15, 1263-1272. (48) Yasutani, Y.; Honsho, Y.; Saeki, A.; Seki, S. Polycarbazoles: Relationship between Intra- and Intermolecular Charge Carrier Transports. Synth. Met. 2012, 162, 1713-1721. (49) Lee, W.; Lee, C.; Yu, H.; Kim, D.-J.; Wang, C.; Woo, H. Y.; Oh, J. H.; Kim, B. J. Side Chain Optimization of Naphthalenediimide–Bithiophene-Based Polymers to Enhance the Electron Mobility and the Performance in All-Polymer Solar Cells. Adv. Funct. Mater. 2016, 26, 1543-1553. (50) Ohto, T.; Masai, H.; Terao, J.; Matsuda, W.; Seki, S.; Tsuji, Y.; Tada, H. Enhancement of Carrier Mobility through Deformation Potential in Metal-Containing Insulated Molecular Wires. J. Phys. Chem. C 2016, 120, 26637-26644.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table of Content By exploiting the lexicographical sequence-patterns of donor (D)-acceptor (A) PTB7 copolymer, the quantitative sequence-property relationship is presented. The D–D/A–A combinations exhibits enhancement of hole mobility and photocurrent compared to the traditional D–A pattern.

ACS Paragon Plus Environment

Page 18 of 18