Solvent-Dependent Stoichiometry in Perylene–7,7,8,8

Oct 24, 2014 - (13) The electrical conductivities of single-crystal charge transfer compounds range ... TTF-TCNQ(24) (tetrathiafulvalene-tetracyanoqui...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/crystal

Solvent-Dependent Stoichiometry in Perylene−7,7,8,8Tetracyanoquinodimethane Charge Transfer Compound Single Crystals Peng Hu,† Lin Ma,‡ Ke Jie Tan,† Hui Jiang,*,† Fengxia Wei,† Chuhuan Yu,† Katelyn P. Goetz,§ Oana D. Jurchescu,§ L. E. McNeil,∥ Gagik G. Gurzadyan,‡ and Christian Kloc*,† †

School of Materials Science and Engineering, Nanyang Technological University, 639798 Singapore School of Physical and Mathematical Sciences, Nanyang Technological University, 637371 Singapore § Department of Physics, Wake Forest University, Winston-Salem, North Carolina 27109, United States ∥ Department of Physics and Astronomy, The University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599-3255, United States ‡

S Supporting Information *

ABSTRACT: Perylene forms three binary compounds with TCNQ. Using a physical vapor transport (PVT) method, mixtures of P1T1 (perylene1−TCNQ1), P2T1 (perylene2−TCNQ1), and P3T1 (perylene3−TCNQ1) crystals can be grown simultaneously. Solution-grown crystals exhibit a stoichiometry that depends on the solvent used and not on the initial stoichiometry of the perylene and TCNQ: only single crystals of P1T1 were grown from toluene, and only P3T1, from benzene. Precipitation−dissolution theory was employed to explain this phenomenon. We found that P1T1 is less soluble than P3T1 in toluene, which resulted in its preponderant growth from this solvent. Using the same argument, we could also explain the growth of P3T1 from benzene. Crystallization of P2T1 from solution was not possible. In addition, steady-state and timeresolved fluorescence measurements indicate that the difference in solubility of TCNQ in benzene and toluene leads to different fluorescence spectra due to selective crystallization of P1T1 and P3T1 from toluene and benzene.



the work on (BEDT-TTF)-(Cl2TCNQ)23 [(bis(ethylenedithio)tetrathiafulvalene)-(2,5-dichloro-7,7,8,8-tetracyanoquinodimethane)], TTF-TCNQ24 (tetrathiafulvalene-tetracyanoquinodimethane), and (BEDSe-TTF)-(F4TCNQ)25 [(bis(ethylene-diselena)tetrathiafulvalene-tetracyanoquinodimethane)], systematic studies of this topic remain rare. The majority of the reported organic charge transfer compounds are binary compounds. In contrast to the case of inorganic binary compounds, where well-documented libraries of phase diagrams provide useful information on the stoichiometry control and phase coexistence, it is striking that such phase diagrams for organic binary compounds are rather rare.26−31 The availability of this information, however, is deemed crucial for controlling the properties of single crystals of charge transfer binary compounds, which are strongly linked to their structures. One example is the degree of charge transfer, which is coupled to the electrical properties. That, in turn, is affected not only by the

INTRODUCTION

Single crystals of organic charge transfer compounds are suitable for the study of charge transfer effects1−5 due to their low concentration of structural imperfections (i.e., grain boundaries) and high purity.6−10 The degree of charge transfer between the donor and acceptor has been shown to play an important role in organic superconductivity,11 antiferromagnetism,12 electron− phonon interactions,12 and energy storage.13 The electrical conductivities of single-crystal charge transfer compounds range from insulating14 to semiconducting,15 metallic,16 and even superconducting,17 as we described in a recent review.18 Therefore, these compounds are considered for use in many applications including photovoltaic devices19−21 or field-effect transistors.22 Given the presence of two different chemical species, the donor and acceptor units, which exhibit different physical properties (solubilities, sublimation temperatures, etc.), crystal growth of charge transfer binary compounds is generally more complex that of monomolecular systems, especially for systems in which multiple stoichiometries are possible. Reports focusing on this subject are quite limited: with few exceptions, including © XXXX American Chemical Society

Received: August 14, 2014 Revised: October 16, 2014

A

dx.doi.org/10.1021/cg501206f | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Solubility Measurements for Perylene, TCNQ, P1T1, and P3T1. To measure the solubility of the charge transfer compounds and individual components, solvent was added to a known quantity of single crystal in a sealed test tube to avoid evaporation of solvent. The temperature was slowly increased to the target temperature. Then, the solution was stirred at target temperature overnight (longer than 20 h). Saturation was confirmed by the presence of residual undissolved crystals. Then, 1 mL quantities of the saturated solutions were placed into glass bottles and weighed before and after solvent evaporation. The solubility of the single crystal was calculated from the difference in weight. The same procedure was repeated for benzene and toluene. Characterization. The crystal structures of the P1T1 and P3T1 single crystals were obtained using a Bruker SMART APEX Π singlecrystal diffractometer (X-ray radiation, Mo Kα, λ = 0.71073 Å). The steady-state fluorescence emission and excitation spectra were recorded on a spectrofluorometer (Fluorolog-3, HORIBA Jobin Yvon). The timeresolved fluorescence measurements were carried out at room temperature by the time-correlated single photon counting (TCSPC) technique with a resolution of 10 ps (PicoQuant PicoHarp 300). The second harmonic of a titanium−sapphire laser (Chameleon, Coherent Inc.) at 400 nm (100 fs, 80 MHz) was used as the excitation source.

type of molecules of which the binary compound is built but also by the molecular packing and molecular stoichiometry.12,32 Recently, several studies also indicated that knowledge of the stability of stoichiometric complexes is essential in tuning their charge transfer properties.33−35 An understanding of the relation between the stability and charge transfer properties of co-crystals is crucial indeed for the successful growth of multiplestoichiometry compounds and control of stoichiometry transformations in such compounds. Perylene and TCNQ form three binary compounds with different stoichiometries, P1T1 (perylene1−TCNQ1), P2T1 (perylene2−TCNQ1),36 and P3T1 (perylene3−TCNQ1). Two stable compounds of perylene and TCNQ were previously reported: P1T137,38 and P3T1.39,40 These studies report on the effect of stoichiometry and molecular packing on the physical properties such as charge transfer and optical and electrical properties as a function of composition. Perylene41 is a donor with a high ionization energy42 (6.96 eV), and TCNQ (7,7,8,8tetracyanoquinodimethane) is a strong electron acceptor with a high electron affinity43 (3.0 eV). Charge transfer compounds formed from TCNQ with various donors exhibit unique properties such as metal-to-insulator transitions and exceptionally high conductivity.18,44 It was also shown that the stoichiometry of a binary organic charge transfer compound depends upon the stoichiometry of the starting materials.24,45 Nevertheless, in the present study, we observed an interesting phenomenon: the final stoichiometries of single-crystal P1T1 (perylene1−TCNQ1) and P3T1 (perylene3−TCNQ1) are not affected by the stoichiometry of the starting materials but by the solvent in which the perylene− TCNQ crystals were grown. The P1T1 crystals were grown from toluene, whereas the P3T1 crystals were grown from benzene, regardless of the acceptor/donor stoichiometry in the starting materials. In contrast, when utilizing the physical vapor transport (PVT) method, a mixture of monomolecular crystals, P1T1 (perylene1−TCNQ1), P2T1 (perylene2−TCNQ1), and P3T1 (perylene3−TCNQ1), is obtained.36 In this work, we grew P1T1 and P3T1 from solution (the only stoichiometries available using the solution growth method) and employed solubility data to analyze the effect of the choice of solvent on the stoichiometry of the perylene−TCNQ charge transfer single crystals. Steady-state optical spectra and time-resolved fluorescence were measured in a mixture of perylene and TCNQ in toluene or benzene, revealing that the difference in the solubilities of TCNQ in benzene and toluene leads to different fluorescence spectra due to selective crystallization of P1T1 and P3T1 from toluene or benzene.





RESULTS AND DISCUSSION The single-crystal charge transfer compounds (perylene)n(TCNQ)1 with perylene and TCNQ stoichiometries of 1:1 (P1T1) and 3:1 (P3T1) were grown from toluene and benzene, respectively, for all three ratios of starting materials. The stoichiometry of the final single crystals was found to be influenced only by the solvent, not by the stoichiometry of the starting materials. Furthermore, we dissolved P1T1 crystals in benzene to make a saturated solution at 80 °C and placed the test tubes in a temperature-controlled water bath. The initial temperature of the water bath was set at 80 °C, with a cooling rate of 0.5 °C/h. When the solution returned to room temperature, we obtained P3T1 crystals. Similarly, we dissolved P3T1 crystals in toluene to make a saturated solution at 80 °C and cooled it slowly to room temperature, obtaining P1T1 crystals. Needle-like P1T1 single crystals (Figure 1A) were

EXPERIMENTAL SECTION

Figure 1. Optical images of P1T1 crystals from toluene (A) and P3T1 crystals from benzene (B).

Single-Crystal Growth. All reagents (99% purity), including perylene, TCNQ, benzene, and toluene, were purchased from SigmaAldrich. The perylene and TCNQ were purified at 180 and 220 °C, respectively, via physical vapor transport (PVT).46,47 The solvents were used without further purification. Perylene and TCNQ with mole ratios of 3:1, 2:1, and 1:1 were weighed and placed in capped test tubes. Benzene or toluene was added to each test tube for crystal growth. After ultrasonication at room temperature for 15 min, the test tubes were placed in a temperature-controlled water bath. The initial temperature of the water bath was set at 80 °C, with a cooling rate of 0.5 °C/h. The starting materials were nearly all dissolved at 80 °C after 2 h. A small quantity of precipitated starting materials remained in the tube, indicating that the solution was saturated. Single crystal growth was observed with decreasing temperature, and crystals were isolated from the solution when the bath reached 20 °C.

grown from toluene, and cube-like P3T1 crystals (Figure 1B) were grown from benzene. The P2T1 stoichiometry was not obtained from any solvent investigated in this study. The molecular packing of P1T1 and P3T1 is depicted in Figure 2A,B. In the P1T1 crystals, the layers of perylene molecules are separated by layers of TCNQ molecules. The molecular packing of P3T1 differs significantly from that of P1T1: here, two perylene molecules are located between each pair of TCNQ molecules, and a third perylene molecule is located between the perylene−TCNQ stacks. Selected crystal structure information is listed in Table 1. B

dx.doi.org/10.1021/cg501206f | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 2. Molecular packing of P1T1 (A) and P3T1 (B).

in which [D] is the donor concentration and [A] is the acceptor concentration in the solvent. In a precipitation−dissolution equilibrium state, the following relationship between the donor and acceptor is attained:

Table 1. Crystallographic Data for P1T1 and P3T1 Derived Using Single-Crystal X-ray Diffraction formula crystal system space group lattice parameter a (Å) lattice parameter b (Å) lattice parameter c (Å) lattice parameter α (deg) lattice parameter β (deg) lattice parameter γ (deg) cell volume (Å3) formula units per cell Z temperature (K) R-factors

P1T1

P3T1

C32H16N4 Monoclinic P21/c 7.3116(19) 10.858(3) 14.500(4) 90 90.276(5) 90 1151.2(5) 2 293 0.0450

C72H40N4 Triclinic P1̅ 10.411(3) 10.862(1) 12.574(2) 66.456(1) 66.295(3) 89.225(4) 1175.1(4) 1 293 0.0496

m[D] = n[A]

Equation 2 can be presented as follows: K sp =

s=

{ }

⎡ n ⎤n [A]m = ⎢ ⎥ [A]m + n ⎣m⎦

(4)

[A] [D] = = m n

K sp

m+n

n

(n / m ) m

m+n

=

m+n

K sp mmn n

(5)

dissolve

(Perylene)1(TCNQ)1 XoooooooooY Perylene + TCNQ precipitate

(6)

dissolve

(Perylene)3 (TCNQ)1 XoooooooooY 3Perylene + TCNQ precipitate

(7)

Therefore, according to eq 5, the relationships between the solubility and the solubility product constant for P1T1 and P3T1 are s(P1T1) = s(P3T1) =

K sp

(8)

K sp

4

27

(9)

The enthalpy and entropy for reactions 6 and 7 can be calculated from the solubility product constant according to the following equation:49

(1)

The solubility product constant (Ksp) is K sp = [D]n [A]m

n

For P1T1 and P3T1, the precipitation−dissolution equilibrium equations are as follows:

dissolve

precipitate

n [A] m

The relationship between the solubility and solubility product constant is

Precipitation−dissolution equilibrium theory is widely used to explain the salt effect and common-ion effect in a separate precipitation of ionic compounds from solution.26,29,30,48,49 Recently, it has been successfully applied to organic cocrystals.26,50 Here, we adopted the same theory to understand the growth of the organic charge transfer compounds. Donors partly transfer electrons to acceptors; thus, the donors are partly positively charged (δ+), whereas the acceptors are partly negatively charged (δ−). Therefore, precipitation−dissolution equilibrium theory can also be applied to organic charge transfer systems. The charge transfer compound can be written as DnAm, in which D represents the donor, A, the acceptor, and n and m describe the stoichiometry of the donors and acceptors. Thus, the precipitation−dissolution equilibrium of DnAm in a given solvent can be presented as follows: DnA m XoooooooooY n D + m A

(3)

ln K sp = −

(2)

ΔH ΔS + RT R

(10)

Table 2. Solubility of P3T1 and P1T1 in Toluene at Different Temperatures (mol/L)

toluene benzene

P1T1 P3T1 P1T1 P3T1

40 °C

50 °C

60 °C

70 °C

80 °C

0.0006(7) 0.0029(6) 0.0025(4) 0.0014(1)

0.0008(3) 0.0031(7) 0.0025(8) 0.0016(6)

0.0008(6) 0.0037(4) 0.0030(2) 0.0017(6)

0.0009(2) 0.0048(3) 0.0032(9) 0.0021(1)

0.0010(1) 0.0056(7) 0.0046(4) 0.0026(4)

C

dx.doi.org/10.1021/cg501206f | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

°C, whereas the opposite is true for the values corresponding to P3T1. The Ksp data indicate that P1T1 is more soluble than P3T1 in benzene and that P3T1 is more soluble than P1T1 in toluene. The relationship between Ksp and solubility depends on the stoichiometry of the dissociation reaction.52 Accordingly, any direct comparison of Ksp values can be made only if the compounds have the same dissociation stoichiometry. To determine the thermodynamic data, the relationship between ln Ksp and 1/T was plotted in Figure 4. The slope and intercept were determined from the linear fit (see eq 10). The enthalpy (ΔH) and entropy (ΔS) are presented in Table 5. The solubility product constants calculated for P1T1 and P3T1 in toluene and benzene at 25 °C are presented in Table 6. From eq 1, the precipitation constant, which indicates the ability to form the charge transfer compound from the donor and acceptor, is the reciprocal of Ksp. The Gibbs energy can be determined by using the Ksp values at different temperatures. The temperature dependence of the Gibbs energy to form P1T1 and P3T1 in the two selected solvents is shown in Figure 5. The Gibbs energy for P1T1 in toluene (∼−40 kJ/mol) is lower than that in benzene (∼−30 kJ/mol); the Gibbs energy of P3T1 in benzene (∼−60 kJ/mol) is lower than that in toluene (∼−52 kJ/ mol). This is the reason that P1T1 is formed in toluene and P3T1, in benzene. The reaction quotient, Q, from reactions 6 and 7, can be calculated using concentrations of perylene in benzene or toluene and concentrations of TCNQ in benzene or toluene according to eqs 15 and 16.

Equation 10 is linear in 1/T, with a slope equal to −ΔH/R and a y intercept of ΔS/R. The enthalpy and entropy of dissolution can be calculated using the temperature dependence of the solubility product constant. The solubility data of P1T1 and P3T1 are listed in Table 2. Note that in solution the chemicals with low solubility will precipitate first.51 The solubility of P1T1 is greater than that of P3T1 in benzene and lower than that of P3T1 in toluene. Consequently, P1T1 will precipitate first in toluene, and P3T1 will precipitate first in benzene. The following equilibria are assumed to occur: precipitate

Perylene + TCNQ XoooooooooY (Perylene)1(TCNQ)1

(11)

dissolve

precipitate

3Perylene + TCNQ XoooooooooY (Perylene)3 (TCNQ)1

(12)

dissolve

In toluene, because P1T1 is less soluble than P3T1, the following equilibrium will exist: (Perylene)3 (TCNQ)1 ⇄ 3Perylene + TCNQ ⇄ (Perylene)1(TCNQ)1 + 2Perylene

(13)

Consequently, P1T1 forms in toluene, in agreement with our experimental observations. Similarly, in benzene, because P1T1 is more soluble than P3T1, the following equilibrium will exist: 3(Perylene)1(TCNQ)1 ⇄ 3Perylene + 3TCNQ ⇄ (Perylene)3 (TCNQ)1 + 2TCNQ

(14)

Consequently, P3T1 crystals form in benzene. All equilibria indicating the crystal growth routes are depicted in Figure 3.

Q(P1T1) = [perylene][TCNQ]

(15)

Q(P3T1) = [Perylene]3 [TCNQ]

(16)

Then, we compared Q with the Ksp values. If Q > Ksp, then precipitation will occur, but if Q < Ksp, then no precipitates will form. From Figure 6, the solubilities of perylene and TCNQ (concentration in the saturation solution) in toluene are 5.1 and 3.2 mmol/L at 25 °C, respectively. The Q values of P1T1 and P3T1 are 1.63 × 10−5 and 4.24 × 10−10, respectively. Compared with the data in Table 6, the reaction quotient is larger than the solubility product constant for P1T1 (1.63 × 10−5 > 8.88 × 10−8) and the reaction quotient is smaller than the solubility product constant for P3T1 (4.24 × 10−10 < 4.83 × 10−10) at room temperature. This indicates that P1T1 will form and P3T1 will dissolve in toluene. In the same way, the solubility of perylene and TCNQ in benzene at 25 °C is 5.4 and 0.7 mmol/L, respectively. At room temperature, the reaction quotient is larger than the solubility product constant for P3T1 (1.10 × 10−10 > 3.48 × 10−11) and the reaction quotient is smaller than the solubility product constant for P1T1 (3.78 × 10−6 < 3.98 × 10−6), which indicate that P1T1 will dissolve and P3T1 will precipitate in benzene.

Figure 3. All possible equilibria in the perylene−TCNQ system.

The Ksp values for P1T1 and P3T1 in toluene and benzene at various temperatures calculated from the solubility data according to eqs 8 and 9 are provided in Tables 3 and 4. These data show that the Ksp values for P1T1 in toluene are smaller than those in benzene for the temperature range 40−80

Table 3. Solubility Product Constant of P1T1 in Toluene and Benzene at Different Temperatures 40 °C toluene benzene

50 °C −7

1.12 × 10 (±0.11 × 10−7) 6.46 × 10−6 (±0.25 × 10−6)

60 °C −7

70 °C −7

1.73 × 10 (±0.11 × 10−7) 6.68 × 10−6 (±0.25 × 10−6)

1.87 × 10 (±0.11 × 10−7) 9.14 × 10−6 (±0.25 × 10−6) D

80 °C −7

2.12 × 10 (±0.11 × 10−7) 1.08 × 10−5 (±0.25 × 10−6)

2.54 × 10−7 (±0.11 × 10−7) 2.16 × 10−5 (±0.25 × 10−6)

dx.doi.org/10.1021/cg501206f | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Table 4. Solubility Product Constant of P3T1 in Toluene and Benzene at Different Temperatures toluene benzene

40 °C

50 °C

60 °C

70 °C

80 °C

2.08 × 10−9 (±0.17 × 10−9) 1.08 × 10−10 (±0.13 × 10−10)

2.74 × 10−9 (±0.17 × 10−9) 2.07 × 10−10 (±0.13 × 10−10)

5.31 × 10−9 (±0.17 × 10−9) 2.64 × 10−10 (±0.13 × 10−10)

1.48 × 10−8 (±0.17 × 10−9) 5.27 × 10−10 (±0.13 × 10−10)

2.79 × 10−8 (±0.17 × 10−9) 1.32 × 10−9 (±0.13 × 10−10)

Figure 4. Relationship between ln Ksp and 1/T (the black lines represent the fitting curves).

Figure 6. Solubility data of perylene and TCNQ in benzene and toluene.

Table 5. Thermodynamic Data for P3T1 and P1T1 in Toluene and Benzene P1T1 ΔH (kJ/mol) ΔS (J/mol)

The higher solubility of TCNQ in toluene than in benzene causes the intensive steady-state absorption and fluorescence observed in the perylene−TCNQ solution in benzene, as discussed below. The steady-state absorption and fluorescence spectra of perylene in benzene and toluene are identical (Figure 7A) because the perylene solubility in the two solvents is equal. The perylene fluorescence spectra in both solvents are in good agreement with the reported perylene monomer emission when the reabsorption effect is taken into consideration.53 The optical absorption of TCNQ in toluene, however, is over 10 times stronger than that in benzene due to the higher concentration. At room temperature, the fluorescence of TCNQ is not detected in either solvent. As a control experiment, we performed the spectroscopic characterization of stock solutions (perylene + TCNQ 1:1) in benzene and toluene. Figure 7B presents the steady-state spectra of both solutions. Figure 8 shows the fluorescence kinetics at λ = 472 nm. It can be observed that the perylene fluorescence lifetime in the stock solution (perylene + TCNQ 1:1) in benzene is 4.2 ns, a value that is a little longer than the perylene radiative lifetime reported.53 This 10% lifetime lengthening is due to reabsorption.54 However, the perylene fluorescence lifetime in toluene is 1.6 times shorter than that in benzene, which indicates the perylene fluorescence is quenched by the highly concentrated TCNQ in toluene (high solubility of TCNQ in toluene). It should be mentioned that the decay kinetics is independent of the emission wavelength both in benzene and in toluene. A similar situation is observed in steadystate fluorescence (Figure 7B), but the intensity difference is much larger. Because of the high solubility of TCNQ in toluene, the absorption edge extends to 550 nm; therefore, the intensity of the steady-state fluorescence from perylene in toluene is much (4.4 times) weaker than that in benzene.

P3T1

toluene

benzene

toluene

benzene

17.06 −77.75

26.23 −16.98

62.78 32.28

54.17 −18.47

Table 6. Solubility Product Constants of P3T1 and P1T1 in Toluene and Benzene at 25 °C P1T1 toluene Ksp

8.88 × 10

P3T1 benzene

−8

−6

3.98 × 10

toluene −10

4.83 × 10

benzene 3.48 × 10−11

Figure 5. Temperature dependence of ΔG to form P1T1 and P3T1 in toluene and benzene.

E

dx.doi.org/10.1021/cg501206f | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

stoichiometry of the starting materials: the crystals with lower solubility will precipitate first. As a result, only P1T1 is formed from toluene because P1T1 is less soluble than P3T1 in this solvent. As P3T1 is less soluble than P1T1 in benzene, only P3T1 was crystallized from benzene. Moreover, steady-state and timeresolved fluorescence measurements indicate that the high solubility of TCNQ in toluene results in weak fluorescence of perylene from the toluene solution due to formation of P1T1 that utilizes the perylene from the solution.



ASSOCIATED CONTENT

S Supporting Information *

CIF files and checkCif reports of P1T1 and P3T1 grown from solutions. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Authors

*(H.J.) E-mail: [email protected]. *(C.K.) E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was conducted by the NTU-HUJ-BGU Nanomaterials for Energy and Water Management Programme under the Campus for Research Excellence and Technological Enterprise (CREATE), which is supported by the National Research Foundation, Prime Minister’s Office, Singapore. K.P.G. thanks the NSF EAPSI for supporting her work at NTU through a fellowship. O.D.J. and L.E.M. acknowledge NSF support under grant DMR-1105147.

Figure 7. Steady-state absorption and fluorescence spectra: (A) Perylene and TCNQ stock solutions in benzene and toluene; (B) perylene + TCNQ 1:1 in benzene and toluene. (The spectrophotometers used give correct results for the absorption only at values of A < 2.5 due to sensitivity limitations.)



REFERENCES

(1) Reese, C.; Bao, Z. J. Mater. Chem. 2006, 16, 329−333. (2) Jiang, L.; Dong, H.; Hu, W. J. Mater. Chem. 2010, 20, 4994−5007. (3) El-Zaria, M. E. Spectrochim. Acta, Part A 2008, 69, 216−221. (4) Andrade, S. M.; Costa, S. M. B.; Pansu, R. J. Colloid Interface Sci. 2000, 226, 260−268. (5) Goetz, K. P.; Fonari, A.; Vermeulen, D.; Hu, P.; Jiang, H.; Diemer, P. J.; Ward, J. W.; Payne, M. E.; Day, C. S.; Kloc, C.; Coropceanu, V.; McNeil, L. E.; Jurchescu, O. D. Nat. Commun. 2014, 5, 5642. (6) Hasegawa, T.; Takeya, J. Sci. Technol. Adv. Mater. 2009, 10, 1−16. (7) Reese, C.; Bao, Z. N. Mater. Today 2007, 10, 20−27. (8) Iwasa, Y.; Asanuma, H.; Shimotani, H.; Takenobu, T.; Nishikawa, T.; Takeya, J. Mol. Cryst. Liq. Cryst. 2006, 455, 277−285. (9) Buurma, A. J. C.; Jurchescu, O. D.; Shokaryev, I.; Baas, J.; Meetsma, A.; de Wijs, G. A.; de Groot, R. A.; Palstra, T. T. M. J. Phys. Chem. C 2007, 111, 3486−3489. (10) Shokaryev, I.; Buurma, A. J. C.; Jurchescu, O. D.; Uijttewaal, M. A.; de Wijs, G. A.; Palstra, T. T. M.; de Groot, R. A. J. Phys. Chem. A 2008, 112, 2497−2502. (11) Kondo, H.; Moriya, T. J. Phys. Soc. Jpn. 1998, 67, 3695−3698. (12) Singleton, J. J. Solid State Chem. 2002, 168, 675−689. (13) Takahashi, K.; Horino, K.; Komura, T.; Murata, K. Bull. Chem. Soc. Jpn. 1993, 66, 733−738. (14) Avci, R.; Flynn, C. P. Phys. Rev. Lett. 1978, 41, 428−431. (15) Zhu, L. Y.; Yi, Y. P.; Li, Y.; Kim, E. G.; Coropceanu, V.; Bredas, J. L. J. Am. Chem. Soc. 2012, 134, 2340−2347. (16) Ferraris, J.; Cowan, D. O.; Walatka, V.; Perlstein, J. H. J. Am. Chem. Soc. 1973, 95, 948−949. (17) Singleton, J.; Mielke, C. Contemp. Phys. 2002, 43, 63−96. (18) Goetz, K. P.; Vermeulen, D.; Payne, M. E.; Kloc, C.; McNeil, L. E.; Jurchescu, O. D. J. Mater. Chem. C 2014, 2, 3065−3076.

Figure 8. Fluorescence decay kinetics of perylene + TCNQ 1:1 in benzene (blue) and toluene (red) at an emission wavelength of 472 nm.



CONCLUSIONS In summary, the single-crystal charge transfer compounds (perylene) 1 (TCNQ) 1 (P1T1) and (perylene) 3 (TCNQ) 1 (P3T1) were grown from solvents via slow cooling. P1T1 and P3T1 exhibit monoclinic and triclinic symmetries, respectively. Solution-grown single crystals of P1T1 are needle-like, and P3T1 crystals are cube-like. We found that the stoichiometry of the binary crystal is influenced by the type of solvent and not by the F

dx.doi.org/10.1021/cg501206f | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

(19) Shaheen, S. E.; Ginley, D. S.; Jabbour, G. E. MRS Bull. 2005, 30, 10−19. (20) Hoppe, H.; Sariciftci, N. S. J. Mater. Res. 2004, 19, 1924−1945. (21) Clarke, T. M.; Durrant, J. R. Chem. Rev. 2010, 110, 6736−6767. (22) Wakahara, T.; D’Angelo, P.; Miyazawa, K.; Nemoto, Y.; Ito, O.; Tanigaki, N.; Bradley, D. D. C.; Anthopoulos, T. D. J. Am. Chem. Soc. 2012, 134, 7204−7206. (23) Hasegawa, T.; Akutagawa, T.; Nakamura, T.; Sasou, Y.; Kondo, R.; Kagoshima, S.; Mochida, T.; Saito, G.; Iwasa, Y. Synth. Met. 1999, 103, 1804−1805. (24) Odom, S. A.; Caruso, M. M.; Finke, A. D.; Prokup, A. M.; Ritchey, J. A.; Leonard, J. H.; White, S. R.; Sottos, N. R.; Moore, J. S. Adv. Funct. Mater. 2010, 20, 1721−1727. (25) Mochida, T.; Moriyama, H.; Hasegawa, T.; Kondo, R.; Kagoshima, S.; Iwasa, Y. Synth. Met. 1999, 102, 1678−1679. (26) Nehm, S. J.; Rodríguez-Spong, B.; Rodríguez-Hornedo, N. Cryst. Growth Des. 2005, 6, 592−600. (27) ter Horst, J. H.; Cains, P. W. Cryst. Growth Des. 2008, 8, 2537− 2542. (28) Chiarella, R. A.; Davey, R. J.; Peterson, M. L. Cryst. Growth Des. 2007, 7, 1223−1226. (29) Jayasankar, A.; Reddy, L. S.; Bethune, S. J.; Rodríguez-Hornedo, N. Cryst. Growth Des. 2009, 9, 889−897. (30) Alhalaweh, A.; George, S.; Boström, D.; Velaga, S. P. Cryst. Growth Des. 2010, 10, 4847−4855. (31) Dobrowolski, M. A.; Garbarino, G.; Mezouar, M.; Ciesielski, A.; Cyranski, M. K. CrystEngComm 2014, 16, 415−429. (32) van Mullekom, H. A. M.; Vekemans, J. A. J. M.; Havinga, E. E.; Meijer, E. W. Mater. Sci. Eng., R 2001, 32, 1−40. (33) Trask, A. V.; van de Streek, J.; Motherwell, W. D. S.; Jones, W. Cryst. Growth Des. 2005, 5, 2233−2241. (34) Karki, S.; Frišcǐ ć, T.; Jones, W. CrystEngComm 2009, 11, 470− 481. (35) Halasz, I.; Puškarić, A.; Kimber, S. A. J.; Beldon, P. J.; Belenguer, A. M.; Adams, F.; Honkimäki, V.; Dinnebier, R. E.; Patel, B.; Jones, W.; Štrukil, V.; Frišcǐ ć, T. Angew. Chem., Int. Ed. 2013, 52, 11538−11541. (36) Vermeulen, D.; Zhu, L. Y.; Goetz, K. P.; Hu, P.; Jiang, H.; Day, C. S.; Jurchescu, O. D.; Coropceanu, V.; Kloc, C.; McNeil, L. E. J. Phys. Chem. C 2014, 118, 24688−24696. (37) Tickle, I. J.; Prout, C. K. J. Chem. Soc., Perkin Trans. 2 1973, 720− 723. (38) Ida, T.; Yakushi, K.; Kuroda, H. J. Chem. Phys. 1989, 91, 3450− 3455. (39) Ishii, K.; Yakushi, K.; Kuroda, H.; Inokuchi, H. Bull. Chem. Soc. Jpn. 1984, 57, 3043−3047. (40) Truong, K. D.; Bandrauk, A. D. Chem. Phys. Lett. 1976, 44, 232− 235. (41) Jiang, H.; Zhang, K. K.; Ye, J.; Wei, F.; Hu, P.; Guo, J.; Liang, C.; Chen, X.; Zhao, Y.; McNeil, L. E.; Hu, W.; Kloc, C. Small 2013, 9, 962− 962. (42) Shchuka, M. I.; Motyka, A. L.; Topp, M. R. Chem. Phys. Lett. 1989, 164, 87−95. (43) Compton, R. N.; Cooper, C. D. J. Chem. Phys. 1977, 66, 4325− 4329. (44) Anderson, P. W.; Lee, P. A.; Saitoh, M. Solid State Commun. 1973, 13, 595−598. (45) Seo, H.; Hotta, C.; Fukuyama, H. Chem. Rev. 2004, 104, 5005− 5036. (46) Jiang, H.; Kloc, C. MRS Bull. 2013, 38, 28−33. (47) Jurchescu, O. D.; Baas, J.; Palstra, T. T. M. Appl. Phys. Lett. 2004, 84, 3061−3063. (48) Ahn, S. Y.; Kang, S. J. Am. Ceram. Soc. 2000, 83, 1489−1494. (49) Euler, W. B.; Kirschenbaum, L. J.; Ruekberg, B. J. Chem. Educ. 2000, 77, 1039−1040. (50) Good, D. J.; Rodríguez-Hornedo, N. Cryst. Growth Des. 2009, 9, 2252−2264. (51) Sharpe, A. G.; Housecroft, C. E. Inorganic Chemistry, 2nd ed.; Prentice Hall: Harlow, England, 2005.

(52) Tro, N. J. Principles of Chemistry: A Molecular Approach;; Prentice Hall: Upper Saddle River, NJ, 2010. (53) Katoh, R.; Sinha, S.; Murata, S.; Tachiya, M. J. Photochem. Photobiol., A 2001, 145, 23−34. (54) Kaplanová, M.; Č ermák, K. J. Photochem. 1981, 15, 313−319.

G

dx.doi.org/10.1021/cg501206f | Cryst. Growth Des. XXXX, XXX, XXX−XXX