4,5-, 3,6-, and 3,4,5,6-tert-butylsulfanylphthalonitriles: synthesis and comparative structural and...

9
ORIGINAL RESEARCH 4,5-, 3,6-, and 3,4,5,6-tert-Butylsulfanylphthalonitriles: synthesis and comparative structural and spectroscopic analyses Ufuk Kumru Fabienne Dumoulin Erwann Jeanneau Fatma Yuksel Yari Cabezas Yunus Zorlu Vefa Ahsen Received: 22 June 2011 / Accepted: 21 July 2011 / Published online: 17 August 2011 Ó Springer Science+Business Media, LLC 2011 Abstract Three tert-butylsulfanylphthalonitriles have been prepared with optimized synthetic procedures. Their comparative structural analyses have been completed, with a focus on IR and NMR spectroscopy and refined X-ray structural data. Miscellaneous parameters such as UV absorption, melting points, and related polarity of the compounds are summarized. Keywords Phthalonitrile tert-Butanethiol tert-Butylsulfanylphthalonitrile Single-crystal IR NMR Structural analysis Introduction Phthalonitriles are among the most widely used precursors for phthalocyanine synthesis [1]. Despite their use as high- tech materials in the recent decades [2, 3], phthalocyanine derivatives still suffer from aggregation which prevent wider general uses. Such aggregation properties, due to the large aromatic core including eight nitrogen atoms, are a limitation for their use in several applications as it affects their desired properties and in particular UV–vis spectra. Aggregation could be suppressed to some extend by several structural factors. The nature of the metal (possibly axially substituted) is one of them [4]. Another important and widely used possibility is to tune the nature, number, and position of the substituents, an easily properties-modulat- ing parameter: the nature, number, and position of sub- stituents can strongly influence the aggregation and modulate a phthalocyanine’s optical, photophysical, and photochemical properties [413]. Substitution by bulky substituents is indeed a general strategy to limit one of the main phthalocyanines’ problems. In the particular case of N-bridged phthalocyanines, it proved to affect the central iron atoms electronic state [14]. The presence of tert-butyl substituents on phthalocyanines is therefore common, and tert-butylphthalonitrile is even commercially available. Nevertheless it has the disadvantage to lead to mixtures of positional isomers of phthalocyanines, complicating their characterization and preventing deep structure-related activities effects studies. This drawback is prevented by using isomerically pure octasubstituted (peripherally or nonperipherally) or persubstituted phthalocyanines bearing bulky substituents. The synthetic accessibility of the pre- cursors is then essential. Alkylthiosubstitution, on a synthetic point of view, offers the advantage of an easy synthesis of the precursors: the condensation of alkylthiols to dichlorophthalonitrile is a very well-known method [15, 16], when the more recently described condensation performed on 3,6-bis(4 0 -methyl- phenylsulfonyloxy)phthalonitrile proved to be effective with several alkanethiols [17]. Alkylthiosubstitution offers in addition the advantage to shift the electronic absorption properties towards the near-infrared region of the spectrum [18], a strongly desired effect for several applications [1921]. This effect is reinforced by non-peripheral U. Kumru F. Dumoulin F. Yuksel Y. Cabezas Y. Zorlu V. Ahsen (&) Department of Chemistry, Gebze Institute of Technology, P. O. Box 141, 41400 Gebze, Kocaeli, Turkey e-mail: [email protected] E. Jeanneau CNRS UMR 5615, Laboratoire des Multimate ´riaux et Interfaces, Universite ´ Lyon1, 69622 Villeurbanne, France Y. Cabezas Ecole Nationale Supe ´rieure de Chimie de Clermont-Ferrand, Clermont Universite ´, BP 10448, 63000 Clermont-Ferrand, France 123 Struct Chem (2012) 23:175–183 DOI 10.1007/s11224-011-9850-8

Upload: independent

Post on 18-Nov-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

ORIGINAL RESEARCH

4,5-, 3,6-, and 3,4,5,6-tert-Butylsulfanylphthalonitriles: synthesisand comparative structural and spectroscopic analyses

Ufuk Kumru • Fabienne Dumoulin •

Erwann Jeanneau • Fatma Yuksel •

Yari Cabezas • Yunus Zorlu • Vefa Ahsen

Received: 22 June 2011 / Accepted: 21 July 2011 / Published online: 17 August 2011

� Springer Science+Business Media, LLC 2011

Abstract Three tert-butylsulfanylphthalonitriles have

been prepared with optimized synthetic procedures. Their

comparative structural analyses have been completed, with

a focus on IR and NMR spectroscopy and refined X-ray

structural data. Miscellaneous parameters such as UV

absorption, melting points, and related polarity of the

compounds are summarized.

Keywords Phthalonitrile � tert-Butanethiol �tert-Butylsulfanylphthalonitrile � Single-crystal �IR � NMR � Structural analysis

Introduction

Phthalonitriles are among the most widely used precursors

for phthalocyanine synthesis [1]. Despite their use as high-

tech materials in the recent decades [2, 3], phthalocyanine

derivatives still suffer from aggregation which prevent

wider general uses. Such aggregation properties, due to the

large aromatic core including eight nitrogen atoms, are a

limitation for their use in several applications as it affects

their desired properties and in particular UV–vis spectra.

Aggregation could be suppressed to some extend by several

structural factors. The nature of the metal (possibly axially

substituted) is one of them [4]. Another important and

widely used possibility is to tune the nature, number, and

position of the substituents, an easily properties-modulat-

ing parameter: the nature, number, and position of sub-

stituents can strongly influence the aggregation and

modulate a phthalocyanine’s optical, photophysical, and

photochemical properties [4–13]. Substitution by bulky

substituents is indeed a general strategy to limit one of the

main phthalocyanines’ problems. In the particular case of

N-bridged phthalocyanines, it proved to affect the central

iron atoms electronic state [14]. The presence of tert-butyl

substituents on phthalocyanines is therefore common, and

tert-butylphthalonitrile is even commercially available.

Nevertheless it has the disadvantage to lead to mixtures of

positional isomers of phthalocyanines, complicating their

characterization and preventing deep structure-related

activities effects studies. This drawback is prevented by

using isomerically pure octasubstituted (peripherally or

nonperipherally) or persubstituted phthalocyanines bearing

bulky substituents. The synthetic accessibility of the pre-

cursors is then essential.

Alkylthiosubstitution, on a synthetic point of view, offers

the advantage of an easy synthesis of the precursors: the

condensation of alkylthiols to dichlorophthalonitrile is a

very well-known method [15, 16], when the more recently

described condensation performed on 3,6-bis(40-methyl-

phenylsulfonyloxy)phthalonitrile proved to be effective

with several alkanethiols [17]. Alkylthiosubstitution offers

in addition the advantage to shift the electronic absorption

properties towards the near-infrared region of the spectrum

[18], a strongly desired effect for several applications

[19–21]. This effect is reinforced by non-peripheral

U. Kumru � F. Dumoulin � F. Yuksel � Y. Cabezas � Y. Zorlu �V. Ahsen (&)

Department of Chemistry, Gebze Institute of Technology,

P. O. Box 141, 41400 Gebze, Kocaeli, Turkey

e-mail: [email protected]

E. Jeanneau

CNRS UMR 5615, Laboratoire des Multimateriaux et Interfaces,

Universite Lyon1, 69622 Villeurbanne, France

Y. Cabezas

Ecole Nationale Superieure de Chimie de Clermont-Ferrand,

Clermont Universite, BP 10448, 63000 Clermont-Ferrand,

France

123

Struct Chem (2012) 23:175–183

DOI 10.1007/s11224-011-9850-8

substitution, which is known as well to limit aggregation

compared to analogous peripheral substitution [4, 7]. This

effect was highlighted by concomitantly published first

reports by Cook’s [22] and Nyokong [23] of octa

non-peripherally alkylsulfanyl substituted phthalocyanines,

followed by more derivatives [24]. Isomerically pure

phthalocyanines substituted by tert-butylsulfanyl moieties

have been reported in the recent years by Zimcik [25–27]

prepared from 4,5-disubstituted phthalonitrile 1 (Fig. 1). To

enlarge the scope of possible precursors, we prepared the

whole set of tert-butylsulfanylsubstituted phthalonitriles

(1, 2, and 3) represented in Fig. 1. The synthesis of phthalo-

nitriles 1 [25, 26] and 3 [28] has been previously reported.

We optimized the synthesis and present in this paper the

comparative analytical data of the three phthalonitriles.

Experimental

Materials and methods

tert-Butanethiol 4 and tetrafluorophthalonitrile (7) were

purchased from Aldrich and used as received. 4,5-Dichlor-

ophthalonitrile (5) [15] and 3,6-bis(40-methylphenylsulfo-

nyloxy) phthalonitrile (6) [17] were prepared following

described procedures.

Synthesis of 4,5-bis(tert-butylsulfanyl)phthalonitrile (1)

tert-Butanethiol (13.3 g, 147 mmol), 4,5-dichlorophthalo-

nitrile (3) (9.7 g, 49 mmol), and anhydrous potassium

carbonate (80 g, 590 mmol) were stirred in N,N0-dimeth-

ylformamide (DMF) (100 mL) under a nitrogen atmo-

sphere overnight at room temperature. Water was then

added to the reaction mixture and the stirring continued 2 h

more. The resulting precipitate was then filtered, thor-

oughly washed with water, and crystallized from ethanol.

Yield: 12.8 g (86%). C16H20N2S2, MW 304.47. White

powder, mp: 163–165 �C. 1H NMR (CDCl3): d, ppm 7.86

(2H, s, Ar–CH), 1.46 (18H, s, CH3). 13C NMR (CDCl3): d,

ppm 146.96 (Ar–CH), 136.99 (Ar–C-S), 115.19 (Ar–C–

CN), 112.56 (CN), 49.46 (C–S), 31.04 (CH3). ATR-IR:

mmax, cm-1 3068, 2978, 2964, 2939, 2899, 2867, 2228

(CN), 1505, 1451, 1396, 1366, 1283, 1221, 1158 (Ar–S–C),

1105, 931, 873 (C–H). LC–MS (ESI) m/z: calcd for

C16H20N2NaS2: 327.099; found 327.107.

Synthesis of 3,6-bis(tert-butylsulfanyl)phthalonitrile (2)

tert-Butanethiol (5.7 g, 64 mmol) was dissolved in DMF

(50 mL) under a nitrogen atmosphere and 2,3-dicyano-

1,4-phenylene bis(4-methylbenzenesulfonate) (6) (10 g,

21 mmol) and anhydrous potassium carbonate (14.6 g,

106 mmol) were added. The mixture was stirred under a

nitrogen atmosphere during 3 days at room temperature.

Water was added and the resulting precipitate was filtered

then dissolved in dichloromethane. The aqueous filtrate

was extracted by dichloromethane (3 9 50 mL). The

organic extracts were further treated with 5% sodium car-

bonate solution (2 9 250 mL). The organic phase was

dried on sodium sulfate, filtered, and evaporated. The

product was crystallized from ethanol. Yield: 2.2 g (34%).

C16H20N2S2, MW 304.47. Pale yellow powder, mp:

177–178 �C. 1H NMR (CDCl3): d, ppm 7.84 (2H, s,

Ar–CH), 1.41 (18H, s, CH3). 13C NMR (CDCl3): d, ppm

140.86 (Ar–CH), 139.75 (Ar–C–S), 124.94 (Ar–C–CN),

114.82 (CN), 50.55 (C–S), 31.07 (–CH3). ATR-IR: mmax,

cm-1 3071, 3054, 2979, 2962, 2941, 2923, 2897, 2868,

2228 (CN), 1553, 1534, 1456, 1441, 1387, 1365, 1171,

1158 (Ar–S–C), 865 (C–H). LC–MS (ESI) m/z: calcd for

C16H20N2NaS2: 327.099; found 327.090.

Synthesis of 3,4,5,6-tetra(tert-

butylsulfanyl)phthalonitrile (3)

tert-Butanethiol (4.7 g, 40 mmol) and 3,4,5,6-tetra-

fluorophthalonitrile (7) (1 g, 5 mmol) were dissolved in

DMF (80 mL) under a nitrogen atmosphere, then anhy-

drous potassium carbonate (27 g, 25 mmol) was added.

The mixture was stirred under a nitrogen atmosphere at

room temperature for 5 days and then stirred at 80 �C 5 h

further. Water was added and the resulting precipitate was

filtered then dissolved in dichloromethane. The aqueous

filtrate was extracted by dichloromethane (3 9 50 mL).

The organic extracts were further treated with 5% sodium

carbonate solution (2 9 250 mL), then water. The organic

phase was dried on sodium sulfate, filtered, and evaporated.

The product was crystallized from ethanol. Yield: 1.4 g

(60%). C24H36N2S4, MW 480.82, mp: 216–218 �C. 1H

NMR (CDCl3): d, ppm 1.25, 1.39 (s, CH3). 13C NMR

(CDCl3): d, ppm 148.01, 159.50 (Ar–CH), 126.42 (Ar–C–

CN), 115.72 (CN), 53.72, 54.01 (C–S), 31.56, 31.72 (CH3).

ATR-IR: mmax, cm-1 2959, 2921, 2899, 2862, 2230 (CN),

1487, 1470, 1455, 1363, 1305, 1218, 1147 (Ar–S–C), 1023,

928, 804 (C–H).

Fig. 1 Structure of phthalonitriles 1, 2, and 3

176 Struct Chem (2012) 23:175–183

123

X-ray data collection and structure refinement

For each compound, a suitable crystal was selected and

mounted on a Bruker AXS SMART-APEXII CCD dif-

fractometer using Mo radiation (k = 0.71073 A).

Indexing was performed using APEX2 [29] (Difference

Vectors method). Data integration and reduction were

carried out with SaintPlus 6.01. [30]. Absorption correction

was performed by multi-scan method implemented in

SADABS [31]. Space groups were determined using

XPREP implemented in APEX2 [29]. The structures of

compound 1 and 3 were solved by direct methods with

SIR97 [32] while the structure of compound 2 was solved

with the charge-flipping algorithm implemented in

SUPERFLIP [33]. The least-square refinement on F2 was

achieved with the CRYSTALS software [34].

All non-hydrogen atoms were refined anisotropically.

The hydrogen atoms were all located in a difference map,

but those attached to carbon atoms were repositioned

geometrically. The H atoms were initially refined with soft

restraints on the bond lengths and angles to regularize their

geometry (C–H in the range 0.93–0.98 A) and Uiso(H) (in

the range 1.2–1.5 times Ueq of the parent atom), after which

the positions were refined with riding constraints.

Further details may be found in Table 1.

Results and discussion

Synthesis

1 was prepared as described in literature for other thiols

[15, 16], by reacting tert-butanethiol (2-methylpropane-2-

thiol) (4) with 4,5-dichlorophthalonitrile (5) in the presence

of potassium carbonate in DMF at room temperature

overnight (Scheme 1). 1 was obtained in 86% yield,

Table 1 Crystal data and structure refinement for 1, 2, and 3

Crystal parameters 1 2 3

Empirical formula C16H20N2S2 C16H20N2S2 C24H36N2S4

Formula weight (g mol-1) 304.48 304.48 480.83

Temperature (K) 293(2) 293(2) 293(2)

Wavelength (A) 0.71073 0.71073 0.71073

Crystal system Triclinic Orthorhombic Monoclinic

Space group P - 1 Cmca P21/n

Unit-cell dimensions

a (A) 9.395(4) 7.931(1) 17.850(3)

b (A) 9.577(4) 8.782(1) 9.554(2)

c (A) 11.080(5) 24.080(2) 18.087(3)

a (�) 80.79(1) 90 90

b (�) 75.76(1) 90 117.939(5)

c (�) 63.31(1) 90 90

Crystal size (mm) 0.32 9 0.31 9 0.31 0.43 9 0.12 9 0.10 0.40 9 0.32 9 0.21

V (A3) 861.9(6) 1677.2(3) 2725.0(9)

Z 2 4 4

qcalcd (mg m-3) 1.173 1.206 1.172

l (mm-1) 0.30 0.31 0.36

F(000) 324 648 1032

h range for data collection (�) 2.8–30.4 1.7–27.6 1.3–25.0

h/k/l -10,12/-11,12/-14,14 -10,10/-11,11/-31,28 -21,21/-11,11/-21,21

Reflections collected 11393 13327 30558

Independent reflections 3906 1046 4812

Tmax and Tmin 0.91 and 0.91 0.964 and 0.970 0.89 and 0.93

Data/restraints/parameters 3905/181/0 1046/62/2 4812/271/0

Goodness-of-fit on F2 0.96 0.97 1.04

Final R indices [I [ 2r(I)] R1 = 0.047, wR2 = 0.128 R1 = 0.040, wR2 = 0.101 R1 = 0.042, wR2 = 0.113

Largest diff. peak and hole (e A-3) -0.27 and 0.41 -0.42 and 0.36 -0.34 and 0.81

Struct Chem (2012) 23:175–183 177

123

superior than the reported conditions using sodium hydride

in DMF in the presence of copper oxide, even after opti-

mization [16, 25, 26].

For the preparation of 3,6-dialkylsulfanylphthalonitriles,

two activated precursors are known to undergo nucleo-

philic substitution by thiolates: 1,2-dicyano-3,6-bis(tolu-

enesulfonyl)benzene (6) and the more active analogous

1,2-dicyano-3,6-bis(trifluorosulfonyl)benzene. In our case,

satisfying yields were obtained by stirring 4 and 6 in the

presence of potassium carbonate in DMF at room tem-

perature during 3 days (Schemes 2, 3).

The reaction of tetrachlorophthalonitrile with thiols has

been described in detail, with the successive substitution of

the chloride atoms by different thiols and phenols, includ-

ing 4 [35, 36]. The analogous 3,4,5,6-tetra(octylsulfanyl)-

phthalonitrile was prepared using ionic liquids as the solvent

in quite good yields [37]. In our case, the substitution was

performed on the tetrafluorophthalonitrile 7, known to react

with nucleophiles [38]. The reaction was completed in DMF

using potassium carbonate as the base.

We manage to improve the yields of compounds com-

pared to their previously described synthesis (Table 2).

Nevertheless it is still desirable to increase it, especially in

the case of 2. Its reproducible yield, lower compared to the

tetrasubstituted derivative 3, is probably due to a lower

reactivity of the corresponding tosylated precursor 6

compared to the fluorinated derivative 7. It will be useful in

the future to use a more active 3,6-disubstituted phthalo-

nitrile, such as for example the corresponding trifluoro-

sulfinate previously used for similar reactions.

To easily check and compare the melting point and

polarity (quantified here by the Rf in a same eluent) of each

phthalonitriles, these data are summarized in Table 3.

Single-crystal structural analysis

The structures and geometries of compounds 1, 2, and 3 in

the solid state were established by single-crystal X-ray

structural analysis. ORTEP representation of these three

structures is shown in Figs. 2, 3, and 4, respectively. The

inter-molecular interactions for the three compounds are

gathered in Table 4.

In compound 1, one of the butylsulfanyl groups is nearly

situated in the same plane as the phenyl ring with an angle

of 16.62(7)� between the mean plane of the phenyl ring and

the plane that goes through atoms C5, S1, and C9. On the

other hand, the second butylsulfanyl group is nearly

Scheme 1 Preparation of the 4,5-bis-tert-butylsulfanylphthalonitrile 1

Scheme 2 Preparation of the 3,6-bis-tert-butylsulfanylphthalonitrile 2

Scheme 3 Preparation of the 3,4,5,6-tetra-tert-butylsulfanylphthalo-

nitrile 3

Table 2 Optimized synthetic conditions and yields for 1, 2, and 3

Compound Previous works Our results

Yield (%) Conditions base/solvent equiv.

phthalonitrile/4/base

Ref. Yield (%) Conditions base/solvent, equiv.

phthalonitrile/4/base

1 65 NaH/Cu2O/DMF 1/2/2.5 [26] 86 K2CO3/DMF, 1/3/12

2 Not reported 34 K2CO3/DMF, 1/3/5

3 55 1/8 [28] 60 K2CO3/DMF, 1/8/40

Table 3 Melting point and polarity of the three phthalonitriles

Phthalonitrile 1 2 3

mp (�C) 163–165 177–178 216–218

Rf (hexane/ethyl acetate 4:1) 0.73 0.53 0.80

178 Struct Chem (2012) 23:175–183

123

perpendicular to the phenyl ring with an angle of 80.90(7)�between the plane formed by C6, S2, and C13 and the

plane of the phenyl ring.

The molecules form dimeric units through a double

C–H���N interaction. These interactions are completed by

p���p intermolecular contacts which lead to the formation

of chains running along the a axis of the unit-cell. Finally

those chains interact through C–N���p contacts to form

layers perpendicular to the c axis of the unit-cell (Fig. 5).

In compound 2, both butylsulfanyl groups are perpen-

dicular to the phenyl ring with an angle of 90.00(4)�between the plane formed by C2, S1, and C6 and the plane

of the phenyl ring.

The molecules form infinite chains along the a axis of

the unit-cell through two C–H���N interactions, as shown in

Fig. 6. These interactions are completed by double C–N���pand p���p intermolecular contacts which lead to the for-

mation of two-dimensional sheets in the (ab) plane of the

unit-cell.

The butylsulfanyl groups in compound 3 are nearly

perpendicular to the phenyl ring with angles ranging from

79.58(8)� to 86.6(1)� and the groups are disposed alterna-

tively over and under the phenyl ring. The steric hindrance

of the four butylsulfanyl groups leads to puckering of the

phenyl ring with the following Cremer and Pople param-

eters [39]: Q = 0.104(2) A, h = 69(1)� and u = 291(1)�.

Two molecules form a dimer through a double C–N���pinteraction and a dipole–dipole interaction between anti-

parallel cyano groups. The observed C–H���N and C–H���Cinteractions lead to a two-dimensional framework (Fig. 7).

ATR-IR

The ATR-IR spectrum of 1, 2, and 3 are presented in

Fig. 8. They are quite similar, only the relative vibration

peak of the C–H bonds in the fingerprint area, much more

intense for 3 than for 1 and 2 can be noticed, in accordance

with the molecular structure. The CN peaks are sharp and

centered at 2230 cm-1.

UV absorption

Each phthalonitriles has a maximum absorption at 275 nm,

with similar extinction coefficients (log e at 5.4).

1H and 13C NMR

The symmetric structures of the phthalonitriles lead to

simple NMR spectra (Figs. 9, 10).

A strong effect of the tetrasubstitution compared to the

disubstitution pattern is observed for the protons and car-

bons of the tert-butyl groups.

Fig. 2 Molecular structure of 1 (ORTEP, 30% probability ellipsoids)

Fig. 3 Molecular structure of 2 (ORTEP, 30% probability ellipsoids).

Symmetry codes: (i) 2 - x, y, z; (ii) x, 1 - y, 1 – z; (iii) 2 - x, 1 - y,

1 - z

Fig. 4 Molecular structure of 3 (ORTEP, 30% probability ellipsoids)

Struct Chem (2012) 23:175–183 179

123

Methyl protons of the disubstituted phthalonitriles 1 and

2 resonate nearly at the same frequency, when those of the

tetrasubstituted phthalonitrile 3 resonate separately into

two singlets separated by 0.15 ppm (Fig. 9). Similarly,

when the methyl carbons of the disubstituted phthalonitr-

iles 1 and 2 have very close chemical shifts, the tetrasub-

stitution pattern induces a differentiation of 0.15 ppm. The

frequencies of the quaternary carbons are shifted to lower

fields by nearly 4 ppm compared to the disubstituted

derivatives (Fig. 10). The resonance of the nitrile carbon is

less affected by the substitution pattern.

Table 4 Intermolecular interactions of the three phthalonitriles

Compound p���p interactions between phenyl rings

Cg Alpha (�) Beta (�) Cg–Cg (A)

1 C3–C4–C5–C6–C7–C8 0 19.18 3.898(2)

2 C2–C3–C3ii–C2ii–C3iii–C3i 0 54.25 5.917(1)

C–N���p interactions

Y–X���Cg N���Cg (A) C–N���Cg (�) C���Cg (A)

1 C1–N1���(C3–C4–C5–C6–C7–C8) 3.656(4) 110.5(2) 4.194(3)

2 C4–N5���(C2–C3–C3ii–C2ii–C3iii–C3i) 3.492(4) 103.6(3) 3.922(4)

3 C4–N5���(C2–C3–C6–C7–C13–C14) 3.569(2) 121.7(2) 4.282(3)

C–H���X interactions

C–H���X H���N (A) C–H���N (�) C���N (A)

1 C7–H2���N2i 2.741(2) 164.4(1) 3.658(3)

2 C3–H31���N5ii 2.449(3) 169.5(2) 3.363(4)

3 C29–H291���N26iii 2.715(2) 154.1(2) 3.603(3)

C28–H283���N26iii 2.698(2) 152.5(2) 3.596(3)

C17–H173���C25iv 2.779(2) 166.3(2) 3.728(4)

Alpha: dihedral angle between the planes; Beta: angle between the planes; Cg–Cg: distance between ring centroids

Symmetry codes: (i) 1 - x, -y, 2 – z; (ii) 1 - x, y, z; (iii) � ? x, 1/2 - y, 1/2 ? z; (iv) x, 1 ? y, z

Fig. 5 View of one of the sheet of compound 1 showing the inter-

molecular interactions as dotted lines (blue: C–N���p, green: p���p,

red: C–H���N) (Color figure online)

Fig. 6 View of one of the sheet of compound 2 showing the inter-

molecular interactions as dotted lines (blue: C–N���p, green: p���p,

red: C–H���N) (Color figure online)

180 Struct Chem (2012) 23:175–183

123

Conclusion

We reported here the optimized syntheses of three pht-

halonitriles likely to play an increased role in the chemistry

of phthalocyanine as they bear bulky substituents, able to

lower the aggregation of the corresponding phthalocya-

nines without leading to isomeric mixtures. Their complete

analysis gathered key parameters that have been compared

and interpreted relatively to the structural variations. The

synthesis of the corresponding phthalocyanines is in

progress.

Supplementary materials

CCDC-798223, 829161, and 829162 contain the supple-

mentary crystallographic data for this paper. These data

can be obtained free of charge at www.ccdc.cam.ac.uk/

conts/retrieving.html [or from the Cambridge Crystallo-

graphic Data Centre (CCDC), 12 Union Road, Cambridge

Fig. 7 View of one of the sheet of compound 3 showing the inter-

molecular interactions as dotted lines (blue: C–N���p, light blue:

cyano–cyano interactions, red: C–H���N, black: C–H���C) (Color

figure online)

Fig. 8 ATR-IR spectrum

of 1, 2, and 3

Struct Chem (2012) 23:175–183 181

123

CB2 1EZ, UK; fax: ?44(0)1223-336033; email: deposit@

ccdc.cam.ac.uk].

Acknowledgments The Scientific and Technological Research

Council of Turkey TUBITAK (PIA Turkey-France bilateral project

Bosphorus 109M356) is gratefully acknowledged. YC thanks the

long-life learning Erasmus funding.

References

1. Sharman WM, Lier JE (2003) In: Kadish KM, Smith KM,

Guilard R (eds) The porphyrin handbook, vol 15. Academic

Press, New York, pp 1–62

2. Claessens CG, Hahn U, Torres T (2008) Chem Rec 8:75

3. de la Torre G, Claessens CG, Torres T (2007) Chem Commun 2000

Fig. 9 1H NMR spectrum of 1, 2, and 3

Fig. 10 13C NMR spectrum of 1, 2, and 3

182 Struct Chem (2012) 23:175–183

123

4. Snow AW (2003) In: Kadish KM, Smith KM, Guilard R (eds)

The porphyrin handbook, vol 17, Academic Press: New York,

pp 129–176

5. Rio Y, Rodriguez-Morgade MS, Torres T (2008) Org Biomol

Chem 6:1877

6. Luk’yanets EA, Nemykin VN (2010) J Porphyr Phthalocyanines

14:1

7. Zorlu Y, Un I, Dumoulin F (2009) J Porphyr Phthalocyanines

13:760

8. Nemykin VN, Luk’yanets EA (2010) ARKIVOC (i)136–208

9. Ogunsipe A, Chen JY, Nyokong T (2004) New J Chem 28:822

10. Kobayashi N, Ogata H, Nonaka N, Luk’yanets EA (2003) Chem

Eur J 9:5123

11. Choi MTM, Li PPS, Ng DKP (2000) Tetrahedron 56:3881

12. Zorlu Y, Dumoulin F, Durmus M, Ahsen V (2010) Tetrahedron

66:3248

13. Dumoulin F, Durmus M, Ahsen V, Nyokong N (2010) Coord

Chem Rev 254:2792

14. Isci U, Dumoulin F, Ahsen V, Sorokin AB (2010) J Porphyr

Phthalocyanines 14:324

15. Wohrle D, Eskes M, Shigehara K, Yamada A (1993) Synthesis

194

16. Gurek AG, Bekaroglu O (1994) J Chem Soc Dalton Trans 1419

17. Mbambisa G, Tau P, Antunes E, Nyokong T (2007) Polyhedron

26:5355

18. Gregory P (2000) J Porphyr Phthalocyanines 4:432

19. Wainwright M (2010) Color Technol 126:115

20. Qian G, Wang ZY (2010) Chem Asian J 5:1006

21. Wang ZY, Zhang J, Wu X, Birau M, Yu G, Yu H, Qi Y,

Desjardins P, Meng X, Gao JP, Todd E, Song N, Bai Y, Beaudin

AMR, LeClair G (2004) Pure Appl Chem 76:1435

22. Burnham PM, Cook MJ, Gerrard LA, Heeney MJ, Hughes DL

(2003) Chem Commun 2064

23. Ogunbayo TB, Nyokong T (2009) Polyhedron 28:2710

24. Sakamoto K, Ohno-Okumura E, Kato T, Sogaa H (2010) J Por-

phyr Phthalocyanines 14:47

25. Kostka M, Zimcik P, Miletin M, Klemera P, Kopecky K, Musil Z

(2006) J Photochem Photobiol A 178:16

26. Zimcik P, Miletin M, Novakova V, Kopecky K, Nejedla M, Stara

V, Sedlackova K (2009) Aust J Chem 62:425

27. Novakova V, Zimcik P, Miletin M, Kopecky K, Ivincova J (2010)

Tetrahedron Lett 51:1016

28. Volkov KA, Avramenko GV, Negrimovskii VM, Luk’yanets EA

(2007) Russ J Gen Chem 77:1108

29. Bruker (2008) APEX2 (Version 2008.1-0). Bruker AXS Inc.,

Madison

30. Bruker (2001). SAINT-V6.28A. Data Reduction Software

31. Sheldrick GM SADABS. Program for empirical absorption cor-

rection. University of Gottingen, Germany

32. Altomare A, Burla MC, Camalli M, Cascarano GL, Giacovazzo

C, Guagliardi A, Grazia A, Moliterni G, Polidori G, Spagna RJ

(1999) Appl Cryst 32:115

33. Palatinus L, Chapuis GJ (2007) Appl Cryst 40:786

34. Betteridge PW, Carruthers JR, Cooper RI, Prout K, Watkin DJJ

(2003) Appl Cryst 36:1487

35. Volkov KA, Avramenko GV, Negrimovskii VM, Luk’yanets EA

(2007) Russ J Gen Chem 77:1126

36. Volkov KA, Negrimovskii VM, Avramenko GV, Luk’yanets EA

(2008) Russ J Gen Chem 78:1794

37. Lo P-C, Cheng DYY, Ng DKP (2005) Synthesis 7:1141

38. Birchall JWI, Haszeldine RN, Morley JO (1970) J Chem Soc C 456

39. Cremer D, Pople JA (1975) J Am Chem Soc 97:1354

Struct Chem (2012) 23:175–183 183

123