• No results found

Platinum bisphosphine complexes of 1,8 naphthosultone

N/A
N/A
Protected

Academic year: 2020

Share "Platinum bisphosphine complexes of 1,8 naphthosultone"

Copied!
17
0
0

Loading.... (view fulltext now)

Full text

(1)

Platinum

bis

phosphine complexes of 1,8-naphthosultone

Louise M. Diamond, Fergus R. Knight, David B. Cordes, Amy L. Fuller, Alexandra M. Z.

Slawin and J. Derek Woollins*

School of Chemistry, University of St Andrews, St Andrews, Fife KY16 9ST

(2)

Abstract

A series of three platinum (II) bisphosphine complexes 1-3 [Pt(1-(SO2),8-(O)-nap)(PR3)2] (where R3 = Ph3, Ph2Me, Me2Ph) have been prepared by metathesis from cis-[Pt(PR3)2Cl2)] and the dilithium salt of 1,8-naphthosultone. The novel compounds were fully characterised by X-ray crystallography, multinuclear NMR, IR and MS. The molecular structures of 1-3 were compared by measuring the peri-distance, splay angle magnitude, peri-atom displacement, naphthalene ring torsions, aromatic ring orientations and the geometry around the platinum centre. The platinum metal adopts a distorted square planar geometry in all three complexes which causes deformation of the naphthalene system. The degree of molecular deformation does not decrease upon going from 1 to 3 as anticipated, instead a competition between steric effects and intramolecular interactions causes 3 to display distortion intermediate of 1 and 2.

Introduction

The naphthalene backbone has a unique geometry which allows the study of non-bonded intramolecular interactions.1 Heteroatoms that are substituted at the peri-positions (positions 1- and 8- of the naphthalene ring) are forced to occupy space that is closer than the sum of their van der Waals radii, resulting in unique interactions.2,3 Although steric strain occurs when the hydrogen atoms are replaced by larger heteroatoms, a great array of peri-substituted naphthalenes have been prepared.4 Preparation is possible because of the naphthalene unit's ability to relieve strain through distortion. Firstly, attractive interactions can occur which relieve the strain by causing formation of weak or strong bonds between the substituents.2,3,4 Secondly, repulsive interactions cause the peri-bonds to distort in-plane or out-of-plane or distort the naphthalene backbone away from its normal geometry.2,3,4 Furthermore, complexation to a bridging metal species will also relieve strain; with the naphthalene system providing the correct spacial arrangement for bidentate coordination.

In the late 1970’s and early 1980’s Teo and co-workers5 coordinated tetrathionaphthalene (TTN), tetrachlorotetrathionaphthalene (TCTTN) and tetrathiotetracene (TTT) to a Pt(PPh3)2 centre through oxidative addition reactions with [Pt(PPh3)4]. Due to the structural similarity of these compounds to naphthalene, we used this oxidative reaction to study the coordination chemistry of 1,8-dichalcogen naphthalenes and the oxidised derivatives of naphtho[1,8-cd]1,2-dithiole, to platinum bisphosphines.6 An alternative metathetical reaction was also utilised where the parent 1,8-dichalcogen naphthalene was reduced with lithium triethylborohydride to form the dilithio-1,8-dichalcogenato naphthalene. The dilithio-species was then reacted with cis-[Pt(PR3)2Cl2] (where R = Ph or Me) in THF at room temperature to yield [Pt(1,8-E2-nap)(PR3)2] (where E = S or Se].6 As a continuation of this work, we chose to study the coordination chemistry of the commercially available 1,8-naphthosultone, which up until now had not been investigated.

(3)

Results and Discussion

Compounds 1-3 were synthesised and their crystal structures were determined. The three complexes were characterised by multinuclear NMR, IR spectroscopy, mass spectrometry and the homogeneity of the complexes was confirmed by microanalysis.

Utilising our previously reported metathetical method,6 the sulfur-oxygen bond in 1,8-naphthosultone was reduced with two equivalents of lithium triethylborohydride to form a yellow solution of the dilithio-species. Subsequent addition of the dilithio-species to a suspension of the appropriate

cis-dichlorobis(phosphine)platinum in THF resulted in the formation of [Pt(1-(SO2),8-(O)-nap)(PPh3)2] (1), [Pt(1-(SO2),8-(O)-nap)(PPh2Me)2] (2)and [Pt(1-(SO2),8-(O)-nap)(PMe2Ph)2] (3).

Scheme 1 Reaction route for synthesis of [Pt(1-(SO2),8-(O)-nap)(PPh3)2] (1), [Pt(1-(SO2),8-(O)-nap)(PPh2Me)2] (2)and

[Pt(1-(SO2),8-(O)-nap)(PMe2Ph)2] (3).

The 31P{1H} NMR spectra of the three complexes (Table 1) all display similar AX patterns with appropriate platinum satellites. For example, the phosphorus resonances and 1J(31P-195Pt) coupling constants of 1 are δ(P

A) = 19.1 ppm (2504 Hz) and δ(PX) = 10.6 ppm (3876 Hz), which are assigned to the phosphine groups trans to the SO2 group and the oxygen atom, respectively. Complexes 2 and 3 are characterised accordingly (Table 1). In previously characterised [(PPh3)2Pt] complexes containing (-S(O)R) and -(SR) ligands, the phosphine groups

trans to the (-S(O)R) group have smaller 1J(31P-195Pt) coupling constants because of the larger trans influence of the (-S(O)R) moiety.6

The trans influence, also known as the structural trans effect, is defined as the ability of a ligand to weaken the bond trans to itself.7 The trans influence is a thermodynamic concept and must not be mistaken for the trans effect which is a kinetic concept. The trans effect, otherwise known as the kinetic trans effect, describes the ability of a ligand to alter the lability of a ligand trans to itself.7 If the ligand is a strong σ-donor ligand or π-acceptor ligand then it increases the rate of substitution of a ligand that lies trans to itself.8

(4)

group is a stronger π-acceptor than the oxygen atom resulting in the SO2 group having a larger trans influence than the oxygen atom and hence the smaller 1J(31P-195Pt) coupling constant value has been assigned to the phosphine group trans to the SO2 group.

Table 131P{1H} NMR data for complexes 1 to 3.

Product Chemical Shifts [ppm] Coupling Constants [Hz]

δ(31P

A) δ(31Px) 1J(PA-Pt) IJ(PX-Pt) 2J(PA-PX)

1 19.1 10.6 2504 3876 19

2 3.0 -2.4 2537 3700 19

3 -9.6 -17.6 2504 3603 21

It has been shown that higher 1J(31P-195Pt) coupling constants are observed for more electron withdrawing R groups, with 1J(31P-195Pt) decreasing as phenyl groups are replaced with electron donating methyl groups.8 It can be seen from Table 1 that there is a steady decrease in value for 1J(31P

X-195Pt) coupling constant as the phenyl groups are replaced by methyl groups, as expected. This effect is less apparent for the signals associated with the phosphorus trans to the SO2 group. The 2J(31PA- 31PX) values of 19-21 Hz are indicative of unsymmetrically substituted cis-platinum diphosphine complexes.6

It was also noted, when the products were left overnight in chlorinated solvent, the solution changed colour from yellow to green. 31P NMR of the green solution showed degradation of product, with the only peaks being those from the appropriate platinum chloro phosphine complex. Similar behaviour was also observed with the platinum complexes of naphthalene-1,8-dichalcogens and the oxidised derivatives of naphtho[1,8-cd ]-1,2-dithiole.6 The authors found that addition of small quantities of triethylamine to work-up and NMR solvents prevented decomposition; leading them to conclude that free hydrogen chloride in the solvent was causing decomposition. The same can be concluded in this instance.

X-ray Investigations

In order to analyse the impact on the peri-naphthalene system caused by forming platinum complexes, the crystal structure of 1,8-naphthosultone (L) was first studied.

The crystal structure of 1,8-naphthosultone shows a peri sulfur-oxygen bond distance of 1.6407(14) Å. This is shorter than the 2.44 Å9peri-distance in unsubstituted naphthalene; which is expected as there is a bond between the peri-substituents. However, it is longer than the typical sulfur-oxygen bond in a (C-O-SO2-C) system (1.577 Å)10 due to the rigid naphthalene C

(5)

of the bay angles is 333.31(12)°; less than those of naphthalene confirming a favourable interaction is occurring between the peri-atoms resulting in a peri-bond. Little distortion of the naphthalene carbon skeleton is observed with torsion angles deviating by 0.62° and 0.86° from the ‘ideal’ 180°. Minor out-of-plane distortion is also observed with the oxygen atom sitting 0.018(1) Å above the naphthyl plane. The sulfur atom, however, essentially lies on the naphthyl plane.

Figure 1 X-ray crystallography numbering scheme for L.

Table 2 Selected Bond Lengths [Å] and Angles [°] for ligand L.

Peri-region distance

E(1)-E(9) 1.6407(14)

Bond lengths

C(1)-S(1) 1.760(3)

C(9)-E(9) 1.402(4)

S(1)-O(1) 1.424(3)

S(1)-O(2) 1.4243(18)

C(1)-C(2) 1.372(2)

C(2)-C(3) 1.422(4)

C(3)-C(4) 1.379(3)

C(4)-C(5) 1.423(3)

C(5)-C(10) 1.405(4)

C(5)-C(6) 1.421(3)

C(6)-C(7) 1.373(3)

C(7)-C(8) 1.425(4)

C(8)-C(9) 1.358(3)

C(9)-C(10) 1.406(3)

C(1)-C(10) 1.390(3)

Peri-region bond angles

E(1)-C(1)-C(10) 106.63(15)

(6)

E(9)-C(9)-C(10) 111.78(17)

∑ of bay angles 333.31(22)

Splay anglea -26.69

Bond angles

C(4)-C(5)-C(6) 128.3(3)

E(9)-E(1)-C(1) 95.15(10)

E(1)-E(9)-C(9) 111.54(13)

O(1)-S(1)-O(2) 118.53(13)

O(1)-S(1)-E(9) 107.01(9)

O(2)-S(1)-E(9) 107.31(9)

C(1)-S(1)-O(1) 113.03(11)

C(1)-S(1)-O(2) 112.77(12)

Out of plane displacement

S(1) +0.001(1)

O(9) +0.018(1)

Central naphthalene ring torsion angles

C(6)-C(5)-C(10)-C(1) 179.14(16)

C(4)-C(5)-C(10)-C(9) -179.38(16)

aSplay angle: ∑ of the three bay region angles - 360.

Figure 2 The molecular structure of 1,8-naphthosultone(L)with H atoms omitted for clarity.

(7)

of 1, 2 and 3 are shown in Figures 4, 5 and 6, respectively. In all three structures, as expected from the NMR studies, the 1,8-naphthosultone acts as a bidentate ligand, coordinating to the platinum via the sulfur and the oxygen atom to form a six-membered chelate ring. The central platinum metal adopts a distorted square planar geometry in each case, with angles deviating significantly from the ideal (90°).

Figure 3 The angles (°) associated with the square planar geometry of the platinum metal in complexes 1, 2 and 3.

The Tolman cone angles for PPh3, PPh2Me and PPhMe2 are 145°, 136° and 122o, respectively.11 Therefore, one would anticipate a steady contraction in the P-Pt-P angle on going from 1 to 3 and this is observed.

In 1 the P(1)-Pt(1)-P(2) angle of 99.73(5)° is enlarged to greater than the ideal 90° in order to accommodate the bulky cis-triphenylphosphine groups. The same applies to the P(1)-Pt(1)-P(2) angle of 98.71(11)° in 2. However, it is less enlarged than that of 1 due to the replacement of a phenyl group with a less sterically demanding methyl group. The S(1)-Pt(1)-O(9) angles, 81.59(11)° and 84.5(2)°, are smaller than 90° due to the restriction imposed by the peri-naphthalene geometry which fixes the sulfur and oxygen atoms. The S(1)-Pt(1)-P(1) angles, [91.14(5)° for 1 and 92.91(11)° for 2], are larger than the O(9)-Pt(1)-P(2) angles, [87.66(11)° for 1 and 85.7(2)° for 2].

There are no significant differences in the bond lengths around the platinum centre between 1 and 2. Whilst the effect of trans influence on bond weakening does not always cause bond lengthening,7 in this case, the Pt(1)-P(1) and Pt(1)-P(2) bond lengths are significantly different as a consequence of the differing trans influence of the oxygen and the SO2. The Pt(1)-P(1) bond (2.3319(15) Å for 1 and 2.338(3) Å for 2) which is trans to SO2 is slightly longer than the Pt(1)-P(2) bond (2.2480(16) Å for 1 and 2.252(3) Å for 2) which is trans to the oxygen.

As expected the peri-distance has been elongated due to the breaking of the sulfur-oxygen bond and insertion of the platinum. The non-bonded sulfur-oxygen distance being 2.842(3) Å for 1 and 2.928(7) Å for 2 compared to the bonded distance of 1.6407(14) Å in the free ligand L. However, the peri-distance in the two complexes is still shorter than the sum of the van der Waals radii for sulfur and oxygen by 12-14%. The increase in

peri-distance causes an increase in the angles of the bay region, with positive splay angles of 9° and 13° being seen for 1 and 2, respectively. This is significantly greater than in 1,8-naphthosultone where there is a negative splay angle of 27° due to the presence of the sulfur-oxygen bond.

(8)

accommodate the diphenylmethylphosphine group without causing great distortion to the naphthalene-carbon framework. In 1,8-naphthosultone the sulfur atom lies on the naphthyl plane and the oxygen atom sits 0.018 Å above the plane. In 1 the sulfur atom now lies 0.426(1) Å below the naphthyl plane and the oxygen atom 0.271(1) Å above the plane. This distortion is due to a combination of the rigid naphthalene backbone, the maintenance of a square planar environment around the platinum centre and the bulky triphenylphosphine groups. The torsion angles also deviate from the free ligand by ca. 5.5-7.0°. In comparison, complex 2 shows little difference in distortion when compared to the free ligand. The oxygen atom has not moved from its original position and the sulfur atom now sits 0.057(1) Å above the plane. The torsion angles show slight differences which can be deemed negligible.

In L O(1)-S(1)-O(9) is 107.01(9)° and O(2)-S(1)-O(9) angle is 107.31(9)°. In compounds 1 and 2 there is a noticeable difference in these angles. The O(1)-S(1)···O(9) angles are 78.20(1)° and 90.96(1)°, and the O(2)-S(1)···O(9) angles are 164.68(1)° and 154.31(1)° for 1 and 2, respectively. O(1)-S(1)···O(9) has been reduced whereas O(2)-S(1)···O(9) has been enlarged, this is a consequence of the S(1)-O(9) peri-bond being broken and the ligand forming a complex with a Pt-PR3 species.

The differences observed between 1 and 2 are due to a combination of steric effects and weak intramolecular interactions. In 1 two phenyl rings, one from each of the two PPh3 groups, adopt face to face alignment leading to a weak π···π interaction (Figure 4) with a Cg(C23-C28)····Cg(C29-C34) distance of 3.605(1) Å. By replacing a bulky phenyl group in 1 with a methyl group in 2, not only is there a reduction in strain on the naphthalene system, but there is also a change in the type and strength of intramolecular interaction. In 2 the weak π···π interaction has been replaced by a stronger C-H····π interaction, where the phenyl ring, from one PPh2Me group, is interacting with a methyl group, from the other PPh2Me group (Figure 4) giving a C11-H11 ····Cg(C25-C30) distance of 2.792(1) Å.

(9)

One would expect that replacing another phenyl with a methyl would result in less steric strain and in turn even less distortion of the naphthalene system. However, for complex 3 this proved not to be the case. In 3 it would appear that the intramolecular interaction may dominate over steric effects, which in turn causes the complex to become more distorted than that of 2, but less than that of 1. Like 2, 3 also has a C-H····π interaction (Figure 5) with a C18-H18····Cg(C19-C24) distanceof2.665(1) Å. This distance is shorter than that in 2, resulting in a stronger interaction which in turn will have more effect on the molecular structure of the overall complex.

Figure 5 The crystal structures of 3 showing the weak C-H···π interaction (with other hydrogen atoms omitted for clarity).

Compound 3 has a P(1)-Pt(1)-P(2) angle of 98.42(13)° (Figure 3), although this is greater than the ideal 90° square planar geometry, it is smaller than those of 1 and 2, which are 99.73(5)° and 98.71(11)°, respectively. The S(1)-Pt(1)-O(9) angle of 83.1(3)° is intermediate to those of 1, 81.59(11)°,and 2, 84.5(2)°, this is due to the dominating intramolecular interactions that occur in 3. The S(1)-Pt(1)-P(1) angles, [91.14(5)° for 1 and 92.91(11)° for 2], are larger than the O(9)-Pt(1)-P(2) angles, [87.66(11)° for 1 and 85.7(2)° for 2]. The same is true for 3 with an S(1)-Pt(1)-P(1) angle of 96.98(14)° and an O(9)-Pt(1)-P(2) angle of 82.6(3)°.

Compound 3 has similar bond lengths around the platinum centre to 1 and 2. The Pt(1)-P(1) bond length of 2.325(4) Å and Pt(1)-P(2) bond length of 2.234(4) Å are significantly different as a consequence of the differing

trans influence of the oxygen and the SO2. This is also seen for 1 and 2.

(10)

sulfur and oxygen by 12-14 %. The increase in peri-distance causes an increase in the angles of the bay region, with positive splay angles of 9°, 13° and 11° being seen for 1, 2 and 3 respectively.

Complex 3 displays distortion of the naphthalene backbone that is intermediate between complex 1 and 2. The non-bonded peri-distance and the splay angle in 3 are also intermediate resulting in a more relaxed geometry than 1, but a more strained geometry than 2. In 1,8-naphthosultone the sulfur atom lies on the naphthyl plane and the oxygen atom sits 0.018 Å above the plane. By comparison, in 3 the sulfur atom now lies 0.379(1) Å above the naphthyl plane and the oxygen atom 0.231(1) Å below the plane. Again this distortion is intermediate to that of 1 and 2 and is due to a combination of the rigid naphthalene backbone, the maintenance of a square planar environment around the platinum centre and the intramolecular interaction that is occurring within the complex. The torsion angles also deviate from the free ligand by ca. 3.5-7.0°.

The O(1)-S(1)···O(9) angle in 3 is 81.00(1)° and the O(2)-S(1)···O(9) angle is 162.24(1)°. The O(1)-S(1)···O(9) angles are 78.20(1)° and 90.96(1)°, and the O(2)-S(1)···O(9) angles are 164.68(1)° and 154.31(1)° for 1 and 2, respectively. Therefore, the angles seen in 3 are intermediate between 1 and 2. This falls into the pattern seen with other features of these complexes and we speculate that this is due to the competition between steric effects and intramolecular interactions, discussed earlier.

As already mentioned, in all three structures the 1,8-naphthosultone acts as a bidentate ligand, coordinating to the platinum via the sulfur and the oxygen atom to form a six-membered chelate ring. This six-membered PtSOC3 ring can be described as having a twisted envelope type conformation with the S···O vector as the hinge (Figure 6). In 1 and 3, C(1), C(10) and C(9) all lie in a plane but S(1) and O(9) are displaced above and/or below this plane, resulting in a non-planar, twisted PtSOC3 ring. In 2 the S(1) and O(9) atoms are located closer to the C3 plane and subsequently the six-membered ring is less twisted. The displacement of Pt(1) from the mean O(9), S(1), C(1), C(10), C(9) plane is comparable for 1 and 3 with distances of 1.343(1) Å and 1.335(1) Å, respectively. Whereas, 2 has a lower Pt(1) displacement of 1.149(1) Å. The angle of the hinge is also comparable for 1 and 3 with values of 54.22(1)° and 54.77(1)°, respectively. 2 has a smaller hinge angle of 44.34(1)°. These differences are due to there being less distortion of the naphthalene backbone in 2.

(11)

Figure 7 X-ray crystallography numbering scheme for 1, 2 and 3.

Table 3 Selected Bond Lengths [Å] and Angles [°] of 1 to 3.

Compound 1 2 3

Peri-region distance and sub-van der Waals contacts

S(1)···O(9) 2.842(3) 2.928(7) 2.889(6)

ΣrvdW - S(1)···O(9); % ΣrvdWa 0.478;86 0.391; 88 0.431; 87

Naphthalene bond lengths

C(1)-C(2) 1.373(8) 1.384(16) 1.36(3)

C(2)-C(3) 1.411(9) 1.388(18) 1.40(3)

C(3)-C(4) 1.354(9) 1.379(16) 1.34(3)

C(4)-C(5) 1.435(9) 1.437(18) 1.41(3)

C(5)-C(10) 1.430(8) 1.450(16) 1.44(3)

C(5)-C(6) 1.408(9) 1.433(16) 1.44(3)

C(6)-C(7) 1.370(9) 1.342(19) 1.37(3)

C(7)-C(8) 1.407(9) 1.415(18) 1.44(3)

C(8)-C(9) 1.380(9) 1.385(14) 1.40(3)

C(9)-C(10) 1.443(9) 1.414(16) 1.42(3)

C(1)-C(10) 1.422(9) 1.432(14) 1.415(18)

Peri-region bond angles

E(1)-C(1)-C(10) 120.7(4) 124.8(9) 123.3(13)

C(1)-C(10)-C(9) 123.2(5) 124.9(10) 123.5(14)

E(9)-C(9)-C(10) 125.0(6) 123.5(9) 124.4(14)

∑ of bay angles 368.9(9) 373.2(16) 371.2(24)

Splay angleb 8.9 13.2 11.2

Bond angles

C(4)-C(5)-C(6) 122.2(6) 119.7(11) 122.4(15)

(12)

O(1)-S(1)···O(9) 78.20(1) 90.96(1) 81.00(1)

O(2)-S(1)···O(9) 164.68(1) 154.31(1) 162.24(1)

C(1)-S(1)-O(1) 105.3(3) 105.2(5) 104.9(7)

C(1)-S(1)-O(2) 107.0(3) 104.6(5) 105.2(7)

Out of plane displacement

O(1) -0.426(1) +0.057(1) +0.379(1)

S(9) +0.271(1) +0.023(1) -0.231(1)

Central naphthalene ring torsion angles

C(6)-C(5)-C(10)-C(1) 174.4(5) -178.7(10) 172.8(12)

C(4)-C(5)-C(10)-C(9) 173.4(6) -178.8(10) 176.6(11)

Platinum geometry - bond lengths

Pt(1)-P(1) 2.3319(15) 2.338(3) 2.325(4)

Pt(1)-P(2) 2.2480(16) 2.252(3) 2.234(4)

Pt(1)-E(1) 2.2764(15) 2.275(3) 2.291(4)

Pt(1)-E(9) 2.067(4) 2.074(7) 2.057(10)

Platinum geometry - bond angles

E(1)-Pt(1)-E(9) 81.59(11) 84.5(2) 83.1(3)

E(1)-Pt(1)-P(1) 169.05(5) 166.09(10) 162.30(13)

E(1)-Pt(1)-P(2) 91.14(5) 92.91(11) 96.98(14)

E(9)-Pt(1)-P(1) 87.66(11) 85.7(2) 82.6(3)

E(9)-Pt(1)-P(2) 171.78(11) 167.7(2) 173.7(3)

P(1)-Pt(1)-P(2) 99.73(5) 98.71(11) 98.42(13)

Intramolecular Interactions

Cg(C23-C28)····Cg(C29-C34) 3.605(1) -

-Cg(C25-C30)····H11-C11 - 2.792(1)

Cg(C19-C24)····H18C18 - - 2.665(1)

a van der Waals radii used for calculations: rvdW(S) 1.80 Å, rvdW(O) 1.52 Å9

(13)

Table 5 Crystallographic data for ligand L and compounds 1-3.

L 1 2 3

Empirical Formula C10H6O3S C48H40Cl4O3P2PtS C38H36Cl4O3P2PtS C26H28O3P2PtS

Formula Weight 206.22 1095.69 971.61 677.60

Temperature (°C) -148.0 -180.0 -180.0 -180.0

Crystal Colour, Habit colourless, chip yellow, prism yellow, prism yellow, prism

Crystal Dimensions (mm3) 0.300 X 0.120 X 0.070

0.150 X 0.080 X 0.010

0.100 X 0.050 X 0.050

0.100 X 0.030 X 0.030

Crystal System triclinic orthorhombic monoclinic monoclinic

Lattice Parameters a = 6.8825(6) Å a = 11.0438(5) Å a = 31.923(14) Å a = 15.555(5) Å

b = 8.0514(7) Å b = 23.6199(13) Å b = 10.964(5) Å b = 9.956(4) Å

c = 8.8857(8) Å c = 33.8528(19) Å c = 22.755(10) Å c = 16.277(6) Å

α = 63.904(2)° α =90° -

-β = 76.979(3)° β = 90° β = 101.374(11)° β = 95.138(9)°

γ = 87.447(3)° γ = 90° -

-Volume (Å3) 429.99(7) 8830.6(8) 7808(6) 2551(2)

Space Group P-1 Pbca C2/c Cc

Z Value 2 8 8 4

Dcalc (g/cm3) 1.593 1.648 1.653 1.793

F000 212 4352 3840 1328

μ(MoKα) (mm-1) 0.348 3.582 4.024 5.804

No. of Reflections Measured 4535 54214 21896 7437

Rint 0.0279 0.0906 0.0749 0.1049

Min and Max Transmissions 0.802 - 0.976 0.1000 - 0.7381 0.524 - 0.818 0.409 - 0.840

Independ. Reflection (No. Variables)

1966(127) 8067(533) 6816(442) 3316(288)

Reflection/Parameter Ratio 15.48 15.14 15.42 11.51

Residuals: R1 (I>2.00σ(I)) 0.0432 0.0460 0.0817 0.0544

Residuals: R (All Reflections) 0.0593 0.0697 0.0996 0.0569

Residuals: wR2 (All Reflections) 0.0964 0.0972 0.2258 0.1302

Goodness of Fit Indicator 1.108 1.091 1.102 1.056

Maximum peak in Final Diff. Map

0.31 e-/Å3 2.279 e-/Å3 7.11 e-/Å3 2.64 e-/Å3

Minimum peak in Final Diff. Map

-0.37 e-/Å3 -1.277 e-/Å3

-2.45 e-/Å3 -1.51 e-/Å3

Conclusion

The work presented in this paper follows on from our previous studies of platinum bisphosphine complexes bearing 1,8-dichalcogen naphthalenes and the oxidised derivatives of naphtho[1,8-cd]1,2-dithiole as ligands.6 Herein, we present the synthesis of a series of three platinum (II) bisphosphine complexes 1-3 [Pt(1-(SO2 ),8-(O)-nap)(PR3)2] (where R3 = Ph3, Ph2Me, Me2Ph), which were prepared by metathesis from cis-[Pt(PR3)2Cl2)] and the dilithium salt of 1,8-naphthosultone.

The 31P{1H} NMR spectra of complexes 1-3 display the expected AX pattern with appropriate platinum satellites. The smaller 1J(31P-195Pt) coupling constant value was assigned to the phosphine group trans to the SO

2

(14)

The naphthalene framework was found to undergo deformation in all three complexes as a result of the platinum metal trying to maintain a square planar geometry. One would expect the degree of molecular deformation to decrease upon going from 1 to 3 as the alkyl group on the phosphorus is changed from Ph3 to Me2Ph. However, a competition between steric effects and intramolecular interactions causes 3 to display distortion intermediate of 1 and 2.

Experimental Section

All experiments were carried out under an oxygen- and moisture-free nitrogen atmosphere using standard Schlenk techniques and glassware. Reagents were obtained from commercial sources and used as received. Dry solvents were collected from a MBraun solvent system. Elemental analyses were performed by Stephen Boyer at the London Metropolitan University. Infra-red spectra were recorded for solids as KBr discs in the range 4000-300 cm-1 on a Perkin-Elmer System 2000 Fourier transform spectrometer. 1H and 13C NMR spectra were recorded on a Bruker Avance 300 MHz spectrometer with δ(H) and δ(C) referenced to external tetramethylsilane. 31P were recorded on a Jeol GSX 110 MHz spectrometer with δ(P) referenced to external phosphoric acid. All measurements were performed at 25 °C. All values reported for NMR spectroscopy are in parts per million (ppm). Coupling constants (J) are given in Hertz (Hz). Mass spectrometry was performed by the University of St. Andrews Mass Spectrometry Service. Electrospray Mass Spectrometry (ESMS) was carried out on a Micromass LCT orthogonal accelerator time of flight mass spectrometer. All

cis-dichlorobis(phosphine)platinum reagents were prepared following standard literature procedures.12

[Pt(1-(SO2),8-(O)-nap)(PPh3)2] (1): Lithium triethylborohydride (1.50 mL of a 1.0 M solution in THF, 1.50

mmol) was added to a solution of 1,8-naphthosultone (0.16 g, 0.78 mmol) in THF (20 mL). An immediate colour change occurred from colourless to pale yellow. After stirring for 30 min, the resulting solution was transferred via a stainless steel cannula to a suspension of cis-dichlorobis(triphenylphosphine)platinum (0.31 g, 0.39 mmol) in THF (10 mL). The mixture was stirred for 24 hours resulting in a yellow solution. The solution was evaporated to dryness under reduced pressure and re-dissolved in the minimum amount of dichloromethane. The product was purified by column chromatography using a dichloromethane/ hexane/ ethyl acetate eluent to yield a yellow solid (0.22 g, 61%); mp 184-186 °C (deccomp.); elemental analysis (Found: C, 57.8; H, 3.2. Calc. for C46H36O3P2PtS.1/2CH2Cl2: C, 57.9; H, 3.8%); IR (KBr disc) vmax cm-1 3464w, 3052w, 1619w, 1555s, 1482s, 1436vs, 1368s, 1316w, 1275s, 1207s, 1188s, 1164w, 1100vs, 1066vs, 1029w, 999w, 912w, 821w, 787w, 745s, 693vs, 645w, 617w, 585w, 542vs, 524vs; δH(300 MHz; CDCl3; 25 °C; Me4Si) 8.05 (1 H, dd, 3JHH 6.9 Hz, 4J

HH 0.6 Hz, Nap 4-H), 7.87 (1H, dd, 3JHH 8.1 Hz, 4JHH 0.8 Hz, Nap 2-H), 7.52-7.45 (6 H, m, P-Phenyl), 7.41-7.35 (1 H, m, Nap 3-H), 7.34-7.24 (11 H, m, P-Phenyl), 7.15-7.04 (13 H, m, Nap 5-H, P-Phenyl), 6.81-6.75 (1 H, m, Nap 6-H), 5.43 (1 H, d, 3J

HH 7.2 Hz, Nap 7-H); δC(75.5 MHz; CDCl3; 25 °C; Me4Si) 134.6 (d, 2JCP 11.4 Hz), 134.2 (d, 2J

CP 11.4 Hz), 131.7(s), 131.0 (d, 4JCP 2.1 Hz), 130.7 (d, 4JCP 2.1 Hz), 128.5-128.2 (m), 127.7(s), 124.2(s), 119.8(s), 116.3(s), 115.6(s); δP(110 MHz; CDCl3; 25 °C; H3PO4) 10.6 (d, 1JP-Pt = 3876 Hz, 2JP-Pt = 19 Hz ), 19.1 (d, 1J

(15)

The other 1,8-naphthosultone complexes were similarly prepared and the physical and spectral data are as follows.

[Pt(1-(SO2),8-(O)-nap)(PPh2Me)2] (2): From 1,8-naphthosultone (0.16 g, 0.79 mmol), lithium

triethylborohydride (1.50 mL of a 1.0 M solution in THF, 1.50 mmol), and

cis-dichlorobis(diphenylmethylphosphine)platinum (0.26 g, 0.39 mmol) to yield a yellow solid (0.06 g, 19%); mp 194-196 °C (decomp.); elemental analysis (Found: C, 53.8; H, 4.0. Calc. for C36H32O3P2PtS: C, 53.9; H, 4.0%); IR (KBr disc) vmax cm-1 3438s, 3052w, 2369w, 2345w, 1555s, 1495w, 1437vs, 1369vs, 1280s, 1201s, 1105vs, 1060vs, 892s, 821w, 788w, 739s, 693s, 645w, 587w, 534s, 516s; δH(300 MHz; CDCl3; 25 °C; Me4Si) 8.14 (1 H, dd, 3J

HH 7.3 Hz, 4JHH 1.1 Hz, Nap 4-H), 7.84 (1H, dd, 3JHH 8.2 Hz, 4JHH 1.0 Hz, Nap 2-H), 7.44-7.37 (6 H, m, Nap 3-H, P-Phenyl), 7.31-7.24 (8 H, m, Nap 5-H, P-Phenyl), 7.17-7.12 (8 H, m, P-Phenyl), 6.85-6.79 (1 H, m, Nap 6-H), 5.85 (1 H, d, 3J

HH 7.5 Hz, Nap 7-H), 2.38 (3 H, d, 2JHH 11.1 Hz, P-CH3), 1.26 (3 H, 2JHH 10.0 Hz, P-CH3);

δC(75.5 MHz; CDCl3; 25 °C; Me4Si);132.9 (d, 2JCP 3.3 Hz), 132.8 (d, 2JCP 3.1 Hz), 132.2(s), 131.6 (d, 4JCP 2.1 Hz), 131.1 (d, 4J

CP 2.2 Hz), 129.2 (d, 3JCP 2.8 Hz), 129.0 (d, 3JCP 3.3 Hz), 128.0(s), 124.7(s), 119.7(s), 116.6(s), 115.6(s), 16.0 (d, 1J

CP 43.2 Hz), 11.6 (d, 1JCP 34.5 Hz); δP(110 MHz; CDCl3; 25 °C; H3PO4) 3.0 (d, 1JP-Pt = 2537 Hz, 2J

P-Pt = 19 Hz ), -2.4 (d, 1JP-Pt = 3700 Hz, 2JP-Pt = 19 Hz); MS (ES+): m/z 823.50 (20%, M+ + Na).

[Pt(1-(SO2),8-(O)-nap)(PMe2Ph)2] (3): From 1,8-naphthosultone (0.16 g, 0.80 mmol), lithium

triethylborohydride (1.50 mL of a 1.0 M solution in THF, 1.50 mmol) and

cis-dichlorobis(dimethylphenylphosphine)platinum (0.22 g, 0.41 mmol) to yield a yellow solid (0.05 g, 17%); mp 186-188 °C (decomp.); elemental analysis (Found: C, 46.2; H, 4.3. Calc. for C26H28O3P2PtS: C, 46.1; H, 4.2%); IR (KBr disc) vmax cm-1 3426w, 3052w, 2963w, 2917w, 1555s, 1496w, 1439vs, 1370vs, 1284vs, 1263s, 1187vs, 1106vs, 1055vs, 1031s, 954s, 912s, 847w, 824s, 789s, 746s, 719w, 694s, 646w, 618w, 585s, 534s; δH(300 MHz; CDCl3; 25 °C; Me4Si) 8.30 (1 H, dd, 3JHH 7.2 Hz, 4JHH 0.8 Hz, Nap 4-H), 7.89 (1H, dd, 3JHH 8.2 Hz, 4JHH 0.8 Hz, Nap 2-H), 7.91-7.88 (1 H, m, Nap 3-H), 7.48-7.08 (12 H, m, Nap 5-H, Nap 6-H, P-Phenyl), 6.91 (1 H, d, 3J

HH

7.0 Hz, Nap 7-H), 1.88 (3 H, d, 2J

HH 11.5 Hz, P-CH3), 1.30 (3 H, 2JHH 10.7 Hz, P-CH3); δC(75.5 MHz; CDCl3; 25 °C; Me4Si) 132.5 (s), 131.4-130.7 (m), 129.3 (d, 3JCP 10.6 Hz), 129.1 (d, 3JCP 10.8 Hz), 128.1(s), 125.0(s), 119.6(s), 117.1(s), 115.3 (s), 14.9 (d, 1J

CP 43.2 Hz), 11.1 (d, 1JCP 34.7 Hz); δP(110 MHz; CDCl3; 25 °C; H3PO4) -9.6 (d, 1J

P-Pt = 2504 Hz, 2JP-Pt = 21 Hz ), -17.6 (d, 1JP-Pt = 3603 Hz, 2JP-Pt = 21 Hz); MS (ES+): m/z 699.59 (45%, M+ + Na).

Crystal Structure Analyses

(16)

instrument with graphite-monochromated Mo Kα radiation (λ = 0.710 73 Å). The data for the complexes was collected and processed using CrystalClear (Rigaku).13 The structures were solved by Patterson or direct methods and expanded using Fourier techniques. Non-hydrogen atoms were refined anisotropically, and hydrogen atoms were refined using a riding model. For 3 the Flack parameter was 0.019(13). All calculations were performed using CrystalStructure14 and SHELXL-97.15 Crystal structure images were generated using Olex2.16

Acknowledgments

The X-ray crystal structures were determined by Prof. Alexandra M. Z. Slawin and Dr David B. Cordes. Elemental analyses were performed by Stephen Boyer at the London Metropolitan University and Mass Spectrometry was performed by Caroline Horsburgh at the University of St Andrews. The work in this project was supported by the Engineering and Physical Sciences Research Council (EPSRC).

Supporting Information

Appendix A. Supplementary data

CCDC 997791-997794 contains the supplementary crystallographic data for sulfam, methylsulfam,

1-3. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html,

or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax:

(+44) 1223-336-033; or e-mail: deposit@ccdc.cam.ac.uk

References

[1] C. A. Coulson, R. Daudel and J. M. Robertson, Proc. R. Soc. London, Ser. A, 1951, 207, 306; D. W. Cruickshank, Acta Crystallogr., 1957, 10, 504; C. P. Brock and J. D. Dunitz, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1982, 38, 2218; J. Oddershede and S. Larsen, J. Phys. Chem. A, 2004, 108, 1057.

[2] V. Balasubramaniyan, Chem. Rev., 1966, 66, 567.

[3] F. R. Knight, A. L. Fuller, M. Bühl, A. M. Z. Slawin and J. D. Woollins, Chem. Eur. J., 2010, 16, 7503 and the references therein.

(17)

[5] B. K. Teo, F. Wudl, J. H. Marshall and A. Krugger, J. Am. Chem. Soc., 1977, 99, 2349; B. K. Teo, P. A. Snyder-Robinson, Inorg. Chem., 1978, 17, 3489; B. K. Teo, P. A. Snyder-Robinson, Inorg. Chem., 1979, 18, 1490; B. K. Teo, P. A. Snyder-Robinson, Inorg. Chem., 1981, 20, 4235.

[6] S. M. Aucott, H. L. Milton, S. D. Robertson, A. M. Z. Slawin, G. D. Walker and J. D. Woollins, Chem. Eur. J., 2004, 10, 1666; S. D. Robertson, PhD Thesis, University of St Andrews (UK), 2005. and the references therein.

[7] B.J. Coe, S.J. Glenwright, Coord. Chem. Rev., 2000, 203, 5; T. G. Appleton, H. C. Clark and L. E. Manzer,

Coord. Chem. Rev., 1973, 10, 335; A. Pidcock, R. E. Richards and L. M. Venanzi, J. Chem. Soc. A, 1966, 1707-1710, F. R. Knight, A. L. Fuller, A. M. Z. Slawin and J. D. Woollins, Polyhedron, 2010, 29 (8), 1956.

[8] D. F. Shriver and P. W. Atkins, Inorganic Chemistry, 3rd ed., Oxford University Press, Oxford, 1999.

[9] A. Bondi, J. Phys. Chem., 1964, 68, 441.

[10] F. H. Allen, O. Kennard, D. G. Watson, L. Brammer, A. G. Orpen and R. Taylor, J. Chem. Soc. Perkin Trans II, 1987, S1.

[11] C. A. Tolman, Chem. Rev., 1977, 77 (3), 313.

[12] D. Drew and J. R. Doyle, Inorg. Synth., 1972, 13, 47; F. J. Ramos-Lima, A. G. Quiroga, J. M. Pérez, M. Font-Bardía, X. Solans, C. Navarro-Ranninger, Eur. J. Inorg. Chem., 2003, 8, 1591; J. Bailar, H. Itatani, Inorg. Chem., 1965, 4, 1618.

[13] CrystalClear 2.0; Rigaku Corp., Tokyo, 2010. CrystalClear Software User’s Guide; Molecular Structure Corporation ©, The Woodlands, TX, 2011.

[14] CrystalStructure 4.0: Crystal Structure Analysis Package; Rigaku and Rigaku/MSC, The Woodlands, TX

77381,200.

[15] G. M. Sheldrick, SHELX97, Acta Crystallogr., Sect. A,, 2008, 64, 112.

[16] OLEX2: O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard and H. Puschmann, J. Appl. Cryst., 2009,

Figure

Table 1  31 P{ 1 H} NMR data for complexes 1 to 3.
Figure 1 X-ray crystallography numbering scheme for L.
Figure 2 The molecular structure of 1,8-naphthosultone (L) with H atoms omitted for clarity.
Figure 3 The angles (°) associated with the square planar geometry of the platinum metal in complexes 1, 2 and 3.
+6

References

Related documents

For the Life of the World is mailed to all pastors and congregations of The Lutheran Church—Missouri Synod in the United States and Canada and to anyone interested in the work

Connections to the equipment (DIGA Recorder, HD Video Camera, Player theatre, Amplifier, etc.) with an HDMI cable allow you to interface them automatically (p. These features

Pressing the MENU button while in the Media Browser mode will access the Picture, Sound and Settings menu options.. Pressing the MENU button again will exit from

Any two-weeks of course INactivity on the CONNECT tutorial at any time or for any reason during the semester will result in you being assigned a final course grade of “FX”.

Economic impact assessments are the most commonly used form of measuring the economic benefits of cultural sector organisations, so a second case study that used this method has

the chromatin, whereas in the former arrangement the entire force transmitted by the kinetochore would converge on a single Cse4 nucleosome (Santaguida and Musacchio 2009).

The present study aimed at comparing the neopterin level of idiopathic infertile and normospermic male groups as well as specifying its relationship with

Figure 2: Genesis of Khadi and Village Industries Commission (KVIC). Today the Khadi and Village Industries have a