Introduction

Organic conductors that are lightweight, flexible, and have excellent molecular designability are widely used as essential materials in modern organic electronic devices. The most industrially successful materials are polymer conductors owing to their high conductivities and ease of synthesis1. However, analyzing their atomic-level structures and conduction mechanisms is still challenging because of the inhomogeneity of polymers with different chain lengths. To address this issue, researchers are investigating molecular conductors with precisely defined molecular weights2,3,4,5,6. The crystal-based structures help to understand the structure–conductivity relationships and address the conduction mechanism. Charge-transfer complexes3,4,5,6 are typical molecular conductors and are categorized as either “segregated-stack complexes,” in which donors (D) and acceptors (A) are separately stacked (DDDD…/AAAA…) or “(alternatingly) mixed-stack complexes,” in which donors and acceptors are alternately stacked (DADADADA….; Fig. 1a). The difference in the stacking forms of complexes composed of the common donors and acceptors dramatically affects the conductivities. Segregated-stack complexes, such as tetrathiafulvalene (TTF) analogs–7,7,8,8-tetracyano-p-quinodimethane (TCNQ) analogs (the molecular structures are shown in Supplementary Fig. 1)7,8, hexamethylenetetraselenafulvalene–TCNQ9, and TTF analogs–metal dithiolenes10, exhibit excellent room-temperature conductivities (σrt) of up to 103 S cm–17,9, although variations on such highly conducting examples are still limited5,6. Several hybrid materials between DD-stack and DA-mixed-stack complexes have shown relatively high σrt (10–2 S cm–1)11,12,13,14. In contrast, mixed-stack complexes are typically poor electrical conductors that show low σrt (below 10–4 S cm–1)15,16,17,18,19,20,21 (Fig. 1b), although they are more frequently constructed, possibly owing to the Madelung energy gain between charged donor and acceptor. Generally, mixed-stack complexes are either in a neutral state (charge-transfer degree from donor to acceptor δ < 0.54, mostly 0 < δ < 0.45,6) with insufficient conductive carriers and weak intermolecular interactions for carrier conduction, or are in a highly ionic state (mostly δ > 0.75)5,6, wherein the carriers are nearly localized and have low mobility, thereby suppressing conductivity. In the early 2000s, a few neutral complexes showed high conductivity22,23, possibly due to the structural disorders within the contaminated ionic domains. However, the structural details and mechanisms remained hidden as the CCDC data were unavailable. We hypothesized that the mixed-stack complexes at the neutral–ionic (N–I) boundary (0.59 < δ < 0.74)5 could have numerous mobile electrons and display excellent electrical conductivity24. The hypothesis was supported by the discovery of segregated-stack complexes with exceptional conductivity in the same boundary area5,6,7,8,9,10. Preceding studies24,25,26 predicted that charge transfer to form mixed-stack complexes with δ at the N–I boundary could be realized by engineering appropriate energy gaps between the highest occupied molecular orbital (HOMO) of donors and the lowest unoccupied molecular orbital (LUMO) of acceptors that satisfy the equation ID – EA (= ΔEREDOX) ≈ EM, where ID is the ionization potential of the donor (related to its HOMO of the donor), EA is the electron affinity of the acceptor (related to its LUMO of the acceptor), and EM is the electrostatic Madelung energy of the donor–acceptor complex (approximately 0.2 eV; Fig. 2). In addition, a consistent orbital symmetry between the donor HOMO and acceptor LUMO27 may be necessary to facilitate the hybridization of frontier orbitals and enable efficient charge transfer. Thus far, a few mixed-stack complexes have partly fulfilled these two requirements: 1) similar energy levels between the donor HOMO and the acceptor LUMO, with the appropriate energy gaps of approximately 0.2 eV, and 2) a consistent orbital symmetry between them; for example, tetrakis(methyltelluro)tetrathiafulvalene (TTeC1-TTF)–TCNQ (δ = 0.56; d’ in Fig. 1b)21 and 3,3’,5,5’-tetramethylbenzidine (TMB)–TCNQ (δ = 0.49; c’ in Fig. 1b)16,28 close to the N–I boundary exhibited relatively high σrt values of 10−2 and 10−5 S cm−1, respectively. Under high-pressure conditions (~9 kbar), δ of TTF–p-chloranil (CA) (v in Fig. 1b) was transiently increased to ≈0.6, exhibiting excellent σrt (7 S cm−1)29, although it exhibits δ ≈ 0.2 and σrt = 10–5 S cm–1 under ambient pressures.

Fig. 1: Structures and electrical conductivities of mixed-stack complexes.
figure 1

a Molecular structures of donors (2X; X = O, S) and acceptors (Fn; n = 4, 2) in the mixed-stack complexes 2XFn. An illustration of the 1D packing mode of 2XFn is shown in the right. b σrt of 1D mixed-stack complexes as a function of charge-transfer degree from donor to acceptor (δ) determined for single-crystal structures. a: Benzothieno[3,2-b]benzothiophene (BTBT)–F415. b: 1,2-Di(2-thienyl)ethylene (DTE)–7,7,8,8-tetracyano-p-quinodimethane (TCNQ)5. c: Naphthalene–TCNQ5. d: Perylene–TCNQ5. e: Dithieno[3,2-b:2’,3’-d]thiophene (DTT)–TCNQ5. f: Anthracene–TCNQ5. g: trans-Stilbene–TCNQ5. h: trans-2,2’,5,5’-Tetramethoxystilbene–TCNQ5. i: Pyrene–TCNQ5. j: Bis(ethylenedithio)tetrathiafulvalene (BEDT-TTF)–TCNQ5. k: 2,3,7,8-Tetramethoxydibenzo[b,e][1,4]thiaselenine (Vn2SSe)–TCNQ6. l: 3,3’,5,5’-Tetramethylbenzidine (TMB)–p-chloranil (CA)16. m: Tetramethyltetraselenafulvalene (TMTSF)–TCNQ6. n: 2,3,7,8-Tetramethoxyselenanthrene (Vn2Se2)–TCNQ6. o: 1,6-Diaminopyrene (DAP)–p-bromanil (BA)22. p: 4,4’-Dimethyltetrathiafulvalene (DMTTF)–BA20. q: Coronene–TCNQ6. r: Dibenzotetrathiafulvalene (DB-TTF)–TCNQ5. s: Tetrathiafulvalene (TTF)–TCNQ17. t: DMTTF–CA20. u: 5,10-Dihydro-5-10-dimethylphenazine (DHMP)–TCNQ6. v: TTF–CA18. w: 2,5-Dichloro-1,4-phenylenediamine (Cl2PD)–2-chloro-5-methyl-7,7,8,8-tetracyano-p-quinodimethane (ClMeTCNQ)20. x: 2-Chloro-5-methyl-1,4-benzenediamine (ClMePD)–2,5-diethyl-7,7,8,8-tetracyano-p-quinodimethane (Et2TCNQ)20. y: α,α’-Dimethylquaterthiophene (DMQtT)–F419. z: ClMePD–2,5-dimethyl-N,N’-dicyanoquinone diimine (DMeDCNQI)20. a’: Tetrakis(methylthio)tetrathiafulvalene (TTC1-TTF)–2-fluoro-5-methyl-7,7,8,8-tetracyano-p-quinodimethane (FMeTCNQ)20. b’: TMB–2,5-dimethyl-7,7,8,8-tetracyano-p-quinodimethane (Me2TCNQ)16. c’: TMB–TCNQ16. d’: Tetrakis(methyltelluro)tetrathiafulvalene (TTeC1-TTF)–TCNQ21. e’: TMB–F216. f’: Phenazine–TCNQ5. g’: TMB–F416. The molecular structures are shown in Supplementary Fig. 1. The δ values for 2XFn are shown with error bars (standard deviation).

Fig. 2: Electronic and optical properties of mixed-stack complexes.
figure 2

Relationship between ΔEREDOX and CT that has two separate linear correlations in the neutral and ionic regions; the two straight lines intersect when ΔEREDOX ≈ EM, resulting in the CT value reaches to the minimum (approximately 0.6 eV). The vertical dashed line indicates the neutral–ionic boundary (Reprinted and modified from Reference26, Copyright 1985, from Taylor & Francis). A: N,N,N’,N’-Tetramethyl-p-phenylenediamine (TMPD)–F4. B: DHMP–TCNQ. C: TMPD–TCNQ. D: TMPD–CA. E: N,N,N’,N’-Tetramethyl-1,6-pyrenediamine (TMDAP)–TCNQ. F: TTF–CA. G: TTF–p-fluoranil (FA). H: Dibenzene TTF–TCNQ. I: Diethyldimethyltetraselenafulvalene–Et2TCNQ. J: TMDAP–FA. K: TTF–dichlorobenzoquinone. L: Perylene–F4. M: Perylene–2,3-dichloro-5,6-dicyano-p-benzoquinone. N: Perylene–tetracyanoethylene. O: Perylene–TCNQ. P: TTF–dinitrobenzene. Q: Perylene–CA. R: Pyrene–tetracyanoethylene. S: Pyrene–CA. T: Anthracene–CA. U: Hexamethylbenzene–CA. V: Naphthalene–TCNE. X: Anthracene–pyromellitic dianhydride (PMDA). Y: Anthracene–tetracyanobenzenze. Z: Phenanthrene–PMDA. The molecular structures are shown in Supplementary Fig. 1.

In this study, we designed and synthesized mixed-stack complexes uniquely located at and near the N–I boundary, enhancing the conductivities of mixed-stack complexes. The combination of the oligo(3,4-ethylenedioxythiophene) analog 2O30,31 or its oxygen/sulfur-substituted analog 2S with tetra- or difluorinated TCNQs Fn fulfills the two requirements for electronic structures (i.e., 2XFn, X = O, S in Fig. 1a). Surprisingly, the donor HOMO and acceptor LUMO with comparable energy levels and well-matched orbital symmetries were highly hybridized in the complexes, enhancing the σrt values of mixed-stack complexes (10−4–0.1 S cm−1). The highest σrt value (0.1 S cm−1) observed for 2SF4 among the combinations of donors 2X and acceptors Fn, which is located at the N–I boundary, is the highest value reported for a structurally defined 1D mixed-stack complex under ambient pressure. Furthermore, 2SF4 exhibited a transition in electrical conductivity to a slightly less conductive phase accompanied by significant changes in its optical, dielectric, and magnetic properties. This emphasizes the unique electronic structure of the mixed-stack complex at the N–I boundary that displays a structural perturbation between electron-itinerant and localized states.

Results and discussion

Molecular electronic structures

First, 2S with non-bulky methylthio groups was synthesized in 52% yield after three-step transformations from 2,2’-bi(3,4-ethylenedithiothiophene) 132 (Fig. 3a, Supplementary Note 3, Supplementary Figs. 7, 41, 42, and Supplementary Table 1). The electrochemical properties of 2O30 and 2S were analyzed via cyclic voltammetry (CV). The voltammograms revealed the first oxidation potentials at 0.622 and 0.805 V (vs. Ag/AgCl) for 2O and 2S, respectively (Supplementary Note 4 and Supplementary Fig. 3), estimating HOMO levels of 2O/2S (−4.94/−5.13 eV) that are comparable to the LUMO levels of F4/F2 (−4.98/−4.74 eV). The gaps between the donor HOMO and acceptor LUMO (i.e., ΔEREDOX in Fig. 2) are estimated to be in the range of −0.04 to 0.39 eV (Table 1), which is consistent with those predicted by density functional theory (DFT) calculations (Fig. 4a, Supplementary Note 5, Supplementary Figs. 4, 6, Supplementary Tables 6, 7, 9, and 11). The 1e-oxidized 2O/2S structures show highly symmetrical molecular orbital shapes spread over the entire molecules and consistent horizontally nodal patterns with an average periodicity of 2.0 Å that correspond well with those of 1e-reduced F4/F2 with an average periodicity of 1.7–1.8 Å (Fig. 4b, Supplementary Fig. 5, Supplementary Tables 8, 10, and 12). Therefore, the combination of 2O/2S as donors and F4/F2 as acceptors ideally has similar energy levels and orbital symmetries between the HOMO of the donor and LUMO of the acceptor, possibly leading to strong hybridization between their orbitals during complexation.

Fig. 3: Synthesis of mixed-stack complexes 2XFn.
figure 3

a Synthesis of donor 2S by bromination followed by lithiation and methylthiolation. b Synthesis of mixed-stack complexes 2XFn by chemical oxidation of donor (2O and 2S) by acceptors (F4 and F2). The δ values were determined using Kistenmacher’s equation37. NBS N-bromosuccinimide, nBu n-butyl, THF tetrahydrofuran, Me methyl.

Table 1 Structural information and physical properties of mixed-stack complexes
Fig. 4: Electronic structure of donor 2S, acceptor F4, and the mixed-stack complex 2SF4 according to theoretical calculations.
figure 4

a Energy levels of orbitals for neutral 2S donor and F4 acceptor calculated by the Gaussian09 program. The energy levels of the donor’s highest occupied molecular orbital (HOMO) and acceptor’s lowest unoccupied molecular orbital (LUMO) were comparable, which is appropriate for their strong hybridization. b Singly occupied molecular orbitals (SOMOs) of 1e-oxidized 2S donor and 1e-reduced F4 acceptor calculated by Gaussian09 program. These orbitals have horizontally nodal patterns with an average periodicity of 1.8–2.0 Å that correspond well with each other (Supplementary Fig. 5). c Band structure calculated by OpenMX38,39. The bandwidth (W) value was determined from the calculated density of states (DOS). The highest occupied crystal orbital (HOCO) and lowest unoccupied crystal orbital (LUCO) at the Γ-point are depicted in red and blue dots, respectively. Γ (0,0,0), P (–0.5,0,0.5), Q (–0.5,0.5,0.5), Y (0,0.5,0). d Highly hybridized HOCO and LUCO between donor and acceptor at the Γ-point calculated by OpenMX38,39. Atoms were colored as follows; white: hydrogen; gray: carbon; blue: nitrogen, red: oxygen; light green: fluorine; yellow: sulfur. Molecular and crystal orbitals with positive and negative phases were colored with magenta and navy, respectively.

Single-crystal X-ray diffraction analyses

Motivated by the potentially ideal combination of 2O/2S donors and F4/F2 acceptors in the electronic structures, we investigated their complexation. The donors and acceptors were mixed in dichloromethane or THF, which led to a gradual color change of the solutions from yellow to dark green, indicative of charge transfer from the donor to the acceptor. After slow solvent evaporation over more than three days, dark-green needle-like single crystals of complexes 2OF4, 2OF2, 2SF4, and 2SF2 were isolated (Fig. 3b). These syntheses were highly reproducible and scalable to a yield of several milligrams in a single sequence of operations. The salts had an astonishingly high solubility in organic solvents such as acetonitrile and were stable for at least several weeks under atmospheric conditions.

X-ray diffraction (XRD) analyses determined that all single-crystal structures belonged to the P21/n space group with inversion centers at the center of gravity of the donors and acceptors and the glide plane symmetry on the 1/2 translation operation along the (a + c)/2 direction (Supplementary Note 6, Supplementary Figs. 9, 10, 12, 14, and Supplementary Table 2). The 2O and 2S donors in single-crystal complexes are nearly planar (Supplementary Fig. 11), similar to the charge-transfer salts of 2O and 2S (i.e., 2O•BF430 and 2S•BF4; Supplementary Figs. 2, 8, and Supplementary Table 1). The fluorine atoms in 2OF2 and 2SF2 exhibited positional disorder, with an occupancy of 94:6 and 86:14, respectively (Supplementary Fig. 9b, d). The donors and acceptors are uniformly and alternatingly π-stacked in columns along the a-axis. The donor–acceptor interplanar distances depend on the molecular size of the donor (i.e., 2O < 2S) and acceptor (i.e., F2 < F4); the distances increased in the order 2OF2 (3.329 Å) < 2OF4 (3.361 Å) < 2SF2 (3.398 Å) < 2SF4 (3.406 Å) (Table 1). Significant intracolumnar short contacts were not observed, suggesting the existence of 1D electronic structures favorable for strong intermolecular interactions. Notably, X-ray diffuse scattering was observed, reminiscent of structural fluctuation (Supplementary Note 6 and Supplementary Fig. 14).

Charge-transfer degree of mixed-stack complexes

The δ values of the complexes were determined using the conventional bond length analyses of the single-crystal structures for donors and acceptors6 in the mixed-stack complexes. An increase in the δ value of the donor from 0 (i.e., a neutral form with a benzenoid structure) to 1 (a radical cation form with a resonance structure) weakens bond alternation (i.e., C–C bonds b and d and C–S bonds g are shortened, while C=C bonds a and c are lengthened; Fig. 5a, b, c, e, f), as is apparent from the single crystals for neutral 2X (X = O30 or S, δ = 0; Supplementary Fig. 7 and Supplementary Table 1) and its 1e-oxidized form 2X•BF430 (δ = 1; Supplementary Fig. 8 and Supplementary Table 1). All bond length alterations of the 2O and 2S donors in the mixed-stack complexes are located between those of 2X and 2X•BF4, indicative of intermediate δ values for the complexes between 0 and 1. Commonly, F4 complexes exhibited weaker bond length alternation than F2 complexes, indicating that the F4 complexes have higher δ values than the F2 complexes. Likewise, the acceptor molecules Fn exhibited weakening of the bond alternation (i.e., C–C bonds i, k, and m are shortened, while C=C bonds j and l are lengthened; Fig. 5a, d, g) upon reduction from neutral33,34 (δ = 0) to their 1e–reduced forms35,36 (δ = 1), suggesting the intermediate δ. Notably, the 2O complexes exhibited weaker bond length alternation than the 2S complexes, suggesting that the 2O complexes have higher δ values than the 2S complexes. These tendencies are consistent with the δ values determined using Kistenmacher’s equation37, which assumes that the bond length changes linearly as δ increases from 0 to 1. The δ values of 2O–F4, 2O–F2, 2S–F4, and 2S–F2 were determined to be 0.79(2), 0.71(4), 0.69(2), and 0.46(3), respectively (Table 1). These values are close to the N–I boundary (0.59 < δ < 0.74), whereas that of 2SF2 lies at the nearly neutral state, as predicted by the comparable energy levels of the donor HOMO and acceptor LUMO by CV analyses (Supplementary Fig. 3) and DFT calculations (Fig. 4a and Supplementary Fig. 6).

Fig. 5: Bond length analysis of mixed-stack complexes.
figure 5

a Bond labels of 2O, 2S, F4, and F2. bg Comparison of C–C and C–S bond lengths (dCC and dCS, respectively) of donors and acceptors in single crystals of mixed-stack complexes, with error bars (standard deviation). The letters along the x-axes of the plots correspond to the bonds depicted in a. As a reference, the bond lengths in single crystals of neutral donors or acceptors (i.e., δ = 0) are shown with dotted lines and 1e-oxidized donors or 1e-reduced acceptors (i.e., δ = 1) are shown with solid lines. Observed dCC of 2O30 (b), dCC of 2S (c), dCC of F433,35 (d), dCS of 2O30 (e), dCS of 2S (f), and dCC of F234,36 (g). i = (i1 + i2)/2.

Electronic structures of mixed-stack complexes

Based on the single-crystal structures of 2XFn as the average structures, we then obtained the band structures using first-principles calculations (OpenMX software; Supplementary Note 7)38,39, wherein the contributions from Coulomb repulsion between the carriers (discussed later) are ignored. All complexes showed band structures with highly dispersed bonding and antibonding orbital-derived bands formed by the donor HOMO and acceptor LUMO (Fig. 4c and Supplementary Fig. 18). The bands were exclusively dispersed along the a-axis (i.e., Γ–X corresponding to the π-stacking direction). The band structures had small or negligible energy gaps (Eg) in proximity to the Fermi level (0.02–0.05 eV for 2OF4 and 2SF2 and <0.01 eV for 2OF2 and 2SF4; Table 1 and Supplementary Fig. 18). The density of states (DOS) was calculated to determine the bandwidth (W) of the complexes in the major occupancies, the values of which were 0.86 eV (2OF4), 0.89 eV (2OF2), 0.88 eV (2SF4), and 0.89 eV (2SF2) (Table 1, Fig. 4c, and Supplementary Fig. 18), where W = WB + WA + Eg; WB and WA are the band dispersions of the bonding and antibonding orbitals, respectively. Among typical organic conductors2, high W values may be characteristic of 1D electronic structures, especially given the comparable energy levels of the donor HOMO and acceptor LUMO and their well-matched orbital symmetries, indicating strong intracolumnar interactions. We next quantified the transfer integrals between a donor and the six neighboring molecules in the single-crystal structures using combined Quantum ESPRESSO (QE)40,41 and RESPACK42 (Supplementary Fig. 27, and Supplementary Table 3). These calculations also identified prominent intracolumnar donor–acceptor interactions (tDA ≈ 0.21 eV in Table 1; t1 and t4 in Supplementary Table 3 and Supplementary Fig. 27), corresponding to W values based on the 1D tight-binding model, and negligible intercolumnar donor–acceptor and donor–donor interactions for all complexes. These findings support their 1D electronic structures, which were also suggested by the band structures. The W and intracolumnar tDA values follow nearly consistent trends: F4 complexes < F2 complexes, possibly because the donor–acceptor interplanar distances are shorter for F2 complexes (as shown in Table 1). Despite the longer interplanar distances, the trend 2O complexes < 2S complexes is primarily due to the presence of multiple S atoms with significant orbitals in the donor.

Notably, we have made a significant discovery using OpenMX calculations38,39, which identified that the HOMO of the donor and LUMO of the acceptor are highly hybridized. The calculated highest occupied crystal orbital (HOCO) and lowest unoccupied crystal orbital (LUCO) at the Γ point of 2XFn appeared equivalently on the donor and acceptor, respectively (Fig. 4d and Supplementary Fig. 19). This contrasts with neutral complexes3, where HOCOs and LUCOs were localized on the donors and acceptors, respectively. This finding supports that 2XFn complexes around the N–I boundary have different electronic structures from previously reported neutral or ionic mixed-stack complexes5,6,15,16,17,18,19,20,21,22,23, leading to significant W values based on their 1D electronic structures and unique electronic structures as discussed later.

Polarized reflectivity measurements

The polarized reflectivity spectra further support the observation that the δ values of 2XFn are close to the N–I boundary. The spectra obtained for the π-stacking direction of the single crystals have peak energies based on the charge-transfer band (CT) at 0.50 eV (2SF2) < 0.64 eV (2SF4) < 0.73 eV (2OF2) < 0.83 eV (2OF4) (Table 1 and Fig. 6a; complexes are located exactly at the N–I boundary, whereas 2SF2 is nearly neutral). The spectrum of 2SF2 has a shoulder peak at approximately 0.64 eV. The shape of this spectrum reflects the neutral-to-N–I boundary state43 (δ = 0.46(3)) or donor–acceptor π-dimerized form43,44 (discussed later), possibly causing the plot for 2SF2 to deviate from the predicted value by δ in Fig. 2. The significant positional disordering of fluorine atoms in 2SF2 may also affect the shape of the spectrum.

Fig. 6: Polarized reflectivity of mixed-stack complexes.
figure 6

a Spectra of mixed-stack complexes obtained in the electric field of light parallel to the π-stacking direction. Bars indicate the top of the peaks. The spectrum of the relatively thin single-crystal 2OF4 exhibits stripe-pattern signals in the low-energy region due to interference from the backside (the raw data is shown in Supplementary Fig. 34a), and thus this part of the spectrum was deleted for clarity. b Temperature-dependent reflectivity of 2SF4 in the electric field of light parallel (//) and perpendicular () to the π-stacking directions for the low-energy region.

Given that the strong hybridization between the donor HOMO and acceptor LUMO in the band structures with nearly negligible Eg, we considered the electronic structures of 2XFn to be like homogeneous donor- or acceptor-stacked charge-transfer salts, with the average structure2. Under this consideration, the observed CT may correspond to the difference between the bonding and antibonding orbital bands separated by carrier-to-carrier Coulomb repulsions in the solid states (Ueff), rather than the donor HOMO–acceptor LUMO energy gap for neutral complexes. Accordingly, the lower CT for the 2S complexes than that for the 2O complexes may correspond to lower Ueff for the 2S complexes. Quantum calculations provided further insights into electronic structures. The Ueff values for the complexes were quantified for donors and acceptors (i.e., Ueff (D) and Ueff (A), respectively) by combining first-principles calculations (QE40,41/RESPACK42 packages) based on the average single-crystal structures. These calculations identified the far lower Ueff (D) and Ueff (A) for 2S complexes compared to those of 2O complexes (Table 1, Supplementary Figs. 2026, and 2833), emphasizing that the combined effects of smaller δ value and the larger π-conjugate area of the donor in the complexes may favorably contribute to lowering Ueff. The cooperative reduction in Ueff (D) and Ueff (A) supports the significant orbital hybridization of donors and acceptors in single crystals. According to these calculations, the 2S complexes have slightly larger W and far lower Ueff than the 2O complexes, conferring superior electrical conductivities45,46,47,48 upon the 2O complexes. However, it is not negligible that nearly neutral 2SF2 has insufficient conductive carriers and positional disorders of fluorine atoms, which may impact the electronic structures that contribute to the conductivity (Supplementary Figs. 1820, 26, 27, 33, and Supplementary Table 3).

Structural fluctuations involved in π-dimerization

It is also notable that the polarized infrared reflectivity spectra of the complexes along the π-stacking direction exhibited multiple sharp peaks in the low-energy region of 0.14–0.19 eV, except for nearly neutral 2SF2 (Fig. 6b, Supplementary Figs. 34, and 35a). The optical conductivity spectrum of 2SF4 (Supplementary Fig. 35b) derived from the infrared spectrum showed a coincidence of the shapes to the Raman spectrum at 293 K (Supplementary Fig. 35d), with varying intensities. Considering the inversion center of the single-crystal structure in a P21/n space group with uniform π-stacked donors and acceptors along the C2 glide symmetry, the Raman-active modes should not be visible in IR, while the IR-active modes should not be visible in Raman. The coincidence of the spectral shapes indicates the dynamic fluctuating of donor–acceptor-π-stacking dimerization based on the electron-molecular vibration (EMV) couplings43,44. We performed a computational study by numerical simulations using the donor–acceptor dimer model49, confirming that the EMV coupling effect is responsible for these signals (Supplementary Note 8, Supplementary Fig. 35c, Supplementary Tables 4, and 5). These results suggest that there is inhomogeneity in the one-dimensional molecular stacking, contradicting the uniform columnar structure predicted by single-crystal XRD (Supplementary Table 2 and Supplementary Fig. 10). This inhomogeneity likely indicates dynamic fluctuations from the average structure, which is supported by the X-ray diffuse scattering observed in XRD at 300 K (Supplementary Fig. 14).

Electrical conductivity measurements

Motivated by the proximity of the δ value to the N–I boundary, high W, and low Ueff for 2S complexes, we subsequently investigated the electrical conductivity of single-crystal mixed-stack complexes (Supplementary Note 9). The σrt values (300 K) determined via the direct-current method along the π-stacking direction (the a-axis) were superior by a few orders of magnitude to those of previously reported typical mixed-stack complexes with σrt (300 K) below 10–4 S cm–1: 4.9 × 10−4 S cm−1 (2OF4), 1.4 × 10−2 S cm−1 (2OF2), 0.10 S cm−1 (2SF4), and 6.9 × 10−3 S cm−1 (2SF2) (Fig. 1b and Table 1) within the ohmic region (Supplementary Fig. 36). Except for the nearly neutral 2SF2, the trend of σrt (2SF4 > 2OF2 > 2OF4) indicates that the closeness of these complexes to the N–I boundary dominantly affected the conductivity. The σrt of 2SF4 (0.10 S cm−1) is the highest value reported to date for a single crystal of a 1D mixed-stack complex, and it is an order of magnitude higher than the previously highest value of 1.0 × 10−2 S cm−1 21. The low Ueff based on the large conjugate area may cooperatively contribute to the high σrt46,47,48, unlike those of the other complexes at the N–I boundary (e.g., complexes e’ and f’ in Fig. 1b).

The activation energies (Ea) near room temperature were determined from the temperature (T) dependence of the resistivity (ρ = σ−1) using the Arrhenius equation. The ρT plots are indicative of semiconducting behavior with relatively small Ea values: 0.113(1) eV (2OF4), 0.178(1) eV (2OF2), 0.200(1) eV (2SF4), and 0.112(1) eV (2SF2) around room temperature (Figs. 7, and 8a, and Table 1), indicating that these complexes are narrow-bandgap semiconductors. The Ea trend was inconsistent with those of σrt, the calculated Eg, and Ueff. Except for the nearly neutral 2SF2, the Ea values increased: 2OF4 < 2OF2 < 2SF4. The trend is explained by the proximity of δ to the N–I boundary which induces the π-dimerization fluctuation between the donor and acceptor; 2SF4 at the N–I boundary showed largest increase in ρ, and highest Ea among the three complexes.

Fig. 7: Temperature-dependent electrical conductivities of mixed-stack complexes.
figure 7

ρ–T plots obtained by a four-probe method for cooling (circles) and heating (plus marks) processes are shown.

Fig. 8: Temperature-dependent physical properties of single-crystal 2SF4 upon cooling.
figure 8

a Electrical resistivity. b Polarized reflectivity for photon energies of 144–178 meV in the electric field of light parallel to the π-stacking direction. c Real part of the dielectric constant εr at 1.0 MHz. d Normalized χspin determined from the calculation using the intensity and ΔBpp of the electron spin resonance (ESR) signals at g ≈ 2.003. The values are normalized by the value at 340 K, with error bars (standard deviation). e ΔBpp of the ESR signals at g ≈ 2.003, with error bars (standard deviation). Gray panel indicates the transition temperature around 282 K.

Abrupt changes in physical properties of 2S–F4 around room temperature

The mixed-stack complex with the highest conductivity at the N–I boundary, 2SF4, exhibited a unique transition with a sharp increase in ρ at a transition temperature (Tc) of 282 K. The transition was followed by an increase in Ea from 0.200(1) eV at high temperatures (288–340 K) to 0.277(3) eV at low temperatures (228–273 K) without significant hysteresis upon cooling and heating (Figs. 7, 8a, and Table 1). The XRD pattern in P21/n space group at 300 K showed significant X-ray diffuse scattering (Supplementary Fig. 14), possibly due to the precursor phenomenon for the π-dimerization, which may result in EMV coupling-based signals at room temperature (Fig. 6b, Supplementary Figs. 34, and 35a). On the other hand, the scattering disappeared (Supplementary Figs. 15 and 16) at 200 K, and the XRD showed superlattice patterns with dimensions of a × 2b × 2c (Supplementary Figs. 13 and 17), possibly implying the π-dimerization along the a-axis (Supplementary Note 6). The broken symmetry upon the possible π-dimerization of the complex may induce the markedly intensified EMV coupling-based signals (e.g., from 0.20 to 0.25) near 282 K during the cooling process while maintaining the energy (Figs. 6b, 8b, and Supplementary Fig. 35a), as well as the large anomaly in the dielectric constants50 (the real part of the dielectric constant is shown in Fig. 8c). The transition events did not cause any significant hysteresis for the cooling and heating processes.

Change in the magnetic characteristics at the transition

The magnetic characteristics revealed via temperature-dependent electron spin resonance (ESR) measurements helped us to address the insights into the electronic structures of the mixed-stack complexes (Supplementary Note 10). The spectra of 2SF4 exhibited a narrow Lorentzian signal with a peak-to-peak width (ΔBpp) of ≈0.2 G at g ≈ 2.003, which is typical for π-electrons2 (Supplementary Figs. 3739). At 282 K, the magnetic susceptibility (χspin) quantified from the spectra decreased sharply to reach a nearly nonmagnetic state (Fig. 8d and Supplementary Fig. 39b), suggesting a spin-Peierls-like singlet formation based on the 1D electronic structure51 during the π-dimerization. This χspin decrease was accompanied by an increase in ΔBpp (Fig. 8e and Supplementary Fig. 39c, d), supporting increased antiferromagnetic interactions in the singlet ground state in the low-temperature region. A decrease in χspin and increase in ΔBpp upon cooling were likewise observed for 2OF2 and 2OF4 at approximately 240 and 200 K, respectively (Supplementary Fig. 39b–d), which may correspond to the increase in Ea in the electrical conductivity at low temperatures (Figs. 7, 8a, and Table 1). The signals of 2SF2 were negligible mainly due to the relatively poor electron spins in the nearly neutral state. These magnetic behaviors were also consistent with the decrease in the magnetic susceptibilities (χ) of the polycrystals of 2SF4 and 2OF4 around the temperatures using a superconducting quantum interface device (SQUID) magnetometer (Supplementary Note 11 and Supplementary Fig. 40). The χ values of 2SF4 and 2OF4 were quantified to be 1.2 × 10−4 and 2.6 × 10−4 emu mol−1, respectively at 293 K. The χ values were relatively high compared to those of previously reported mixed-stack complexes18,50,52, most of which are in nearly nonmagnetic states. These values are comparable to those of typical organic conductors2, emphasizing the presence of abundant conductive electrons.

Surprisingly, in the case of 2SF4, another broad signal with a ΔBpp value of ~3 G appeared at g ≈ 2.004 below 282 K (Supplementary Figs. 3739), and overlapped with the narrow signal at g ≈ 2.003. This suggests the appearance of nearly non-correlated spins (e.g., spin solitons) originating from unpaired donors or acceptors during transient partial-π-dimerization between donor and acceptor. This observation suggests that the conductive mechanism of 2SF4 at the N–I boundary may involve spin solitons that exist because of topological excitations29 that contribute to its high conductivity.

The abrupt change in the electrical, optical, dielectric, and magnetic properties of 2SF4 at high Tc (282 K) under ambient pressures (Fig. 8) may reflect structural perturbation, which can be characteristic of the mixed-stack complexes at the N–I boundary. The neutral20,53 or highly ionic complexes50 transitioned based on π-dimerization to form a structure with more strongly ionic behavior at lower Tc values of 49–81 K. The complexes around the N–I boundary (δ = 0.3–0.5) had relatively high Tc values of 115–220 K20,52 under ambient pressures. Under high-pressure conditions (approximately 9 kbar), TTF–CA (δ ≈ 0.6) transiently exhibits excellent conductivity (7 S cm–1) and a high Tc at approximately 270 K29, whereas it has a low Tc of 81 K under ambient pressures (δ = 0.2, σrt = 105 S cm–1; v in Fig. 1b). The high Tc (282 K) and high σrt (0.1 S cm−1) even at ambient pressures of 2SF4, for which δ is exactly at the N–I boundary (δ = 0.69(2)) with non-hysterical changes in their electronic functionalities such as electrical conductivity, reflectivity, dielectricity, and magnetism may involve the unique electronic perturbation at the N–I boundary between the electron-itinerant and localized states.

In conclusion, we developed highly conductive 1D mixed-stack complexes for which δ is close to the N–I boundary by combining a donor HOMO and acceptor LUMO that have similar energy levels and well-matched orbital symmetries. These complexes have highly hybridized orbitals between the donor and acceptor and exhibited high σrt (10−4 to 0.1 S cm−1) under ambient conditions; the top σrt (0.1 S cm−1) for 2SF4 at δ = 0.69(2) is the highest value reported to date for a single-crystal 1D mixed-stack complex. 2SF4 uniquely exhibited a non-hysterical change in its electrical conductivity at high Tc (282 K) under ambient pressure to a less conductive phase below this temperature. This behavior was simultaneously accompanied by abrupt changes in the low-energy EMV coupling-based reflectivity, dielectric constant, and magnetic properties, reminiscent of the π-dimerization fluctuation that spatiotemporally breaks the symmetry and may widen the band gap to form a band-insulating state and increase Ea. Furthermore, the appearance of a broad ESR signal suggests non-correlated spins of unpaired donors or acceptors, which may enhance the conductivity of 2SF4 through mechanisms such as topological excitation29. In addition, mixed-stack complexes have unique advantages as conductor materials, such as high potential for solution processing and scalable synthesis without special equipment. The high conductivity, high solubility, and air stability of the conductors may provide scope for further investigation of the fundamental physical properties and applications using various external stimuli29,50,53,54. The rich variations in the structure-determining factors in oligomer donors, e.g., monomer unit variations, oligomer length55,56, sequences comprising multiple units56, and end-capping structures could enable precise tuning of the electronic structures toward highly functional mixed-stack complexes.

Methods

Apparatus

Proton (1H) and carbon (13C) nuclear magnetic resonance (NMR) spectra were obtained on a JEOL JNM-AL300 (1H NMR: 300 MHz; 13C NMR: 75 MHz) spectrometer. 1H NMR spectra in CDCl3 were referenced internally to tetramethylsilane as a standard. 13C NMR spectra in CDCl3 were referenced to the solvent resonance. High-resolution mass spectra (HRMS) were obtained on a JEOL JMS-AX500 with a field desorption (FD) probe in the positive mode using cholesterol as an internal standard. Single-crystal XRD measurements were performed using a Rigaku MercuryII CCD X-ray diffractometer (Mo Kα, λ = 0.71073 Å) and BL02B1 (λ = 0.30960 Å) at a synchrotron facility SPring-8 in Japan. A N2-gas-blowing device was employed for the low-temperature measurements. Cyclic voltammetry was performed using an ALS 610DB electrochemical analyzer. Microscopic mid-infrared (mid-IR) and ultraviolet-visible (UV-vis) measurements were performed on a JASCO FT/IR-6100 and FT/IR-6200 and MSV-5200TSO, both of which are equipped with a JASCO IRT-5000 FT/IR microscope. Raman spectroscopy was performed on a RENISHAW inVia Reflex. Dielectric constant measurements were performed using a custom-built alternating-current (AC) impedance probe with a Solartron Impedance Analyzer SI 1260, a Dielectric Interface 1296, and a Quantum Design SQUID magnetometer (MPMS-XL) as the temperature controller. Resistivity measurements were performed on a Quantum Design Physical Properties Measurement System (PPMS) and HUSO HECS 994C with a high-resistance-low-current electrometer KEYTHLEY 6517B equipped with an ADVANTEST R6142 power supply, an ADVANTEST R7210 scanner, and a KEYTHELY 2001 digital multimeter. X-band continuous wave ESR measurements were performed using a Bruker EMXmicro and EMX spectrometers with a N2 flow variable temperature unit ER4131VT and E500 spectrometer with a He flow cryostat ESR900. Static magnetic susceptibility measurements were performed using a MPMS-XL.

Synthesis

Synthesis of 5,5’-bis(methylthio)-2,2’-bi(3,4-ethylenedithiothiophene) 2S

General synthetic procedure and materials sources are shown in Supplementary Notes 1 and 2. To a solution of 2,2’-bi(3,4-ethylenedithiothiophene)32 1 (200 mg, 0.576 mmol) in dichloromethane (20 mL) was dropwise added at 0 °C a solution of N-bromosuccinimide (NBS; 238 mg, 1.34 mmol) in dichloromethane (6 mL), and the reaction mixture was stirred for 6 h. After the addition of a saturated aqueous solution of sodium bicarbonate (100 mL), the mixture was extracted with dichloromethane (3 × 100 mL). The combined organic layer was dried over sodium sulfate and concentrated in vacuo to afford a crude material containing 2 (250 mg) which was used for the subsequent reaction without further purification. Thus, to a solution of the mixture (250 mg) in THF (12 mL) was added nBuLi (1.6 M in hexane, 1.70 mL, 2.68 mmol) at −80 °C. The reaction mixture was stirred for 2 h and then allowed to warm to ambient temperature. To the mixture was dropwise added dimethyl disulfide (500 µL, 5.63 mmol), and the mixture was stirred at ambient temperature for 3 h. After the removal of volatile materials in vacuo and the subsequent addition of a saturated aqueous solution of sodium bicarbonate (100 mL), the mixture was extracted with dichloromethane (3 × 100 mL). The combined organic solvents were dried over sodium sulfate, concentrated in vacuo, and purified by GPC (eluent: chloroform) to afford 2S (131 mg, 0.298 mmol) as a yellowish white powder in 52% total yield for three-step transformations from 1. Physical data of 2S: 1H NMR (CDCl3, 300 MHz) δ 2.44 (s, 6H), 3.14–3.23 (m, 4H), 3.25–3.34 (m, 4H); 13C NMR (CDCl3, 75 MHz) δ 20.1, 27.6, 27.7, 126.6, 126.6, 127.8, 130.5; MS (FD) calcd for C14H14S8 [M+•] 437.8834, found 437.8856. The structural integrity and purity were identified by the NMR spectra (Supplementary Figs. 41 and 42). Block-like orange single crystals were obtained by the liquid–liquid diffusion method with dichloromethane/hexane (1:1, v/v). The single-crystal XRD analyses revealed the structural details (Supplementary Table 1 and Supplementary Fig. 7), confirming a nearly planar structure. In contrast, other solvent combinations afforded crystal-like solids, while their XRD patterns were too complicated to be resolved, suggesting that neutral 2S is polymorphic and could have multiple structures, potentially influenced by molecular twisting.

Typical procedure of mixed-stack complexes

A donor (2O30 or 2S, 10 µmol) was placed on one side of an H-shaped cell and an acceptor (F4 or F2, 10 µmol) was placed on the other side of the cell, respectively. To the cell were slowly added THF (8 mL for 2O–F4 and 2O–F2) or dichloromethane (8 mL for 2S–F4 and 2S–F2), and the mixture was kept at 25 °C in the dark. The slow evaporation of solvents from the mixture over more than three days gave dark green needle-like single crystals which were collected by paper filtration and washed with acetonitrile or 1:1 v/v dichloromethane/hexane (typical size of ~500 × ~40 × ~20 µm3). The single-crystal structures and chemical compositions were identified by single-crystal X-ray structural analyses (Supplementary Figs. 917 and Supplementary Table 2).

CV measurements

We performed CV in a degassed 0.6 mM solution of 2O30 or 2S in dichloromethane containing 100 mM nBu4N•PF6. We used a glassy carbon as the working electrode, a platinum wire as the counter electrode, and a silver–silver chloride electrode (Ag/AgCl in 1 M KCl) as the reference electrode (Supplementary Fig. 3).

Theoretical calculations

Energy levels and shapes of orbitals for donors and acceptors

The energy levels of orbitals for neutral donors (2O30 and 2S), one-electron-oxidized donors (2O•+ and 2S•+), neutral acceptors (F4 and F2), and one-electron-reduced acceptors (F4•– and F2•–) were calculated. The calculations were performed using the optimized structures on the Gaussian09 program (Revision D.01, Gaussian, Inc., Wallingford CT, 2016. https://gaussian.com/) at the DFT level with the (unrestricted) B3LYP functional, the gradient correction of the exchange functional by Becke57,58 and the correlation functional by Lee, Yang and Parr59 and the 6–31G(d) split valence plus polarization basis set60,61,62.

Band structure calculations

All the periodic DFT calculations were performed with the single-crystal structures of mixed-stack complexes as the average structures at 300 K. The calculations were performed by the OpenMX software38,39, based on optimized localized basis functions and pseudopotentials. The basis functions used were H6.0-s2p1, C6.0-s2p2d1, O6.0-s2p2d1, S7.0-s2p2d1f1, N6.0-s2p2d1, and F6.0-s2p2d1 for hydrogen, carbon, oxygen, sulfur, nitrogen, and fluorine, respectively, wherein the abbreviation of basis functions such as C6.0-s2p2d1, C stands for the atomic symbol, 6.0 the cutoff radius (Bohr) in the generation by the confinement scheme, and s2p2d1 means the employment of two, two, and one optimized radial functions for the s-, p-, and d-orbitals, respectively. The radial functions were optimized by a variational optimization method38,39. As valence electrons in the pseudopotentials (PPs), we included 1 s for hydrogen, 2 s and 2p for carbon, oxygen, nitrogen, and fluorine, and 3 s and 3p for sulfur, respectively. All the PPs and pseudo-atomic orbitals (PAOs) we used in the study were taken from the database (2019) in the OpenMX website (https://www.openmx-square.org/), which was benchmarked by the delta gauge method63. Real-space grid techniques were used for the numerical integrations and the solution of the Poisson equation using FFT with the energy cutoff of 220 Ryd64. We performed Brillouin-zone integrations on a 2 × 1 × 2 k-grid; the Fermi-Dirac distribution function at 300 K is employed as a smeared occupation function. We used a generalized gradient approximation (GGA) proposed by Perdew, Burke, and Ernzerhof to the exchange-correlation functional65. The real parts of the HOCO and LUCO at the Γ point (0,0,0) were visualized by VESTA66 (Fig. 4d and Supplementary Fig. 19).

Calculations of transfer integrals

The first-principles calculations were performed using the QE package40,41 with the GGA as the exchange-correlation function65 based on the single-crystal structure data of 2OF4, 2OF2, 2SF4, and 2SF2 at 300 K. In the calculations, SG15 Optimized Norm-Conserving Vanderbilt (ONCV) pseudopotentials67 were used as the pseudopotentials, in which 428 bands corresponding to the electron numbers were considered. The cutoff kinetic energies for wave functions, charge densities, and the mesh of the wave numbers were set as 80, 320 Ry, and 5 × 3 × 3, respectively. After the calculations, the maximally localized Wannier functions (MLWFs) were obtained using RESPACK42, and four bands near the Fermi energy were selected to reproduce the MLWFs (Supplementary Figs. 2126). Initial coordinates of the MLWFs were located at the center of gravity of crystallographically independent two donors and two acceptors in the unit cell, respectively. Supplementary Fig. 20 shows the energy bands near the Fermi energy obtained by QE40,41 and the Wannier interpolation. Based on the MLWFs, transfer integrals (t in Supplementary Fig. 27) were calculated as summarized in Table 1 and Supplementary Table 3.

Calculations of Coulomb interactions

Effective direct Coulomb interactions for donors and acceptors (i.e., Ueff (D) and Ueff (A)) were calculated by RESPACK42, in which the Coulomb interaction was reduced due to screening effects from bands other than the target band. Using a constrained random phase approximation (cRPA) method, we obtained the screened Coulomb interactions by incorporating screening effects from non-target bands. Through previous reports of such applications to different organic material systems46,47,48, we know that the effective model Hamiltonians derived by the cRPA method succeed in evaluating the stability of competing phases in electron-correlated systems. Thus, we estimated the Coulomb interactions by using the cRPA method. The energy cutoff for the dielectric function was set as 5.0 Ry. Static screened direct integrals were calculated as the matrix element of the shielded Coulomb interaction between the Wannier functions on the donors and acceptors, respectively (Supplementary Figs. 2126)42. The values at 0 Å were used as the Ueff (D) and Ueff (A) (Table 1 and Supplementary Figs. 2833).

Optical reflection spectroscopy measurements

Steady-state polarized reflectivity spectra of mixed-stack complexes (2OF4, 2OF2, 2SF4, and 2OF2) in the mid-IR region (0.12–0.97 eV) and UV-vis-near IR region (0.46–1.35 eV) were performed using mid-IR and UV-vis spectrometers equipped with a polarizer by applying an electric field (E) along the π-stacking directions (i.e., the a-axis) and perpendicular to the directions (Figs. 6, 8b, Supplementary Figs. 34, and 35a, b). Raman spectroscopy of 2SF4 was performed using an excitation wavelength of 532 nm (Supplementary Fig. 35d).

Dielectric constant measurements

AC impedance spectroscopy measurements to determine dielectric constant were carried out by the quasi-four-probe method in the frequency range of 1.0 MHz for 160–312 K. A direction parallel to the long axis direction of the crystal shape, corresponding to the π-stacking, was chosen for the AC impedance measurements. Electrodes were made on the two opposite sides of the samples using silver paste and gold wires (15 μm diameter).

Electrical resistivity measurements

Electrical resistivity (ρ) measurements of the single-crystal mixed-stack complexes were performed by the conventional a four-probe method along the long-axis direction of the crystal shape, corresponding to the π-stacking direction. Samples were prepared by attaching gold wires (15 µm diameter) of a single crystal with a conductive carbon paste. The ρ and conductivity (σ) values were derived from the following equation: ρ = σ–1 = R × S / L (S: cross-section, L: length). The typical size of samples was ~500 × ~40 × ~20 µm3.

ESR measurements

The X-band (9.4 GHz) continuous wave ESR experiments were performed on single-crystal 2OF4, 2OF2, 2SF4, and 2SF2 (Fig. 8d, e and Supplementary Figs. 3739). The ESR signals were measured by setting the long axis of the crystal (i.e., the stacking a-axis) along the direction parallel to the magnetic field (i.e., φ ~ 0° or 180°). The typical size of samples: ~500 × ~40 × ~20 µm3.

SQUID measurements

The static magnetic susceptibility of polycrystals 2OF4 and 2SF4 was measured upon cooling from 300 or 320 K to 2 K by <1 K min–1 applying the static magnetic field of 10,000 Oe (Supplementary Fig. 40). The absence of ferromagnetic impurity in the synthesized samples was confirmed from the obtained MH curve at 2 K around −55,000 to 55,000 Oe, which follows a typical Brillouin function. The obtained magnetic susceptibilities (χexp) were plotted after subtracting the contribution of the Curie impurity for S = 1/2 (χcw; 0.3% for 2OF4 and 0.2% for 2SF4) and the contribution from core diamagnetism of 2OF4 and 2SF4 (–3.14 × 10–4 and –3.74 × 10–4 emu mol–1, respectively) estimated from Pascal’s law68.