You are on page 1of 159

Developing Terminal Phosphinidene Complexes: The Quest for Applicability Continues

Halil Akta 2009

The investigations described in this thesis were carried out in the Division of Organic Chemistry, Department of Chemistry and Pharmaceutical Sciences, Faculty of Sciences, Vrije Universiteit, De Boelelaan 1083, NL-1081 HV Amsterdam, The Netherlands.

This research was partly supported by a TOP grant to KL (no. 700.52.308) of the Council for Chemical Sciences of the Netherlands Organization for Scientific Research (NWO/CW).

Cover and chapter illustrations by Okan Akn Printed by Whrmann Print Service ISBN: 978-90-9024-839-4 Halil Akta, Zaandam, the Netherlands, 2009

VRIJE UNIVERSITEIT

Developing Terminal Phosphinidene Complexes: The Quest for Applicability Continues

ACADEMISCH PROEFSCHRIFT

ter verkrijging van de graad Doctor aan de Vrije Universiteit Amsterdam, op gezag van de rector magnificus prof.dr. L.M. Bouter, in het openbaar te verdedigen ten overstaan van de promotiecommissie van de faculteit der Exacte Wetenschappen op dinsdag 8 december 2009 om 13.45 uur in de aula van de universiteit, De Boelelaan 1105 door

Halil Akta

geboren te Karyaka, Turkije

promotor: copromotoren:

prof.dr. K. Lammertsma dr. A.W. Ehlers dr. J.C. Slootweg

anneme ve babama

Reading Committee:

prof.dr. F.M. Bickelhaupt dr. B. de Bruin prof.dr. C.J. Elsevier prof.dr. E. Hey-Hawkins dr. C. Mller

Contents

Chapter 1

Nucleophilic Phosphinidene Complexes: Access and Applicability General introduction Transition metal ligation Generating nucleophilic phosphinidene complexes Salt Metathesis/elimination Insertion/elimination

11 12 13 14 14 20 22 24 25 25 29 29 34 37 38 41 42

1.1 1.2 1.3 1.3.1 1.3.2 1.3.3 1.3.4 1.3.5 1.3.6 1.4 1.4.1 1.4.2 1.4.3 1.4.4 1.5 1.6

-Hydrogen migration
Oxidation/deprotonation Phosphinidene group-transfer Dehydrohalogenation/addition Reactivity of nucleophilic phosphinidene complexes Reactive 16-electron intermediates RP-transfer Insertion into the M=P bond Cycloaddition to the M=P bond Conclusion Outline of this thesis

Chapter 2

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes Introduction NHC functionalization Theoretical calculations on [(Ring)M(NHC)=PH] Reactivity

49 50 50 52 54

2.1 2.2 2.3 2.4

2.5 2.6 2.7 2.8

Transient species Conclusion Computational Section Experimental Section

55 56 56 58

Chapter 3

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes Introduction Results and Discussion Synthesis of NHC-functionalized group 8 phosphinidenes Synthesis of NHC-functionalized group 9 phosphinidenes Theoretical calculations on [(Ring)M(NHC)=PH] Geometries Energy Decomposition Analysis Conclusion Computational Section Experimental Section

69 70 72 72 74 75 76 77 82 82 83

3.1 3.2 3.2.1 3.2.2 3.3 3.3.1 3.3.2 3.4 3.5 3.6

Chapter 4

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides Introduction Results and Discussion Transient species Mechanism Imido complex Phosphinidene complex Conclusion

95 96 97 100 102 102 103 105

4.1 4.2 4.3 4.4 4.4.1 4.4.2 4.5

4.6 4.7

Computational Section Experimental Section

106 107

Chapter 5

3-Diphosphavinylcarbene:
A P2 Analogue of the Dtz Intermediate 119 120 121 123 124 126 128 128 128

5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8

Introduction Synthesis of 3-diphosphavinylcarbenes Synthesis of 1,3-diphospha-3H-indenes Mechanism Reactivity Conclusion Computational Section Experimental Section

Appendix 1 Appendix 2 Samenvatting Curriculum Vitae List of publications Dankwoord

141 143 147 153 155 157

Chapter 1
Nucleophilic Phosphinidene Complexes: Access and Applicability

Halil Aktas, J. Chris Slootweg, and Koop Lammertsma*,

Department of Chemistry and Pharmaceutical Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands

Angew. Chem. 2009, accepted

Abstract: The topic of the research described in this thesis concerns the syntheses, properties, and reactivities of nucleophilic phosphinidene complexes LnM=PR. Emphasis is placed on the electronic tuning of this emerging class of phosphorus reagents, using different ancillary ligands and coordinatively unsaturated transition metal moieties. The difference in applicability of the established stable 18-electron and transient 16-electron phosphinidenes is addressed.

Chapter 1

1.1

General introduction

Phosphinidenes[1] (phosphanylidenes,[2] RP:, A; Scheme 1) are the phosphorus analogues of carbenes (R2C:)[3] and nitrenes (RN:).[4] These exceedingly reactive phosphorus species have been detected only in the gas phase (MS) and in glassy and cryogenic matrices (EPR, IR, UV).[5] Their chemistry remains to be fully explored,[6] which distinguishes them from the explosive growth the carbenes saw. Terminal transition metal-complexed phosphinidenes LnM=PR (B; Scheme 1), which are the phosphorus analogues of the wellestablished carbene complexes, appear, however, to be valuable synthons with rapidly expanding chemistries.[7-10] Enabling were the discoveries in the 1980s of a transient electrophilic species, (OC)5W=P Ph, by the group of Mathey[11] and that of an isolable nucleophilic phosphinidene complex, Cp2W=PMes*, by Lapperts group.[12] Illustrative of the ensuing rapid progress in this field are two reviews by Cowley, one in 1988, entitled The quest for terminal phosphinidene complexes[13a] and the other in 1997 Terminal phosphinidene and heavier congeneric complexes. The quest is over.[13b] Much has happened since as is emphasized in the present thesis, which focuses on neutral nucleophilic 1-phosphinidene complexes. Hitherto, these compounds were considered to have limited applicability, sharply contrasting the electrophilic ones,[1,8] but their potential is far greater than believed.

R P A

MLn R P B

Scheme 1. Free (A) and 1complexed phosphinidene (B).

12

Nucleophilic Phosphinidene Complexes: Access and Applicability

1.2

Transition metal ligation

It is important to recognize the impact of a transition metal group on the phosphinidene RP. Terminal complexed phosphinidenes strongly prefer a singlet ground state, contrasting the uncomplexed species, and are either nucleophilic (Schrock-type)[14] or electrophilic (Fischertype)[15] at the phosphorus atom. An extensive density functional study[16] on LnM=PH (M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, W, Fe, Ru, Os, Co, Rh, Ir and L = CO, PH3, Cp) revealed that the philicity and chemical reactivity of the phosphinidene complex is influenced mainly by the metals spectator ligand L. Those with strong -donor capabilities increase the electron density on the phosphorus atom, enhancing its nucleophilicity. Conversely, spectator ligands with strong -acceptor capabilities electrophilic lower the charge Illustrative concentration is the on P, causing between

behavior.

difference

electrophilic (OC)4Fe=PH and nucleophilic Cp2Cr=PH, which is a reflection of the different magnitude in which charge is transferred from the frontier orbital energies of the transition metal fragments to the phosphorus atom. Indeed all reported phosphinidene complexes with only CO ligands, like (OC)nM=PR (M = W, Mo, Cr, n = 5; M = Fe, n = 4), are known to be transient electrophiles, generated in-situ from appropriate precursors. Their insertion into -bonds, addition to bonds, and coordination to lone pairs is well documented and reviewed.[1,8] More diversity in ancillary ligands is available in cationic complexes [LnM=PR]+ of which stable ones[17] with limited reactivity have been reported.[17e-f,18] The diversity in ligands and transition metals is by far the largest for the nucleophilic phosphinidene complexes, the topic of this chapter. Before advancing, it must be noted that the M=P bond of all LnM=PR complexes have genuine double bond character as established by a DFT bond energy analysis
13

Chapter 1

from which quantitative - and

-bond strengths could be

determined.[16] The M=P interaction increases on going from the first to the second- and third-row transition metals. This chapter emphasizes first the different methodologies to access these entities and then addresses their chemical applicability.

1.3 1.3.1 The

Generating nucleophilic phosphinidene complexes Salt Metathesis/elimination most common route toward nucleophilic phosphinidene

complexes is by combining a metal complex with a halogenated species under expulsion of M+X. There are two possibilities, reacting a Li+ metallocene hydride with a chlorophosphine and conversely reacting a transition metal halide complex with a lithium phosphide, but variations on this salt metathesis/elimination theme exist.

Scheme 2. Salt metathesis with dichlorophosphines.

The

first

stable

18-electron

phosphinidene

complexes

were

synthesized by Lappert and co-workers[12] who reacted lithium metallocene hydride [Cp2MHLi]4 with dichlorophosphine RPCl2 (R = Mes*, (Me3Si)2CH) to obtain Cp2M=PR 1a,b (M = Mo (a), W (b)) as stable, red crystalline material (Scheme 2).[12] The low-field
31P

NMR

chemical shifts (R = Mes*; Mo (1a): () 799.5 and W (1b): 661.1) proved to be characteristic for terminal phosphinidene complexes.[12] The Xray crystal structures show a M=PMes* double bond of 2.370(2) for the molybdenum complex[12] and of 2.349(5) for the tungsten

14

Nucleophilic Phosphinidene Complexes: Access and Applicability

complex[12] with bent MPMes* angles of 115.8(2) and 114.8(5), respectively. The group of Stephan synthesized the first early-transition metal complex, zirconium phosphinidene Cp2(Me3P)Zr=PMes* (4), by salt metathesis of zirconocene dichloride and lithium supermesitylphosphide[19] and from zirconium phosphide [Cp2(Cl)ZrP(H)Mes*] using an alkali metal base,[20] both in the presence of PMe3.[19,20] The X-ray crystal structure reveals a short Zr=PMes* double bond of 2.505(4) , a ZrPMes* angle of 101.4(1), and a long ZrPMe3 single bond of 2.741(5) , indicating weak bonding of the ancillary ligand. The
31P

NMR chemical shift of the phosphinidene is observed at 792.7

ppm. A superior route with a near quantitative yield is the reaction of chloro-bis(5cyclopentadienyl)methylzirconium (2) with lithium

supermesitylphosphide followed by loss of methane from the incipient 3 in the presence of PMe3 (Scheme 3).[19b] Using a similar procedure, Protasiewicz and co-workers reported the related phosphinidene complex Cp2(Me3P)Zr=PDmp 5 (Dmp = 2,6-Mes2C6H3, Scheme 3).[21] Salt metathesis and Lewis base stabilization also enabled the synthesis of hafnium phosphinidene Cp2(Me3P)Hf=PMes* 6,[22] the terminally bonded phosphanylphosphinidene complex Cp2(PhMe2P)Zr=PPtBu2 7,[23] and the uranium complex Cp*2(Me3PO)U=PMes* 8 (Scheme 3),[24] all of which are bent according to their data (
31P 31P

NMR spectroscopic

671 (6), 728 (7), 71 ppm (8)) and solid state structures (M

PC/P 115.53(16) (7), 143.7(3) (8)). Salt metathesis was shown by the Lammertsma group to be equally applicable to the late transition metal iridium, tolerating different ancillary ligands, such as PPh3 and the N-heterocyclic carbene IiPr2Me2 together with a Cp* ligand.[25]

15

Chapter 1

Me Cp2Zr Cl 2

LiPHMes*

Mes* Zr PH Me 3

PMe3 - CH4 Zr P

Mes* PMe3 4

tBu Dmp Zr P PMe3 5 6 Hf P PMe3 7 Mes* Zr P PPhMe2 8 P tBu U P O=PMe3 Mes*

Scheme 3. Metal-complexed phosphinidenes synthesized using salt metathesis and Lewis base stabilization.

Reacting iridium dichloride complex 9 (a: PPh3; b: IiPr2Me2) with LiPHMes* provided 10a,b (Scheme 4).[25] NHC-ligated iridium

phosphinidene complex 10b, characterized by an X-ray crystal structure, strongly resembles that of phosphane complex 10a, both showing the expected bending (IrPMes*: 113.73(7) 10a; 110.76(6) 10b) for a phosphinidene complex with typical M=P double bonding (10a: Ir=P 2.2121(5) ; 10b: Ir=P 2.1959(5) ).[25] The difference in their
31P

NMR resonances (10a: 686.6; 10b: 560.0 ppm) is caused by the

strong donor and moderate acceptor capabilities of the NHC ligand rather than by geometrical differences.[25]

IrCl2 L 9

2 LiPHMes* - H2PMes* - 2 LiCl a L = PPh3 b L = Ii Pr2Me2 L

Mes* Ir P N N Ii Pr 2Me2

10

Scheme 4. Phosphane (10a) and N-heterocyclic carbene (10b) functionalized iridium phosphinidenes.

16

Nucleophilic Phosphinidene Complexes: Access and Applicability

A series of tantalum phosphinidenes [N3N]TaPR 12 ([N3N] = (Me3SiNCH2CH2)3N; R = tBu, Cy, Ph) was reported by Schrock and coworkers who condensed tantalum dichloride complex 11 with lithium phosphides (Scheme 5).[26] The large tetradentate triamidoamine ligand [N3N] with trimethylsilyl groups ensures effective stabilization of the nucleophilic phosphinidene unit, but narrows the space available to it. As a result, the TaPR geometry is almost linear (Cy: 170.9) enforcing both TaP pseudo-triple bonding (Cy: 2.145(7) ) and a high field
31P

NMR chemical shift for the phosphinidene (175.1227.3

ppm). The mechanism by which the phosphinidene complex is generated is not clear. The TaP multiple bond may be formed via dehydrohalogenation with a second phosphide acting as base, whereas a proposed alternative path uses proton abstraction from the tantalum bisphosphide.
R Me3Si P SiMe3 Me3Si Cl Cl SiMe3 N Ta N N Ta N 2 LiPHR Me3Si Me3Si N N -2 LiCl, - RPH2 N N R = t -Bu, Cy, Ph 12 11 0 for R = Ph Li 1 R -PhLi Me3Si Me3Si P N Ta N N 14 SiMe3 N R1X -LiX Me3Si Me3Si N Ta N N N 13 PLi SiMe 3

R1 = Me, Bu, SiMe3, SiMe2Ph

Scheme 5. Linear tantalum phosphinidenes resulting from metathesis and PC bond cleavage.

Interestingly,

the

phosphinidene

substituent

PR

of

12

is

exchangeable. Reaction of the phenyl derivative with lithium


17

Chapter 1

afforded terminal phosphido complex {[N3N]TaP} 13 (Scheme 5),[27] which has a lowfield 31P NMR resonance at 575 ppm in concurrence with a phosphide complex. Subsequent reaction at 35 C with organic halides afforded tantalum phosphinidene complexes

[N3N]Ta=PR 14 (R = Me, n-Bu, SiMe3, SiMe2Ph).


Ar N CH 3 AgOTf Ti 0 N CH - Ag 3 Ar 15 Ar N CH 3 Ti OTf N CH 3 Ar 16 B(C 6F5)3 Ar H C N H H Ti 18 N P Is Ar Ar N CH3 Ti N P Is Ar 17 B(C 6F5)3

LiPHIs - LiOTf - CH 4

Ar = 2,6-iPr2C6H3 Is = 2,4,6-i Pr3C6H2

Scheme 6. Titanium phosphinidene formed by salt metathesis followed by hydrogen abstraction and methide elimination.

Mindiolas group used the sterically hindered -diketiminate ligand to react a titanium complex with a lithium phosphide.[28] The process starts by one electron oxidation of titanium dialkyl complex 15 with AgOTf, followed by reaction of 16 with LiPHIs (Is = 2,4,6-iPr3C6H2), presumably giving putative titanium phosphide

(tBunacnac)Ti(Me)2PHIs and on loss of methane phosphinidene complex (tBunacnac)(Me)Ti=PIs 17 (nacnac


31P

= NMR

[Ar]NC(tBu)CHC(tBu)N[Ar]) (Scheme 6). The diagnostic

chemical shift at 231.5 ppm, the large TiPIs angle of 159.95(7), and short Ti=P bond of 2.1644(7) reveal a pseudo-linear titanium

phosphinidene complex. Treatment with tris(pentafluorophenyl)borane caused methide abstraction to yield the terminal

phosphinidene zwitterion (tBunacnac)TiPIs{H3CB(C6F5)3} 18 of which


18

Nucleophilic Phosphinidene Complexes: Access and Applicability

the X-ray crystal structure showed a short Ti=P bond (2.1512(4) ), a linear TiPIs unit (176.03(5)), and an essentially departed methide group (TiCH3 2.405(3) ).

Scheme 7. Niobium phosphinidenes generated via P4 activation.

Another protocol using a transition metal-complexed phosphide was developed by Cummins and Figueroa. With niobaziridine-hydride complex [Nb(H)(2tBu(H)C=NAr(N[Np]Ar)2] 19 (Np = neopentyl, Ar = 3,5Me2C6H3)[29a] they activated white phosphorus (P4) to form bridged diphosphide complex [(2:2,2P2){Nb(N[Np]Ar)3}2] 20, which on treatment with sodium amalgam gave monomeric 21 (Scheme 7). The formation of this terminal phosphide, having a
31P

NMR chemical

shift of 1010 ppm (!), was supported by an X-ray crystal structure that confirmed its anion-cation separation. Reaction with organic halides resulted in niobium phosphinidene complexes (Ar[Np]N)3Nb=PR 22 (R
19

Chapter 1

= SiMe3, SnMe3, PPh2, PtBu2).[29b] The X-ray crystal structure for the SnMe3 derivative reveals an elongated Nb=P bond (2.2731(8) ) and a PSn single bond with a length (2.4778(8) ) that matches the sum of the covalent radii of phosphorus and tin. The
31P

NMR resonances

(401.3607.0) of the niobium complexes 22 are in accord with a bent phosphinidene.

1.3.2

Insertion/elimination

Inserting an electron-deficient organometallic fragment into a (R)P=X (X = CO, CNPh) bond is an alternative route to phosphinidene complexes. Cowley and co-workers[30] prepared tungsten

phosphinidene (MePh2P)2Cl2W(CO)PMes* 24 by reacting the 16electron tetraphosphine complex (MePh2P)4Cl2W 23 with

phosphaketene (Mes*)P=C=O under elimination of two equivalents of phosphine (Scheme 8). In the phosphinidene product the ketenes RP (axial) and CO (equatorial) moieties end up in a syn fashion, but the mechanism of formation is not known. The short WP distance of 2.169(1) suggest triple bonding. The large WPMes* angle of 168.2(2)) is also reflected in the upfield
31P

NMR chemical shift of

193.0 ppm. The similar reaction with (Mes*)P=C=NPh likely gives the thermally unstable complex (MePh2P)2Cl2W(C=NPh)PMes*.
Mes* O C PMes* - 2 eq. PMePh2 55% 23 P CO Ph2MeP W PMe Ph2 Cl Cl 24

(Ph2MeP)4Cl2W

Scheme 8. Tungsten phosphinidene by P=C double bond cleavage.

Another example concerns the insertion into a PH bond. Oxidative addition of sterically unhindered phenylphosphine to the electron20

Nucleophilic Phosphinidene Complexes: Access and Applicability

poor tris-siloxy tantalum complex 25 (Scheme 9) reportedly gives intermediate phosphide 26 and on 1,2H2 elimination tantalum phosphinidene complex (tBu3SiO)3Ta=PPh 27.[31] The large siloxy groups ensure kinetic stabilization of the bent phosphinidene. Its X-ray crystal structure shows a short Ta=P double bond of 2.317(4) and bent TaPPh moiety of 110.2(4). The preference of a bent over a linear phosphinidene complex was suggested to originate from O(p)Ta(d) backbonding, which prevents the formation of an otherwise more favorable P(p)Ta(d) interaction.
H2PPh t-Bu3SiO t-Bu3SiO t-Bu3SiO Ta Ph PH C 6H6 - H2 t-Bu3SiO t-Bu3SiO t-Bu3SiO Ta P 27 Ph

(t-Bu3SiO)3Ta 25

H 26

Scheme 9. Tantalum phosphinidene formed by PH bond cleavage and subsequent 1,2H2 elimination.

In situ generation of a transient organometallic precursor is also a viable option. For example, compound 28 undergoes reductive elimination and reacts with one equivalent of the primary phosphine Mes*PH2 in the presence of a threefold excess of PMe3 to afford the isolable 29 (Scheme 10).[32]
Cp Ti N tBu3 P 28 tBu3P 29
3 PMe3 , H2PMes*

Cp Ti N

PMe3 P Mes*

Scheme 10. Titanium phosphinidene formed by reductive elimination.

Although

the

structure
31P

of

29

could

not

be

confirmed

crystallography, its

NMR resonances observed at 769.9, 35.3 and

21

Chapter 1

10.3 ppm are indicative for the terminal Tiphosphinidene fragment, the phosphinimide ligand, and the coordinated PMe3, respectively. Additional 1H and 13C{1H} spectra were consistent with this formulation.

1.3.3

-Hydrogen migration

-Hydrogen migration of the initial salt metathesis product is another


route to phosphinidene complexes. The first spectroscopic evidence for such a process was reported by Nieckes group for aminosubstituted complex Cp*2M=PN(H)Mes* 32a,b (M = Mo (a), W (b); Scheme 11a),[33] which is similar to the complex Cp2M=PMes* 1a,b reported by Lappert.[12] Reaction of metal hydride 30a,b with chloroiminophosphine ClP=NMes* is believed to give intermediate 31, based on the observed
31P

NMR resonance at 754 ppm at 40C for

the tungsten complex. Above this temperature the metal hydride presumably undergoes a 1,3hydrogen shift to yield phosphinidene complex 32, which was characterized by
31P

NMR spectroscopy (Mo:

770, W: 663 ppm), but could not be isolated. A 1,3-shift of an -hydrogen was used by Mindiolas group to prepare the titanium and vanadium phosphinidene complexes (nacnac)(CH2tBu)M=PR 35a,b (M = Ti (a), V (b); Scheme 11b). Salt metathesis of titanium alkylidene 33a with LiPHR (R = Cy, Is, Mes*) at low temperature gave putative neopentylidene-phosphide 34a that underwent

-hydrogen

migration

to

give

phosphinidene

(nacnac)(CH2tBu)Ti=PR 35a, (Scheme 11b).[34a] The Mes* derivatives has a short TiP pseudo-triple bond (2.1831(4) ) and a pseudo-linear TiPMes* unit (164.44(5)),[34b] while solution spectra reveal two
31P

NMR resonances (242 and 216 ppm), suggesting the presence of two conformers.[34a] Paramagnetic vanadium complexes (nacnac)(CH2t Bu)V=PR 35b (R = Is, Mes*) were synthesized analogously (Scheme
22

Nucleophilic Phosphinidene Complexes: Access and Applicability

11b).[34c] The X-ray crystal structures reveal a distorted tetrahedral geometry at vanadium, a V=PR bond (R = Is, 2.174(4) ; Mes*, 2.1602(6) ) that is considerably shorter than those for the fourcoordinate vanadium phosphides, and a VPCipso angle that depends on the P-substituent (Is, 136.6(5); Mes*, 153.28(6)).
Cp* a) Cp* M H Li ClP NMes* - LiCl Cp* M Cp* H
31

P NMes*

1,3

Cp* M Cp* P

H N Mes*

30 M = Mo (a), W (b)

32

b)

Ar H N C M N X Ar
33

LiP(H)R - LiX

Ar H N C M N P(H)R Ar
34
i

68-79%

Ar H2 N C M N P R Ar
35

M = Ti (a), V (b); Ar = 2,6- Pr2C6H3; R = Is, Mes*; X = I, OTf

Pi Pr2 c) N Ti Pi Pr2
36

H2 C P Is

Cp d) W P

Mes*

Et3N -Et3N.HCl

Cp W P OC CO
38

Mes* H

H OC Cl H OC
37

KOt-Bu 79%

Cp OC OC H W P

Mes*

39

Scheme 11. Phosphinidene complexes formed by hydrogen migration.

The

generality

of

the

-hydrogen migration[28] was further

demonstrated by the synthesis of titanium(IV)phosphinidene 36 (Scheme 11c), bearing the PNP-pincer ligand N[2-P(CHMe2)2-4methylphenyl]2; imide and alkylidene functionalities can be obtained
23

Chapter 1

by the same approach.[34d] Spectroscopic evidence for a baseinduced 1,2-H shift leading to a phosphinidene complex was provided by the group of Malish (Scheme 11d).[35]

Dehydrohalogenation of phosphine complex 37 with triethylamine gave phosphenium complex 38, which underwent a 1,2-H shift in the presence of KOtBu, likely by a deprotonation-reprotonation

sequence, to yield the evidently preferred phosphinidene complex Cp(CO)2HW=PMes* (39). Whereas this product eluded isolation, it was characterized by its
31P

NMR chemical shift at 819.9 ppm (1JPW =

123 Hz) and a hydride signal in the 1H NMR at 10.03 ppm.

1.3.4

Oxidation/deprotonation

One-electron oxidation of paramagnetic nickel(I) phosphido complex 40 using tropylium hexafluorophosphate was shown by the Hillhouse group to give cationic complex 41 that can be

deprotonated with a strong base to afford nickel(II) phosphinidene complex (dtbpe)Ni=PDmp 42 (dtbpe = 1,2-bis(di-t-

butylphosphino)ethane, Dmp = 2,6-Mes2C6H3; Scheme 12).[36] The structure has a short Ni=P bond of 2.0772(9) and a bent NiPC unit with an angle of 130.78, which is also reflected by its resonance at 970 ppm ( 2JPP = 134 Hz).
PF6- R R Dmp Ni P P RR 40 H P RR 41 P Ni P H Dmp
31P

NMR

R R P

PF6NaN(TMS)2

R R P Ni P P RR 42 Dmp

R= t-Bu; Dmp = 2,6-Mes2C6H3


Scheme 12. Nickel phosphinidene complex formed by an oxidation-deprotonation sequence.

24

Nucleophilic Phosphinidene Complexes: Access and Applicability

1.3.5

Phosphinidene group-transfer

The groups of Mindiola and Protasiewicz jointly demonstrated that phospha-Wittig reagents of the type Me3P=PAr are effective as PAr transfer reagents The for the synthesis of stable phosphinidene reagents

complexes.[37]

phosphanylidene-4-phosphorane

Me3P=PAr (Ar = 2,4,6-tBu3C6H2 and 2,6-Mes2C6H3) can deliver PAr fragments to low-valent early-transition metal complexes 43 and 44 to respectively effect oxidation to form terminal Zr(IV) phosphinidene 5 (Scheme 13a) or generate terminal vanadium(V) phosphinidene complex 45 (Scheme 13b).

Scheme 13. Phosphinidene group-transfer with phospha-Wittig reagents.

1.3.6

Dehydrohalogenation/ligation double dehydrohalogenation of appropriate

Base-induced

precursors in the presence of a suitable donor ligand was used by Lammertsma et al. to synthesize various Group 8 and 9

phosphinidene complexes in an one-pot procedure (Scheme 14a,b). Illustrative is the formation of iridium phosphinidene complex (Cp*)(Ph3P)Ir=PMes* 10a by dehydrohalogenating phosphine
25

Chapter 1

complex 46a with two equivalents of DBU (1,8-biazabicyclo-[5.4.0]undec-7-ene) and capturing the putative 16-electron [(Cp*)Ir=P Mes*] by the donor ligand PPh3 (Scheme 14c).[25a] This mild

procedure proved equally effective for various other donor ligands L (PH2Mes*, PMe3, P(OMe)3, dppe, AsPh3, tBuNC, XyNC, CO; Scheme 14a).[25a] The X-ray crystal structure of the CO ligated complex (Cp*)(CO)Ir=PMes* (Ir=P 2.1783(8) , IrPMes* 113.77(10)) shows a Z conformation for the double bond, which is attributed to the small size of the CO ligand and differs from that of the PPh3 ligated complex that has an E conformation but with otherwise similar structural features. The deshielded
31P

NMR resonance for the CO

ligated complex (805 ppm), as compared to the phosphane analogue (687 ppm), was attributed to the CO acceptor capabilities rather than to geometrical differences.[25] The X-ray crystal structure for the only isolable cobalt phosphinidene complex (Cp)(Ph3P)Co=PMes* (10c) shows a piano-stool geometry with bonding properties (Co=P 2.1102(8) ; CoPMes* 109.00(9)) that closely resemble iridium analogue 10a.[38] The CO-ligated Cocomplex Z(Cp)(CO)Co=PMes* could be observed by its

characteristic low-field 31P NMR resonance at 1047 ppm (!). Illustrative for group 8 phosphinidenes are the ruthenium complexes (6Ar)(L)Ru=PMes* 48a (Ar = benzene, p-cymene; L = PPh3, PMe3, tBuNC; Scheme 14b) with
31P

NMR resonances in the range 801846

ppm, while the non-isolable CO-ligated complex (6Bz)(CO)Ru=P Mes* reportedly has a chemical shift at lower field (897 ppm).[39] Ruthenium phosphinidenes (6pCymene)(R3P)Ru=PMes* 48a (R = Ph, Cy) are also accessible from (6pCymene)RuCl2(PR3) by reaction with DBU and PH2Mes*.[40] The heavier osmium phosphinidene complexes (6Ar)(L)Os=PMes* 48b (Ar = benzene, p-cymene; L =
26

Nucleophilic Phosphinidene Complexes: Access and Applicability

PPh3,

PMe3,

CO)

were

obtained

equally

readily

from

dehydrohalogenating/ligating the primary complexed phosphine (6Ar)OsX2(PH2Mes*) 47b.[39] Similar to the third row group 9 transition metal iridium, also a carbon monoxide ligated osmium complex, (6 Ar)(CO)Os=PMes* (48b), could be isolated. The E isomers were shown to be favored for complexes having large PR substituents (e.g. Mes*) and bulky ligands (e.g. PPh3), the Z isomer for the smaller carbon monoxide ligand, and a mixture of E/Z products for ligands of intermediate size, such as PMe3, or P(OMe)3.[25,38]

Scheme 14. Dehydrohalogenation/ligation for Group 8 (a) and 9 (b) transition metal phosphinidene complexes; c) Synthesis of phosphinidene complex 10a via putative [(Cp*)Ir=PMes*] from iridium precursor 46a.

Dehydrohalogenation/ligation proved also an effective route to introduce Nheterocyclic carbene ligands, such as IiPr2Me2 (1,327

Chapter 1

diisopropyl-4,5-dimethylimidazol-2-ylidene).[41] For example, the onepot reaction of rhodium and iridium precursors 46a,b and ruthenium and osmium precursors 47a,b with three equivalents of IiPr2Me2 yielded the corresponding Group 8 and 9 NHC-ligated

phosphinidene complexes 49,10b and 50a,b respectively (Scheme 15).[42,43] Because NHC is the stronger base (pKa: IiPr2Me2 24.0 in d6DMSO),[44] DBU (pKa: 11.3)[45] cannot be used for the dehydrohalogenation,[25a,38,39,46] thereby necessitating the use of two

equivalents of NHC as base and one as stabilizing ligand. However, the imidazolium salt IiPr2Me2HCl that precipitates can be regenerated to the carbene by deprotonation.[47] The NHC-functionalized

phosphinidene complexes were obtained as colored, air and moisture sensitive, but thermally stable solids with characteristic
31P

NMR resonances (Ir: 560.0 (10b);[25b] Rh: 745.9 (49);[43] Ru: 751.7 (50a);[42] Os: 557.6 (50b)[43]) that reflect shielding due to -donor capacity of the NHC ligand.
5-Cp*
3 Ii Pr2Me2 46, 47 Toluene -2 IiPr2Me2.HCl N M N P N Mes*

6-Ar
M P N

Mes*

49 M = Rh, Ir (10b)

50 M = Ru (a), Os (b); Ar = C 6H6, p-Cymene

Scheme 15. Dehydrohalogenation/ligation using NHCs.

The

X-ray

crystal

structure

of

rhodium

complex

(5-Cp*)(IiPr2Me2)Rh=PMes* 49 shows a Rh=P bond length of 2.1827(7) and an MPMes* angle of 107.65(4) that are similar to those for iridium
28

complex

10b

and

ruthenium

complex

Nucleophilic Phosphinidene Complexes: Access and Applicability

(6-Bz)(IiPr2Me2)Ru=PMes*

50a

(Ru=P

2.2222(8)

RuPMes*

105.82(10)). Both phosphinidene complexes have pronounced MC single bonds with lengths of 2.036(2) (Rh) and 2.091(3) (Ru) that are in the typical range for MNHC complexes.[48]

1.4 To

Reactivity of nucleophilic phosphinidene complexes review the reactivities of the nucleophilic, 18electron

phosphinidene complexes it is relevant to recognize the impact of the putative 16electron [LM=PR]. Whereas their involvement in the reactions often cannot be ascertained, they are evolving as reactive entities that we address here separately.

1.4.1

Reactive 16-electron intermediates

Convincing spectroscopic evidence has been presented for the reactive, 16electron intermediates [LM=PMes*] (L = 5-Cp(*),6-Ar). They presumably form in-situ on dehydrohalogenating primary phosphine complexes, like 46 and 47, and are than captured by a ligand to give the discussed 18-electron phosphinidene complexes. Slowing down the ligation by using the heavily congested carbene IMes (1,3-dimesityl-imidazol-2-ylidene)[42,43] both as base and as ligand in the reaction with (6-pCymene)RuCl2(PH2Mes) 47c gave besides the expected phosphinidene complex 52 also isomer 53 in which the pCymene group is replaced by a (d3-)toluene solvent molecule, indicating the conversion of the 16-electron intermediate

[(6-pCymene)Ru=PMes] 51a to [(6-Toluene)Ru=PMes] 51b (Scheme 16a). Solvent stabilization was demonstrated for the putative 16electron complex carrying the bulky phosphorus substituent 2,6dimesitylphenyl (Dmp). Monitoring by
31P

NMR spectroscopy the

reaction of 54 with DBU in dichloromethane (Scheme 16b) showed


29

Chapter 1

the appearance of a characteristic low-field resonance at 672 ppm, which concurs with the 684 ppm calculated at BP86/TZP for dichloromethane-solvated phosphinidene 55.[49]

Scheme 16. Solvent-stabilized phosphinidene complexes.

In the absence of a stabilizing donor ligand, HeyHawkins and coworkers reported that on dehydrohalogenating tantalum-

complexed primary phosphine Cp*(Cl4)Ta(PH2Is), the 14-electron phosphinidene complex [Cp*(Cl2)Ta=PIs] (56:
31P

488.0 ppm) was

isolated; no mention was made of intermediate products.[50] Lammertsmas group showed that dehydrogenation of primary phosphine complex 46a with the strong phosphazene base tbutylimino-tri(pyrrolidino)phosphorane (BTPP; pKB 26) in the absence of a ligand gave 18-electron complex 57 and dimer [Cp*IrCl2]2 58 (Scheme 17a).[51] Phosphinidene 57 was interpreted to result from [Cp*Ir=PMes*], abstracting PH2Mes* from its precursor. Reaction at low temperatures showed the intermediate formation of 59 (Scheme 17b),
30

Nucleophilic Phosphinidene Complexes: Access and Applicability

resulting

from

[2+2]-cycloaddition
31P

of

[Cp*Ir=PMes*]

and

[Cp*(Cl)Ir=P(H)Mes*], as identified by its

NMR resonances (366 and

126 ppm), and suggests that the first dehydrohalogenation is faster than the second one. The reaction appears sensitive to the size of the substituent on phosphorus as the smaller Mes group gave the intermediate dimetallocycle 61, resulting from dimerization of [Cp*(Cl)Ir=P(H)Mes], and subsequently on dehydrohalogenation dimer 62 (Scheme 17c).

Scheme 17. Dehydrohalogenation in the absence of a stabilizing ligand.

Stephan and co-workers reported on the generation and reactivity of the transient, 16-electron phosphinidene [Cp*2Zr=PR] 64[19] (Scheme 18) that can be generated from primary phosphide complex 63 by elimination of phosphane H2PR. Attempts to isolate 64 were unsuccessful. However, it could be detected by
31P

NMR
31

Chapter 1

spectroscopy at 537 ppm as an unstable [Cp*2Zr=PMes]LiCl adduct when prepared from Cp*2ZrCl2 and LiPHMes in dimethoxyethane (DME).[19a,b] In-situ-generated 64 (R = Mes) is highly reactive and gives 65 by intramolecular C-H insertion and yields metallocycle 66 and phosphane H2PR by reaction with acetonitrile (Scheme 18).[19] The reinsertion of phosphane H2PR to the Zr=P bond of 64 is also feasible and yields complex 67 (and 68 upon reaction with MeCN) irreversibly with elimination of H2 as the deriving force (Scheme 18).[19]

Scheme 18. Formation and reactivity of [Cp*2Zr=PR].

Stephan and co-workers reported other examples of the transient 16-electron [Cp(*)2Zr=PR] 64 (R = SiPh3: 263, Mes*: 478, Cy: 499, Mes: 526, Ph: 579 ppm), all characterized by their
31P

NMR chemical

shifts.[19,20] Majoral and co-workers reported 31P NMR chemical shifts for [Cp(*)2Zr=P(2,4,6(MeO)3C6H2)] 69a,b (Cp: 465, Cp*: 438 ppm) using salt the metathesis approach (Scheme 19a), but dimers and polymeric forms could not be excluded.[52] The heavier hafnium congener [Cp*Hf=PPh] 71 was postulated to be formed from the reaction of precursor compound 70 with NaN(SiMe3)2 as base (Scheme 19b), based on its at 376, but it was not isolated or trapped.[53]
32
31P

NMR resonance

Nucleophilic Phosphinidene Complexes: Access and Applicability

Cl a) Cp(*)2Zr Cl

Li2PR'

CpR2Zr

P R'

69 R' = 2,4,6-(MeO) 3C6H2 Cp* Hf P Cp* 71 Ph

PHPh b) Cp*2Hf 70 I

NaN(SiMe3) 2 - HN(SiMe3)2 - NaI

Scheme 19. Formation of transient Zr and Hf phosphinidenes.

Attempts

to

kinetically

stabilize

the

transient

lanthanide

phosphinidene complex 73 that was generated from 72 by applying the bulky phosphine H2PMes* resulted in CH activation and formation of phosphaindole 74 (Scheme 20). The sterically less hindered H2PMes was shown to give isolable lutetium-dimer 75 presumably from dimerization of the putative monomeric complex.[54]

PR 2 N Lu PR2 72

Si H2PMes* Si - 2 SiMe4 N

PR2 Lu PMes* PR 2 73 PR2 Mes P

Bu

Bu

P H

R= i Pr, Ph

74

R2P Lu N

H 2PMes - 2 SiMe4

Lu P PR2 75

Mes R2P

Scheme 20. Postulated lanthanide phosphinidene 73 and isolable Lu-dimer 75.

33

Chapter 1

1.4.2

RP-transfer

Of all the 18-electron phosphinidene complexes, Zr phosphinidene 4 developed by Stephan is the most extensively studied and a variety of phosphinidene transfer reactions has been developed.[7] The phospha-Wittig reaction that transfers a PR group is the most applied reaction of nucleophilic phosphinidene complexes bearing oxo- or halophilic transition metals such as Zr (Scheme 21). Stephan and coworkers demonstrated that the reaction of zirconium complex 4 with ketones and aldehydes yields phosphaalkenes 76 and the insoluble zirconocene oxide [Cp2ZrO]n, which is easily separated from the product, together with uncoordinated PMe3 (Scheme 21a).[20b] This metathesis reaction is thought to proceed via initial decoordination of PMe3 generating the active 16-electron species [Cp2Zr=PMes*]. Subsequent coordination of the carbonyl species to Zr followed by intramolecular attack of the nucleophilic phosphorus atom gives a 4membered intermediate (Scheme 21a), which via retrocyclisation yields the P=C and Zr=O products. Phosphinidene 4 also undergoes a metathesis reaction with phenylisothiocyanate to give heteroallene EPhN=C=PMes* 77 and the insoluble zirconocene sulfide dimer [Cp2Zr(-S)]2 (Scheme 21b).[20b] In addition, epoxides can be converted into the three-membered phosphiranes 78 via a P/Oexchange (Scheme 21c),[20b] whereas 4 in the presence of gem dihalides and CHCl3 affords phosphaalkene 76b (Scheme 21d).[20b] This approach was successfully extended to the synthesis of phosphirene 79, phospholane 80, and the substituted phosphirane 81 (Scheme 21e,f,g).[20b] The mechanism for formation of phosphirane 79 from Zr-phosphinidene 4 and 1,2-dichloroethane, invoking the 16electron [Cp2Zr=PMes*] was also addressed computationally.[55]

34

Nucleophilic Phosphinidene Complexes: Access and Applicability

Mes* O
a) 4 Cp Zr Cp

P O Mes*

R' R'
- [Cp2Zr=O]n R'

Mes* R'

R'

R'

- PMe3

R' = H, Me, Ph 4 b) Cp Zr Cp 4 - [Cp2Zr=O]n - PMe3 4 - Cp2ZrCl2 - PMe3 4 - Cp2ZrCl2 - PMe3 4 f) - Cp2ZrCl2 - PMe3

76a

Mes*
- [Cp2Zr( -S)]2

PhN=C=S

P N S Ph R P CHPh C 78 HPh Mes* H


76b

N=C=P Ph
77

- PMe3

c)

O CHPh C HPh

d)

CHECl2 E = H, Cl

P E

e)

Cl Cl

Mes* P
79

Cl

Cl

P Mes*
80

Cl
g)

Cl

4 - Cp2ZrCl2 - PMe3

Mes* P
81

Scheme 21. Phosphinidene transfer reactions of phosphinidene 4.

Schrocks group reported that tantalum complex 12 reacts with carbonyl compounds to yield phosphaalkene 76c and Ta=O complex 82 (Scheme 22).[26] As Schrock also showed in earlier work[56] that Ta alkylidenes and carbonyl compounds yield alkenes and Ta=O species, the P/C analogy between phosphinidene and carbene complexes is demonstrated.

35

Chapter 1

Scheme 22. Phosphinidene transfer reaction of tantalum complex 12.

Lammertsma and co-workers showed that the rate of the reaction of phosphinidenes 10 (M = Co (10c), Rh (10d), Ir (10a)) with dihalomethanes to afford phosphaalkene 76d (Scheme 23a)

depends on the halogen atom of the substrate, the oxygen/halogen philicity of the transition metal, and the electronic properties of the ancillary ligand.[25a,38,42,43] The influence of the stabilizing ligand was demonstrated by changing the phosphane donor of complex 48a (L = PPh3) for a NHC carbene ligand in 50a (IiPr2Me2), which accelerates (40 times) the formation H2C=PMes* 76e (Scheme 23b).[42] It was demonstrated that for 50a, the relative -donor/-acceptor ability of the NHC ligands can easily be influenced by a simple substituentcontrolled conformational change.[42]

Scheme 23. Phosphinidene transfer reactions of late transition metal phosphinidenes.

36

Nucleophilic Phosphinidene Complexes: Access and Applicability

1.4.3

Insertion into the M=P bond =

The groups of Stephan and Mindiola reported on the insertion of substrate molecules into the M=P bond. For example, whereas reaction of Zr-complex 4 with PhCN afforded E/Z imido complex 84 in a 1:1 ratio (Scheme 24),[20b] that with dicyclohexylcarbodiimine gave insertion into the Zr=PMes* bond to yield the X-ray crystallographically characterized phosphaguanidino complex 85 (Scheme 24).[20b]
Cy Ph PMes* Cp2Zr N Me3P E/Z-84 PhC N Zr P 4
Scheme 24. Zr=P bond insertion reactions.

Mes* CyN=C=NCy - PMe3

N Cp2 Zr 85 PMes* N Cy

PMe3

The coordinatively unsaturated titanium phosphinidene complex 35a reacts with tBuNC affording the rare 2-(N,C)-phosphaazaallene complex 86 (Scheme 25).[34b] The two 31P NMR resonances at 8.5 and 17.6 ppm for 86 indicate the presence of two isomers in solution. Reaction of 35a with N2CPh2 yielded complex 87, which contains an uncommon phosphinylimide Ti-ligand.[34b] Both complexes (86 and 87) are exceedingly reactive and readily decompose in solution and in the solid state.
t Bu

Ar N Ar

N Ti N

CCH 2tBu t BuNC C PMes*

t BuNC

Ar H 2 N C N2CPh2 Ti N P Mes* Ar 35a

86

t Bu

Ar N N CPh2 Ti N N PMes* Ar CH2t Bu 87


Sc

Ar = 2,6-iPr2C6H 3
heme 25. Insertion reactions into the Ti=P bond.

37

Chapter 1

1.4.4

Cycloaddition to the M=P bond =

[1+2]- and [2+2]-(retro)cycloadditions are important metal-assisted bond forming and bond breaking reactions that are well established for metal alkylidenes,[57] in contrast to the nucleophilic phosphinidene complexes. Only few examples have been reported, the stepwise addition of isocyanides being one. In-situ-generated 16-electron complex [(Cp*)Ir=PR] (R = Mes, Mes*, Dmp) was shown to react with an isocyanide to form 18-electron phosphinidene complex

(Cp*)(XyNC)Ir=PR 10e. Subsequently reaction with ArNC (Ar = Ph, Xy) gave complex 89, presumably via intermediate 88 as indicated by DFT calculations (Scheme 26).[49]
5-Cp*
Ir PR XyNC 10e ArN C

5-Cp*

5-Cp*
Ir P XyNC Ar C R 89 N Ar

Ir P C XyNC 88 N

R = Mes, Mes*, Dmp; Ar = Ph, Xy


Scheme 26. [1+2]-cycloaddition of phosphinidene and isocyanide.

An example of the [2+2]-cycloaddition of C=C and CC multiple bonds to metal phosphinidenes was provided by Stephan et. al.. Zirconium complex 4 reversibly adds to acetylenes to afford phosphametallacycle 90, which has an indicative 31P NMR resonance at 55 ppm (Scheme 27a).[58] Loss of PMe3 from 4 is the rate determining step in this reaction. It was shown that a more expeditious version of this reaction starts with the spontaneous loss of methane from Cp2(Me)ZrPHMes*.[58] Hillhouses group showed that nickel phosphinidene 42 reacts with olefins to form phosphirane cyclo-C2H4PDmp in a stereoselective manner by way of metallacyclobutane 91 (Scheme 27b).[59]
38

Nucleophilic Phosphinidene Complexes: Access and Applicability

Mes* Mes*
a)

Zr P PMe3
4

- PMe3

PhC CR
[Cp2Zr=PMes*] R = Me, Ph 90

P Cp2Zr C R C Ph

R R P
b)

Dmp Dmp C2H4 P (dtbpe)Ni


91

Ni P P RR
42

C2H4 - (dtbpe)Ni(C2H4)

Dmp P

Dmp
c) 42

Ph Ph (dtbpe)Ni
93

PhC CH

P (dtbpe)Ni
92

P Dmp H
t

Bu

Mes*
d)
6

Bu P (Ar)Ru
95 R

-Ar Ru P Mes*
51a

RC CR'

P (Ar)Ru
94 R

R'

R'

Ar = C6H6, p-Cymene; R = Me, Ph, SiMe3; R' = H, Me


Scheme 27. [2+2] cycloaddition reactions.

In addition, 42 was shown to undergo cycloaddition with alkynes to give the putative [2+2]-adduct phosphametallacyclobutene 92 that rearranges to the more stable metallophosphabicyclobutane 93 (Scheme 27c).[60] In-situ-generated ruthenium phosphinidenes 51a also react with alkynes, as shown by Lammertsmas group,[39] to give the stable phosphaallyl complexes 95 (Scheme 27d). It was reasoned that first [2+2]-cycloadduct 94 is formed, which subsequently undergoes CH activation to yield the final product. Analogously, Menye-Biyogo et al. have reported the formation of the putative phosphinidenes 51a from the interaction of phosphinidene complex
39

Chapter 1

(6-pCy)(Cy3P)Ru=PMes* and alkynes by loss of the phosphine ligand.[61] Zwitterionic titanium phosphinidene 18 with its labile borate group was shown to undergo [2+2]-cycloaddition with diphenylacetylene to generate phosphatitanocyclobutene 96 (Scheme 28a),[28] based on its characteristic
31P

(160.7) and

13C

NMR (253.5) chemical shifts.

Complex 18 was demonstrated to be able to function as a precatalyst in the catalytic hydrophosphination of PhCCPh with PhPH2. The proposed mechanism (Scheme 28b) involves PAr transfer of the primary phosphine followed by [2+2]-cycloaddition of diphenylacetylene to form 97, which generates vinylphosphine HP(Ph)Ph=C=CHPh 98 upon reaction with phenylphosphine.[28]
B(C6F5)3 Ar H C N H H Ti N P Is Ar 18 Ph b) 18 PH2Ph - PH 2Is [Ti] PPh PhCCPh [Ti] P Ph 97 Ph PhHP 98 H
Scheme 28. [2+2]-cycloaddition reaction (a) and proposed catalytic PAr transfer in hydrophosphination reaction (b) for Ti phosphinidene. The -diketeminate ligand and the BCH3(C6F5)3 anion in (b) are omitted for clarity.

a)

PhCCPh

Ar Ph H3C B(C6F5)3 N Ph Ti N P Ar Is 96

Ph

PH2Ph Ph

A diphosphorus analogue of the versatile Dtz intermediate that is common in the chemistry of complexed carbenes has been reported
40

Nucleophilic Phosphinidene Complexes: Access and Applicability

by the group of Lammertsma. 3-Diphosphavinylcarbene complex 100 resulted on the DBU-induced reaction of Ir and Ru complexed primary phosphines 46a and 47a with phosphaalkyne Mes*CP. The product obtained with the less congested tBuCP was shown to convert to the 1,3-diphospha-3H-indene complex 101, which

resembles the intermediate of the Dtz benzannulation reaction (Scheme 29).[62] The reversibility of the phosphaalkyne addition was demonstrated by the exchange of Mes*CP in 100b with PPh3 and tBuCP, yielding phosphinidene complex 48a and 101b, respectively.
L Cl Cl Mes* M PH2 Mes*
DBU - DBUHCl

LM PH Cl 99a ,b

RC P, DBU - DBUHCl

LM = 46a, 5 -Cp*Ir 47a, 6 -pCyRu;

R = Mes*, tBu

tBu t Bu LM P P R
100a,b R = Mes*

t Bu LM t Bu
for R = tBu

P P tBu t Bu

101a,b R = tBu

Scheme 29. Synthesis and rearrangement of 3-diphosphavinylcarbene 100.

1.5

Conclusion

The emerging applicability of terminal phosphinidene complexes that are nucleophilic at the phosphorus atom drives the search for novel reagents and new reactions. The past decades has shown many openings to evolve this chemistry. Whereas the focus was initially on stable 18-electron complexes it is evident that the in-situ generated 16-electron analogous are viable reactive intermediates. Much ground still needs to be covered, but it is clear that the broad spectrum of reactions that are common place for transition metal41

Chapter 1

complexed carbenes are also feasible for the phosphinidenes, such as [2+2]-cycloadditions to CC multiple bonds, insertions into single bonds, and phospha-Wittig type reactions. Moreover, the diagonal relationship between phosphorus and carbon in the Periodic Table provides opportunities to mimic the carbene complexes by

conducting mechanistic studies that take advantage of the stabilizing phosphorus atom. Exemplary is the diphospha-Dtz intermediate. There are many more possibilities. With the increasing emphasis on the element phosphorus in conducting organic conversions, in ligands and catalysts, and in the advance of metalassisted organophosphorus chemistry much can be expected from this field.

1.6

Outline of this thesis

In Chapter 2 the synthesis and reactivity of NHC-functionalized Ru phosphinidene (6-Bz)(IiPr2Me2)Ru=PMes* 50a is described.[42] The influence of the stabilizing ligand (PPh3 vs NHC) and the relative donor/-acceptor ability of these ligands bonded to the nucleophilic phosphinidene complexes are investigated computationally. The scope of the dehydrohalogenationligation sequence using NHCs both as Brnsted base and as stabilizing ligand has been successfully extended in Chapter 3 by the synthesis of novel ruthenium, osmium, and rhodium phosphinidene complexes.[43] An extensive computational analysis of the Group 7-9 transition metal complexed phosphinidenes revealed Re, Rh, and Ru as the most reactive transition metals of the Group 7-9 triads. In Chapter 4 in-situ-generated 16-electron complex [(Cp*)Ir=PR] (R = Mes, Mes*, Dmp) were reacted with isocyanide ArNC yielding isolable phosphinidene complexes (Cp*)(ArNC)Ir=PR, which are
42

Nucleophilic Phosphinidene Complexes: Access and Applicability

prone to [1+2] cycloaddition with a second isocyanide ArNC to afford novel iridaphosphirane complexes 89 [Cp*(ArNC) IrPArC =NAr].[49] For the imido analogue [(Cp*)IrNtBu] of Bergman a different mechanism is found with isocyanides. In Chapter 5 the synthesis, mechanism, and reactivity of the novel

3-diphosphavinylcarbene complex 100, a diphosphorus analogue of


the versatile Dtz intermediate, is presented.[62] The product obtained with the less congested tBuCP was shown to convert via an unprecedented rearrangement to the novel 1,3-diphospha-3Hindene complex 101.

References and Notes


[1] (a) K. Lammertsma, Top. Curr. Chem. 2003, 237, 95119. (b) J. C. Slootweg, K. Lammertsma, In Science of Synthesis; Trost, B. M., Mathey, F., Eds.; Georg Thieme Verlag: Stuttgart, 2009; Vol. 42, pp 1536. [2] G. P. Moss, P. A. S. Smith, D. Travenier, Pure Appl. Chem. 1995, 67, 13071375. [3] Isolation of the first carbene complex, (a) E. O. Fischer, A. Maasbl, Angew. Chem. 1964, 76, 645645; Angew. Chem. Int. Ed. 1964, 3, 580581. Isolation of the first free carbene, which can be classified as a push-push carbene, see: (b) A. J. Arduengo, III, R. L. Harlow, M. Kline, J. Am. Chem. Soc. 1991, 113, 361363. For push-pull carbenes, see: (c) A. Igau, H. Grtzmacher, A. Baceiredo, G. Bertrand, J. Am. Chem. Soc. 1988, 110, 64636466. (d) A. Igau, A. Baceiredo, G. Trinquier, G. Bertrand, Angew. Chem. 1989, 101, 617618; Angew. Chem., Int. Ed. Engl. 1989, 28, 621622. For reviews, see: (e) G. Bertrand, Carbene Chemistry (Marcel Dekker, 2002). (f) D. Bourissou, O. Guerret, F. P. Gabba, G. Bertrand, Chem. Rev. 2000, 100, 3992. (g) F. E. Hahn, M. C. Jahnke, Angew. Chem. Int. Ed. 2008, 47, 3122 3172. [4] (a) W. A. Nugent, B. L. Haymore, Coord. Chem. Rev. 1980, 31, 123175. (b) J. K. Brask, T. Chivers, Angew. Chem. 2001, 113, 40824098; Angew. Chem. Int. Ed. 2001, 40, 39603976. (c) P. R. Sharp, J. Chem. Soc., Dalton Trans. 2000, 26472657. (d) D. E. Wigley, Prog. Inorg. Chem. 1994, 42, 239482. (e) L. H. Gade, P. Mountford, Coord. Chem. Rev. 2001, 216217, 6597.

43

Chapter 1

[5] (a) X. Li, S. I. Weissman, T.-S. Lin, P. P. Gaspar, A. H. Cowley, A. I. Smirnov, J. Am. Chem. Soc. 1994, 116, 78997900. (b) G. Bucher, M. L. G. Borst, A. W. Ehlers, K. Lammertsma, S. Ceola, M. Huber, D. Grote, W. Sander, Angew. Chem. 2005, 117, 33533356; Angew. Chem. Int. Ed. Engl. 2005, 44, 32893373. (c) J. Glatthaar, G. Maier, Angew. Chem. 2004, 16, 13141317; Angew. Chem. Int. Ed. 2004, 43, 1294 1296. (d) J. J. Harrison, B. E. Williamson, J. Phys. Chem. A 2005, 109, 13431347. [6] U. Schmidt, Angew. Chem. 1975, 87, 535540; Angew. Chem. Int. Ed. Engl. 1975, 14, 523528. [7] D. W. Stephan, Angew. Chem. 2000, 112, 322338; Angew. Chem. Int. Ed. 2000, 39, 314337. [8] (a) K. Lammertsma, M. J. M. Vlaar, Eur. J. Org. Chem. 2002, 11271138. (b) F. Mathey, N. H. Tran Huy, A. Marinetti, Helv. Chim. Acta. 2001 84, 29382957. [9] L. Weber, Eur. J. Inorg. Chem. 2007, 40954117. [10] F. Mathey, Dalton Trans. 2007, 18611868. [11] a) A. Marinetti, F. Mathey, J. Fischer, A. Mitschler. J. Am. Chem. Soc. 1982, 104, 44844485. (b) A. Marinetti, F. Mathey, J. Fischer, A. Mitschler, J. Chem. Soc., Chem. Commun. 1982, 667668. [12] (a) P. B. Hitchcock, M. F. Lappert, W.-P. Leung, J. Chem. Soc., Chem. Commun. 1987, 12821283. (b) R. Bohra, P. B. Hitchcock, M. F. Lappert, W.-P. Leung, Polyhedron, 1989, 8, 1884. [13] (a) A. H. Cowley, A. R. Barron, Acc. Chem. Res. 1988, 21, 8187. (b) A. H. Cowley, Acc. Chem. Res. 1997, 30, 445451. [14] The first reported Schrock complexes: (a) R. R. Schrock, J. Am. Chem. Soc. 1974, 96, 67966797. (b) R. R. Schrock, J. Am. Chem. Soc. 1978, 100, 33593370. Selected review: (c) R. R. Schrock, Acc. Chem. Res. 1979, 12, 98104. [15] Selected reviews: (a) E. O. Fischer, G. Kreis, C. G. Kreiter, J. Muller, G. Huttner, H. Lorenz, Angew. Chem. 1973, 85, 618620; Angew. Chem., Int. Ed. Engl. 1973, 12, 564565. (b) Transition Metal Carbene Complexes, ed. K. H. Dtz, H. Fischer, P. Hoffmann, F. R. Kreissl, U. Schubert, K. Weiss, VCH, Weinheim, 1983. (c) K. H. Dtz, Angew. Chem. 1984, 96, 573594; Angew. Chem., Int. Ed. Engl. 1984, 23, 587608. (d) L. S. Hegedus, in Comprehensive Organometallic Chemistry II, ed. F. W. Abel, F. G. A. Stone and G. Wilkinson, Pergamon, Oxford, 1995, 12, p. 549. (e) W. D. Wulff, in Comprehensive Organometallic Chemistry II, ed. F. W. Abel, F. G. A. Stone and G. Wilkinson, Pergamon, Oxford, 1995, 12, 469.

44

Nucleophilic Phosphinidene Complexes: Access and Applicability

[16] A. W. Ehlers, E. J. Baerends, K. Lammertsma, J. Am. Chem. Soc. 2002, 124, 2831 2838. [17] (a) B. T. Sterenberg, K. A. Udachin, A. J. Carty, Organometallics 2003, 22, 3927 3932. (b) B. T. Sterenberg, A. J. Carty, J. Organomet. Chem. 2001, 617618, 696 701. (c) B. T. Sterenberg, K. A. Udachin, A. J. Carty, Organometallics 2001, 20, 26572659. (d) B. T. Sterenberg, K. A. Udachin, A. J. Carty, Organometallics 2001, 20, 44634465. (e) J. Snchez-Nieves, B. T. Sterenberg, K. A. Udachin, A. J. Carty, J. Am. Chem. Soc. 2003, 125, 24042405. (f) T. W. Graham, R. P.-Y. Cariou, J. Snchez-Nieves, A. E. Allen, K. A. Udachin, R. Regragui, A. J. Carty, Organometallics 2005, 24, 20232026. [18] (a) T. W. Graham, K. A. Udachin, A. J. Carty, Chem. Commun. 2005, 58905892. (b) B. T. Sterenberg, O. S. Senturk, K. A. Udachin, A. J. Carty, Organometallics 007, 26, 925937. [19] (a) Z. Hou, T. L. Breen, D. W. Stephan, Organometallics 1993, 12, 31583167. (b) Z. Hou, D. W. Stephan, J. Am. Chem. Soc. 1992, 114, 1008810089. (c) J. Ho, Z. Hou, J. Drake, D. W. Stephan, Organometallics 1993, 12, 31453157. (d) J. Ho, D. W. Stephan, Organometallics 1991, 10, 30013003. [20] (a) J. Ho, R. Rousseau, D. W. Stephan, Organometallics 1994, 13, 19181926. (b) T. L. Breen, D. W. Stephan, J. Am. Chem. Soc. 1995, 117, 1191411921. [21] E. Urnezius, K.-C. Lam, A. L. Rheingold, J. D. Protasiewicz, J. Organomet. Chem. 2001, 630, 193197. [22] A. T. Termaten, PhD thesis, Vrije Universiteit Amsterdam, the Netherlands, 2004. [23] J. Pikies, E. Baum, E. Matern, J. Chojnacki, R. Grubba, A. Robaszkiewicz, Chem. Commun. 2004, 24782479. [24]D. S. J. Arney, R. C. Schnabel, B. C. Scott, C. J. Burns, J. Am. Chem. Soc, 1996, 118, 67806781. [25] (a) A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2002, 21, 31963202. (b) A. T. Termaten, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 35773582. [26] C. C. Cummins, R. R. Schrock, W. M. Davis, Angew. Chem. 1993, 105, 758761; Angew. Chem. Int. Ed., 1993, 32, 756759. [27] J. S. Freundlich, R. R. Schrock, W. M. Davis, J. Am. Chem. Soc. 1996, 118, 3643 3655. [28] G. Zhao, F. Basuli, U. J. Kilgore, H. Fan, H. Aneetha, J. C. Huffman, G. Wu, D. J. Mindiola, J. Am. Chem. Soc. 2006, 128, 1357513585.

45

Chapter 1

[29] (a) J. S. Figueroa, C. C. Cummins. J. Am. Chem. Soc. 2003, 125, 40204021. (b) J. S. Figueroa, C. C. Cummins, Angew. Chem. 2004, 116, 10021006; Angew. Chem. Int. Ed. 2004, 43, 984988. [30] A. H. Cowley, B. Pellerin, J. L. Atwood, S. G. Bott, J. Am. Chem. Soc. 1990, 112, 67346735. [31] J. B. Bonanno, P. T. Wolczanski, E. B. Lobkovsky, J. Am. Chem. Soc. 1994, 116, 1115911160. [32] J. D. Masuda, A. J. Hoskin, T. W. Graham, C. Beddie, M. C. Fermin, N. Etkin, D. W. Stephan, Chem. Eur. J. 2006, 12, 86968707. [33] E. Niecke, J. Hein, M. Nieger, Organometallics 1989, 8, 23702371. [34](a) F. Basuli, J. Tomaszewski, J. C. Huffman, D. J. Mindiola, J. Am. Chem. Soc. 2003, 125, 1017010171. (b) F. Basuli, L. A. Watson, J. C. Huffman, D. J. Mindiola, Dalton Trans. 2003, 42284229. (c) F. Basuli, B. C. Bailey, J. C. Huffman, M.-H. Baik, D. J. Mindiola, J. Am. Chem. Soc. 2004, 126, 19241925. (d) B. C. Bailey, J. C. Huffman, D. J. Mindiola, W. Weng, O. V. Ozerov, Organometallics 2005, 24, 13901393. [35] W. Malisch, U.-A. Hirth, K. Grn, M. Schmeusser, O. Fey, U. Weis, Angew. Chem. 1995, 107, 27172719; Angew. Chem. Int. Ed. Engl. 1995, 34, 25002502. [36] R. Melenkivitz, D. J. Mindiola, G. L. Hillhouse, J. Am. Chem. Soc. 2002, 124, 3846 3847. [37] U. J. Kilgore, H. Fan, M. Pink, E. Urnezius, J. D. Protasiewicz, D. J. Mindiola, Chem. Commun. 2009, 45214523. [38] A. T. Termaten, H. Aktas, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2003, 22, 18271834. [39] A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 22002208. [40]R. Menye-Biyogo, F. Delpech, A. Castel, A. Castel, H. Gornitzka, P. Rivire, Angew. Chem. 2003, 115, 57685770; Angew. Chem. Int. Ed. 2003, 42, 56105612. [41] N. Kuhn, T. Kratz, Synthesis 1993, 561562. [42] H. Aktas, J. C. Slootweg, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, J. Am. Chem. Soc. 2009, 131, 66666667. [43] H. Aktas, J. C. Slootweg, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2009, 28, 51665172. [44] R. W. Alder, P. R. Allen, S. J. Williams, J. Chem. Soc., Chem. Comm. 1995, 1267 1268.

46

Nucleophilic Phosphinidene Complexes: Access and Applicability

[45] The pKa of DBU is 11.3, see: (a) G. Hfle, W. Steglich, H. Vorbruggen, Angew. Chem. 1978, 90, 602615; Angew. Chem. Int. Ed. 1978, 17, 569583. The estimated pKa value of DBU is 12 (H2O), see: (b) H. Ripin, D. A. Evans.

http://daecr1.harvard.edu/pdf/evans_pKa_table.pdf. [46] (a) V. Nair, S. Bindu, V. Sreekumar, Angew. Chem. 2004, 116, 52405245; Angew. Chem. Int. Ed. 2004, 43, 51305135. (b) A. E. Mattson, A. R. Bharadwaj, A. M. Zuhl, K. A. Scheidt, J. Org. Chem. 2006, 71, 57155724; and references therein. [47] A. J. Arduengo, III, H. V. R. Dias, R. L. Harlov, M. Kline, J. Am. Chem. Soc. 1992, 114, 55305534. [48]Rh: (a) W. A. Herrmann, C. Kcher, L. J. Gooen, G. R. J. Artus, Chem. Eur. J. 1996, 2, 16271636. (b) W. A. Herrmann, M. Elison, J. Fischer, C. Kcher, G. R. J. Artus, Chem. Eur. J. 1996, 2, 772780. (c) N. M. Scott, R. Dorta, E. D. Stevens, A. Correa, L. Cavallo, S. P. Nolan, J. Am. Chem. Soc. 2005, 127, 35163526. (d) M. Viciano, E. Mas-Marza, M. Sanau, E. Peris, Organometallics 2006, 25, 30633069. Ru: (a); (b); (e) J. Huang, H.-J. Schanz, E. D. Stevens, S. P. Nolan, Organometallics 1999, 18, 23702370. (f) W. Baratta, E. Herdtweck, W. A. Herrmann, P. Rigo, J. Schwarz, Organometallics 2002, 21, 21012106. [49] H. Aktas, J. Mulder, J. C. Slootweg, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, J. Am. Chem. Soc. 2009, 131, 1353113537. [50] U. Segerer, R. Felsberg, S. Blaurock, G. A. Hadi, E. HeyHawkins, Phosphorus Sulfur Silicon, 1999, 146, 477480. [51] A. T. Termaten, T. Nijbacker, Andreas W. Ehlers, M. Schakel, M. Lutz, A. L. Spek, M. L. McKee, K. Lammertsma, Chem. Eur. J. 2004, 10, 40634072. [52] A. Mahieu, A. Igau, J.P. Majoral, Phosphorus Sulfur Silicon 1995, 104, 235239. [53] G. A. Vaughan, G. L. Hillhouse, Organometallics 1989, 8, 17601765. [54] J. D. Masuda, K. C. Jantunen, O. V. Ozerov, K. J. T. Noonan, D. P. Gates, B. L. Scott, J. L. Kiplinger, J. Am. Chem. Soc. 2008, 130, 24082409. [55] T. P. M. Goumans, A. W. Ehlers, K. Lammertsma, J. Organometallic Chem. 2005, 690, 55175524. [56] R. R. Schrock, J. Am. Chem. Soc. 1976, 98, 53995400. [57] R. H. Grubbs, Angew. Chem. 2006, 118, 38453850; Angew. Chem. Int. Ed. 2006, 45, 37603765. [58] T. L. Breen, D. W. Stephan, J. Am. Chem. Soc. 1996, 118, 42044205. [59] R. Waterman, G. L. Hillhouse, J. Am. Chem. Soc. 2003, 125, 1335013351. [60] R. Waterman, G. L. Hillhouse, Organometallics 2003, 22, 51825184.

47

Chapter 1

[61] R. Menye-Biyogo, F. Delpech, A. Castel, V. Pimienta, H. Gornitzka, P. Rivire, Organometallics 2007, 26, 50915101. [62] H. Aktas, J. C. Slootweg, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Angew. Chem. 2009, 121, 31543157; Angew. Chem. Int. Ed. 2009, 48, 31083111.

48

Chapter 2
N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes
Halil Aktas, J. Chris Slootweg, Marius Schakel, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, and Koop Lammertsma*,

Department of Chemistry and Pharmaceutical Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands, and Bijvoet Center for Biomolecular Research, Crystal Structural Chemistry, Utrecht University, Padualaan 8, 3584 CH Utrecht, The Netherlands

J. Am. Chem. Soc. 2009, 131, 66666667

Abstract: Catalyst tuning by changing ligands is a well established protocol in transition metal chemistry. N-Heterocyclic carbenes (NHC) and tertiary phosphines (R3P) are the ubiquitous ligand actors. Here we demonstrate that the relative -donor/-acceptor ability of the NHC ligand itself can be influenced by a simple substituent-controlled conformational change, thereby directly impacting the reactivity of the transition metal complex.

Chapter 2

2.1

Introduction

N-heterocyclic carbenes (NHC)[1] are ubiquitous ligands in transitionmetal chemistry and homogeneous catalysis and serve increasingly often as a replacement for tertiary phosphines (R3P). The two ligand classes exert often subtle but crucially different electronic influences on the properties of catalysts.[2] Exemplary is the enhanced activity of the second-generation Grubbs metathesis catalyst

[(Cy3P)(L)Cl2Ru=CHPh] (Cy = cyclohexyl; L = H2IMes, 1,3-dimesityl-4,5dihydroimidazol-2-ylidene) relative to that of the first-generation catalyst (L = Cy3P), which is caused by the differences in -donor/acceptor ability, [3] shape, and symmetry of the ligands.[4] Does a similar sensitivity apply to the isolobal phosphinidene[5] complexes? We address here the ligand and conformational sensitivities for [(6-C6H6)(L)Ru=PMes*] (L = IiPr2Me2; Mes* = 2,4,6tBu3C6H2 (1), L = Ph3P (2)[6]) by examining their solution-phase chemistry together with their structure-activity parameters modeled by density functional theory (DFT). We simultaneously demonstrate the applicability of phosphinidene complexes to the synthesis of phosphaalkenes (P=C),[5a] which are unique P=ligands[7] and

attractive building blocks for P-functionalized polymers.[8]

2.2

NHC functionalization

The desired novel dark-green crystalline compound 1 (84%) was obtained by a double dehydrohalogenationligation sequence[9] of the phosphine complex [(6-C6H6)RuCl2(PH2Mes*)][6] using three equivalents of IiPr2Me2 in toluene (eq 1). In this reaction, two NHCs act as Brnsted bases, while the third carbene captures the putative 16electron intermediate [(6-C6H6)Ru=PMes*] (3). The single
31P

NMR

resonance of 1 at 751.7 ppm is highly shielded compared to the


50

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

known triphenylphosphine analogue 2 (31P NMR = 845.9 ppm),[6] which is attributed to the -donor capacity of IiPr2Me2 (Table 1).
Mes* Cl Cl Ru PH2
2 Ii Pr2Me2 I iPr2Me 2

Mes* Ru P L (1)

Ru P Mes* 3

-2 Ii Pr2Me2HCl

1 L = IiPr 2Me2

The molecular structure of 1, established unequivocally by a singlecrystal X-ray analysis (Figure 1),[10] has an exact mirror symmetry and shows a two-legged piano stool shape with a characteristic acute C15Ru1P1 angle of 84.88(9), a bent phosphinidene complex with a Ru1P1C1 angle of 105.81(9), and an E configuration for the congested Ru1P1 double bond (2.2222(8) ). This bond is longer than that of the first-generation phosphinidene 2 (2.1988(6) ),[6] whereas its Ru1Bz(cg) bond is correspondingly shorter (1.7390(12) in 1; 1.7560(12)[6] in 2).

Figure 1. Displacement ellipsoid plot (50% probability level) for 1. Only one conformation of the disordered tert-butyl group is shown. Hydrogen atoms are omitted for clarity. Bz denotes the centroid position of the benzene ring. Symmetry operation a: x, 0.5-y, z. Selected bond distances [] and angles and torsion angles []: Ru1P1 2.2222(8), Ru1C15 2.091(3), Ru1Bz(cg) 1.7390(12), P1C1 1.876(3); Ru1 P1C1 105.81(9), P1Ru1C15 84.88(9), N1C15N1a 104.8(3); C2aC1C2C3 18.7(4).

51

Chapter 2

Steric congestion is reflected in the 18.7(4) distortion from planarity of the Mes* ring and in the restricted rotation of the iso-propyl wingtip groups of the NHC fragment, indicated by the two
13C

NMR

resonances at 21.6 and 21.8 ppm. A striking feature is the orthogonal relationship between the NHC and Ru=P units, which contrasts with the in-plane arrangement of the NHC and Ru=C units in the secondgeneration Grubbs catalyst.[11]

2.3

Theoretical calculations on [(Bz)M(L)=PPh] =

To address the effect of the NHC ligand orientation in 1 and the impact of the stabilizing ligand (L = NHC vs R3P) on the properties of 1 and 2, we performed BP86/TZP calculations on model structures (labeled 1', 2', etc.) bearing a P-phenyl group (instead of PMes*) and methyl groups on the NHC (IMe) and phosphine (PMe3) ligands (Figure 2).

C1 Ru P1 C1 Ru C2 P1 C2 P2 Ru P1 C2

1-

1-

Figure 2. BP86/TZP optimized structures of 1-, 1-, and 2 (all Cs symmetry). Selected bond distances [] and angles [] of 1-: RuP1 2.216, RuC1 2.071, RuBz(cg) 1.784, P1C2 1.859; RuP1C2 108.74, C1RuP1 82.83; 1-: RuP1 2.204, RuC1 2.159, P1C2 1.862, RuBz(cg) 1.833; RuP1C2 112.39, C1RuP1 91.64; 2: RuP1 2.209, RuP2 2.342, RuBz(cg) 1.791, P1C2 1.859; RuP1C2 109.33, P1RuP2 85.50.

The optimized geometries of 1'-, in which the IMe ligand and the Ru=P bond are orthogonal, and 2' compare well with the corresponding X-ray structures.[6] But why does the NHC-ligated structure not prefer a coplanar arrangement of IMe and Ru=P (1'-)?
52

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

For the unsubstituted NHC (H instead of Me) the calculations do indeed show a 2.5 kcalmol1 preference for the coplanar form, but the methyl derivative favors the orthogonal conformation by 12.5 kcalmol1. Apparently, the steric congestion induced by the IMe wingtip groups enforces the out-of-plane conformation. This substituent effect causes a reduction in the -acceptor capacity of the IMe fragment (0.18 0.10 e-), making the NHC ligand in 1'- an effective donor (Table 1).
Table 1. Energy Decomposition Analysis for RuL bond of 1-, 1-, and 2.a complex 1- 1- 2
a

Ea 60.8 (0.45, 0.10) 45.9 (0.41) 50.6 (0.52, 0.10)

Ea 6.4 (0.01) 13.5 (0.18) 10.8 (0.12)

E steric 17.1 14.4 18.5

E tot 50.1 45.0 43.0

E prep 9.2 16.6 8.1

BE 41.0 28.4 34.9

in kcalmol1; values in parentheses refer to charge transferred (eV) to (positive) and from the metal center.

The RuL bond properties impact those of the Ru=P bond, which is evident from the energy decomposition scheme in ADF (see Appendix 1). Because the carbene provides less back-bonding than PMe3, the frontier orbitals of the [(6-C6H6)(IMe)Ru] fragment (E(dxz) = 2.27 eV, E(dyz) = 2.49 eV) are higher in energy than those of the Ruphosphine fragment (E(dxz) = 2.61 eV, E(dyz) = 2.86 eV; Table 2). Ru=P bond formation causes transfer of charge from [(6-C6H6)(L)Ru] to the 3PPh fragment (E(px) = 4.59 eV, E(py) = 4.86 eV), which is largest for 1'-. Whereas the Ru=P bonds are of similar lengths (2.216 and 2.209 for 1'- and 2' respectively), the polarity varies with the phosphorus atom, which carries more charge in 1'- (0.113e) than in 2' (0.086e).

53

Chapter 2

Table 2. Calculated d-orbital energies (eV) for the [(C6H6)(PhP)Ru] fragments. Ed occ. occ. occ. occ. occ. 1- 4.13 4.09 3.75 2.49 2.27 1- 4.04 3.95 3.66 2.56 2.49 2 4.47 4.44 4.14 2.86 2.61

2.4

Reactivity

The greater Ru=P bond polarity is reflected in the enhanced reactivity of the NHC-containing phosphinidene 1 (L = IiPr2Me2) over that of first-generation 2 (L = Ph3P) toward diiodomethane (eq 2):[12]

31P

NMR monitoring of the reaction of complex 1 showed the

quantitative formation of phosphaalkene H2C=PMes* (6, 94% isolated yield) within 1 min. at 20 C (t (0 C, C6D6) = 22 min; five equivalents of CH2I2). In contrast, the reaction of phosphine ligated complex 2 with CH2I2 is much slower (t (20 C, toluene) = 60 min.; t (0 C, C6D6) = 925 min.) and also less selective (6, 45%). This difference between 1 and 2 demonstrates that like the difference in catalytic activity of the Grubbs catalysts, the reactivity of the isolobal nucleophilic 18electron phosphinidene complexes can also be readily modified by changing the ancillary ligands. The applicability of the illustrated reaction is underscored by the quantitative regeneration of 1 from the transition metal byproduct [(6-C6H6)(IiPr2Me2)RuI2] (4) with DBU and H2PMes*[13] as determined by
54
31P

NMR (63% isolated yield),

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

thereby demonstrating that ruthenium phosphinidene complexes are viable reagents for the synthesis of phosphaalkenes.

2.5

Transient species

A final aspect to address is the presumed 16-electron phosphinidene intermediate 3, which could not be detected by
31P

NMR

spectroscopy,[14,15] suggesting that if it is indeed formed, it is readily captured by IiPr2Me2 to yield 1. Increasing the steric bulk by using 1,3dimesityl-imidazol-2-ylidene (IMes) to slow the NHC complexation enough for detection was unsuccessful, but monitoring its ligation with the less crowded [(6-C6H6)RuCl2(PH2Mes)], which carries a Mes instead of a Mes* substituent, did have the anticipated effect. Besides dark-brown crystalline [(6-C6H6)(IMes)Ru=PMes] (7a;
31P

NMR

= 752.5 ppm, 65%), small amounts of the corresponding toluene adduct [(6-Tol)(IMes)Ru=PMes] (7b; also observed (eqs 3 and 4).[16]
Mes Ru P L 7a L = IMes (3)
31P

NMR = 736.8 ppm, 3%) were

C6 H6 Ru P Mes 8a Me (4) L 7b L = IMes Ru P Mes IMes 8b


benzene toluene slippage

IMes

C 7H 8 Ru P Mes

The apparent arene exchange is supported by detection of d3-7b when d3-toluene was used as the solvent. Since no ligand exchange was observed for the isolated products, it appears that the 16electron intermediate is prone to arene exchange. BP86/TZP calculations support this view. Simplified 16-electron [(6-C6H6)Ru=PH], which has an energy minimum, reacts barrier-free with toluene to
55

Chapter 2

form the 5.6 kcalmol1 favored [(2-Tol)(6-C6H6)Ru=PH] as an initial adduct in the exchange of the two arene ligands. Associative ring slippage[17] via [(4-Tol)(4-C6H6)Ru=PH] then gives [(6-Tol)(2-

C6H6)Ru=PH] (E = 1.4 kcalmol1), which requires 4.2 kcalmol1 to lose benzene and form the product. Implicitly, this process supports a 16electron intermediate that undergoes ligand exchange of aromatic molecules (8a8b) before being captured by the carbene ligand to give 7b.

2.6

Conclusion

Catalyst tuning is generally sought via a change of ligands because their effect is considered to be constant for a given transition metal complex. We have now demonstrated that the relative -donor/acceptor ability of NHC ligands can easily be influenced by a simple substituent-controlled conformational change. The sterically imposed ligand rotation of the NHC fragment in 1 enhances its reactivity and thereby facilitates the synthesis of phosphaalkene (P=C) building blocks.

2.7

Computational Section

All DFT calculations have been performed with the parallelized Amsterdam density functional (ADF) package (version 2005.01 and 2005.01b).[18] The Kohn-Sham MOs were expanded in a large, uncontracted basis set of Slatertype orbitals (STOs), of a triple- basis set with polarization functions quality, corresponding to basis set TZP in the ADF package. The 1s core shell of carbon and nitrogen and the 1s2s2p core shells of phosphorus were treated by the frozen core approximation. The transition metal centers were described by a triple- basis set for the outer ns, np, nd and (n+1)s orbitals, whereas the shells of lower energy were treated by the frozen core approximation using a small core. All calculations were performed at the

56

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

nonlocal exchange self-consistent field (NL-SCF) level, using the local density approximation (LDA) in the Vosko-Wilk-Nusair parametrization[19] with nonlocal corrections for exchange (Becke88)[20] and correlation

(Perdew86).[21] All geometries were optimized using the analytical gradient method implemented by Versluis and Ziegler,[22] including relativistic effects by the Zeroth Order Regular Approximation (ZORA).[23] All Cartesian coordinates in Angstroms and energies in kcal.mol1 for calculated compounds are available free of charge via the Internet at

htttp://pubs.acs.org. Bonding Energy Analysis. The bonding interactions of the transition metal to ligand bonds (RuP, RuC(NHC), and RuBz) were analyzed with ADF's established energy decomposition[24] into an exchange (or Pauli) repulsion (E Pauli) between the electrons on the two fragments plus electrostatic interaction energy part (E elst) and an orbital interaction energy (charge transfer, polarization) part (E oi). The energy necessary to convert fragments from their ground state equilibrium geometries to the geometry and electronic state they acquire in the complex is represented by a preparation energy term (E prep). The overall bond energy (E tot) is formulated as: E tot=E Pauli+E elst+E oi+E prep Note that E tot is defined as the negative of the bond dissociation energy (BDE), i.e., E tot=E (molecule) E (fragments), thereby giving negative values for stable bonds. The orbital interaction term E oi accounts for interactions between occupied orbitals on one fragment with unoccupied orbitals on the other fragment, including HOMO-LUMO interactions and polarization (empty/occupied orbital mixing on the same fragment). The charge transfer part is the result of both -donation from the ligand to the metal, and -back-donation from the metal into the unoccupied orbitals of the ligand. Instead of separating the charge transfer and polarization parts, we used the extended transition state (ETS) method developed by Ziegler and Rauk to decompose E oi into contributions from each irreducible representation of the interacting system.[24] In systems with a clear , -

57

Chapter 2

separation (a' and a), this symmetry partitioning proves to be most informative.

2.8

Experimental Section

General. All experiments and manipulations were performed under an atmosphere of dry nitrogen or argon with rigorous exclusion of air and moisture using flame-dried glassware. 1H,
13C

and

31P

NMR spectra were

recorded at 300 K on a Bruker Avance 250 (respectively 250.13, 62.90 and 101.25 MHz) or on a Bruker Avance 400 (respectively 400.13, 100.64 and 162.06 MHz) spectrometer. 1H NMR spectra were internally referenced to CHCl3 ( 7.26) or C6D5H ( 7.16), 13C NMR spectra to CDCl3 ( 77.16) or C6D6 ( 128.06) and
31P

NMR spectra externally to 85% H3PO4. IR Spectra were

recorded on a Mattson-6030 Galaxy FT-IR spectrophotometer, high-resolution mass spectra (HRMS) on a Finnigan Mat 900 spectrometer operating at an ionization potential of 70 eV, and fast atom bombardment (FAB) mass spectrometry was carried out using a JEOL JMS SX/SX 102A four-sector mass spectrometer. Melting points were measured on samples in unsealed capillaries on a Stuart Scientific SMP3 melting point apparatus and are uncorrected. Elemental analyses were performed by the Microanalytical Laboratory of the Laboratorium fr Organische Chemie, ETH Zrich, Switzerland. Reagents. Solvents were distilled from sodium (toluene and n-hexane), sodium benzophenone (THF), P2O5 (CH2Cl2), or LiAlH4 (pentanes) and kept under an atmosphere of dry nitrogen. Deuterated solvents were dried over 4 molecular sieves (CDCl3, C6D6). All solid starting materials were dried in vacuo. [(6-C6H6)RuCl2]2,[25] [(6-C6Me6)RuCl2]2,[26] (IiPr2Me2),[27] and 1,3-diisopropyl-4,5-

dimethylimidazol-2-ylidene (IMes),[28]

1,3-dimesityl-imidazol-2-ylidene Mes*PH2,[29] were prepared

[(6-C6H6)RuCl2(PH2Mes*)],[6]

according to literature procedures. MesPH2 was prepared, analogously to IsPH2,[30] by LiAlH4 reduction of MesPCl2. [(6-pCy)RuCl2]2 was purchased from Strem and used as received.

58

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

Experimental Procedures. [(6-C6H6)RuCl2(PH2Mes)]. MesPH2 (0.117 g, 0.77 mmol) was added to a dark red suspension of [(6-C6H6)RuCl2]2 (0.175 g, 0.35 mmol) in CH2Cl2 (15 mL), which was stirred for 1 h at room temperature and filtered to remove insoluble material. After concentrating the solution to a few mL, pentanes (30 mL) was added slowly, which resulted in precipitation of brown microcrystals that were isolated by filtration, washed with pentanes (25 mL), and dried in vacuo to yield [(6-C6H6)RuCl2(PH2Mes)] (0.233 g, 0.58 mmol, 83%). Mp 150 C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 2.36 (s, 3 H, pCH3), 2.52 (s, 6 H, o-CH3), 5.44 (s, 6 H, C6H6), 5.59 (d, 1JHP = 395.7 Hz, 2 H, PH2); 7.04 (d, 4JHP = 2.7 Hz, 2 H, m-Mes).
13C{1H}

NMR (62.90 MHz, CDCl3, 300 K):

21.2 (s, p-CH3), 22.2 (d, 3JCP = 8.7 Hz, o-CH3), 87.8 (d, 2JCP = 3.9 Hz, C6H6), 121.9 (d, 1JCP = 44.8 Hz, i-Mes), 129.7 (d, 3JCP = 8.0 Hz, m-Mes), 140.7 (d, 2JCP = 7.6 Hz, o-Mes), 141.4 (d, 4JCP = 2.5 Hz, p-Mes).
31P

NMR (101.3 MHz, CDCl3, 300 K):

44.8 (t, 1JPH = 395.5 Hz, PH2). IR (KBr, cm1): v = 3088.8 (w), 3061.4 (m), 3019.5 (w), 2967.9 (w), 2930.2 (w), 2365.1 (m, PH), 2344.4 (m, PH), 1601.5 (m), 1436.7 (s), 1391.3 (m). HR FAB-MS: calcd for C15H19PCl2Ru: 401.9645; found: 401.9635. MS m/z (%): 402 (30) [M+], 367 (50) [M+ Cl], 331 (65) [M+ 2Cl], 250 (20) [M+ H2PMes], 154 (100) [H2PMes]. [(6-pCy)RuCl2(PH2Mes)]. Analogous to that described for [(6-

C6H6)RuCl2(PH2Mes)], [(6-pCy)RuCl2]2 (0.400 g, 0.653 mmol) and MesPH2 (0.219 g, 1.437 mmol) were used to give [(6-pCy)RuCl2(PH2Mes)] as an orange powder (0.561 g, 1.22 mmol, 94%). Mp 163 C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 1.19 (d, 3JHH = 7.0 Hz, 6 H, CH(CH3)2), 1.99 (s, 3 H, pCyCH3), 2.34 (s, 3 H, p-CH3), 2.50 (s, 6 H, o-CH3), 2.70 (sp, 3JHH = 7.0 Hz, 1 H, CH(CH3)2), 5.13 (d, 3JHH = 5.7 Hz, 2 H, C6H4), 5.34 (d, 3JHH = 5.7 Hz, 2 H, C6H4), 5.52 (d, 1JHP = 389.8 Hz, 2 H, PH2); 7.01 (d, 4JHP = 2.3 Hz, 2 H, m-Mes).
13C{1H}

NMR (62.90 MHz, CDCl3, 300 K): 18.4 (s, pCy-CH3), 21.6 (d, 5JCP = 1.1 Hz, pCH3), 22.5 (s, CH(CH3)2), 22.9 (d, 3JCP = 8.2 Hz, o-CH3), 31.0 (s, CH(CH3)3), 85.5 (d, 2JCP = 5.7 Hz, C6H4), 89.1 (d, 2JCP = 5.3 Hz, C6H4), 97.6 (s, C6H4), 107.7 (d, 2JCP = 1.0 Hz, C6H4), 122.3 (d, 1JCP = 42.3 Hz, i-Mes), 130.1 (d, 3JCP = 7.8 Hz, m-Mes), 141.4 (d, 2JCP = 7.7 Hz, o-Mes), 141.5 (d, 4JCP = 2.9 Hz, p-Mes).
31P

NMR (101.3

59

Chapter 2

MHz, CDCl3, 300 K): 48.0 (t, 1JPH = 389.9 Hz, PH2). IR (KBr, cm1): v = 3043.0 (w), 2967.2 (m), 2913.3 (w), 2864.7 (w), 2387.6 (m, PH), 2353.7 (m, PH), 1601.4 (m), 1437.9 (s), 1382.3 (s). HR FAB-MS: calcd for C19H27PCl2Ru: 458.0271; found: 458.0270. MS m/z (%): 458 (70) [M+], 423 (100) [M+ Cl], 388 (100) [M+ 2Cl], 271 (35) [M+ Cl H2PMes], 154 (90) [H2PMes]. [(6-C6Me6)RuCl2(PH2Mes)]. Analogous to that described for [(6-

C6H6)RuCl2(PH2Mes)], [(6-C6Me6)RuCl2]2 (1.01 g, 1.50 mmol) and MesPH2 (0.464 g, 3.05 mmol) were used to give [(6-C6Me6)RuCl2(PH2Mes)] as a yellow powder (1.29 g, 2.65 mmol, 88%). Mp 257 C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 1.92 (s, 18 H, C6(CH3)6), 2.29 (s, 3 H, p-CH3), 2.39 (s, 6 H, o-CH3), 5.45 (d, 1JHP = 379.7 Hz, 2 H, PH2), 6.95 (d, 4JHP = 2.1 Hz, 2 H, m-Mes).
13C{1H}

NMR (62.90 MHz, CDCl3, 300 K): 15.7 (s, C6(CH3)6), 21.6 (s, p-CH3), 23.4 (d, 3JCP = 8.0 Hz, o-CH3), 96.1 (d, 3JCP = 3.5 Hz, C6(CH3)6), 119.3 (d, 1JCP = 38.3 Hz, i-Mes), 129.9 (d, 3JCP = 7.6 Hz, m-Mes), 141.1 (d, 4JCP = 2.5 Hz, p-Mes), 142.7 (d, 2JCP = 7.4 Hz, o-Mes).
31P

NMR (101.3 MHz, CDCl3, 300 K): 53.7 (t, 1JPH = 379.8 Hz,

PH2). IR (KBr, cm1): v = 3019.1 (w), 2953.2 (m), 2914.4 (m), 2397.3 (m, PH), 2364.6 (m, PH), 1601.3 (m), 1447.67 (s), 1383.7 (m). HR FAB-MS: calcd for C21H31PCl2Ru: 486.0584; found: 486.0588. MS m/z (%): 486 (65) [M+], 451 (50) [M+ Cl], 413 (50) [M+ 2HCl], 299 (60) [M+ 2HCl H2PMes], 154 (100) [H2PMes]. [(6-C6H6)(IiPr2Me2)Ru=PMes*] (1). Complex [(6-C6H6)RuCl2(PH2Mes*)] (0.082 g, 0.154 mmol) was added to a solution of IiPr2Me2 (0.083 g, 0.462 mmol) in toluene (4 mL) at 78 C. The mixture was stirred for 3 h and was slowly allowed to warm up to room temperature. The resulting dark green suspension was filtered and the salts were extracted with toluene (3 x 1 mL). After concentration of the solvent, dark green crystals of 1 were obtained at 20 C (0.083 g, 0.130 mmol, 84%). Mp 222 C (dec). 1H NMR (250.13 MHz, C6D6, 300 K): 1.17 (d, 3JHH = 7.3 Hz, 6 H, CH(CH3)2), 1.32 (d, 3JHH = 7.3 Hz, 6 H, CH(CH3)2), 1.58 (s, 9 H, p-C(CH3)3), 1.78 (s, 18 H, o-C(CH3)3), 1.89 (s, 6 H, CH3), 4.69 (s, 6 H, C6H6), 5.95 (sp, 3JHH = 7.3 Hz, 2 H, CH(CH3)2), 7.60 (s, 2 H, m-Mes*).
13C{1H}

NMR (62.90 MHz, C6D6, 300 K): 10.4 (s, CH3), 21.6 (s, CH(CH3)2), 21.8 (s,

CH(CH3)2), 32.0 (s, p-C(CH3)3), 32.5 (d, 4JCP = 7.5 Hz, o-C(CH3)3), 34.9 (s, p-

60

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

C(CH3)3), 38.6 (s, o-C(CH3)3), 52.3 (s, CH(CH3)2), 79.9 (s, C6H6), 119.4 (s, mMes*), 124.5 (s, =CCH3), 144.6 (s, p-Mes*), 145.6 (s, o-Mes*), 177.0 (d, 1JCP = 100.7 Hz, i-Mes*), 187.9 (s, N2C).
31P{1H}

NMR (101.3 MHz, C6D6, 300 K): 751.7

(s, Ru=P). HRMS (EI, 70 eV): calcd for C35H55N2PRu: 636.3140; found: 636.3137. MS m/z (%): 636 (12) [M+], 558 (8) [M+ C6H6], 536 (8) [M+ CH(CH3)2 C(CH3)3], 319 (20) [M+ Mes* NCH(CH3)2 CH3], 180 (20) [IiPr2Me2]. Anal. Calcd for C35H55N2PRu: C, 66.10; H, 8.72; N, 4.41. Found: C, 66.32; H, 8.83; N, 4.32. Alternatively, 1 can be prepared from [(6-C6H6)(IiPr2Me2)RuI2] (4) as follows: DBU (35.1 L, 0.235 mmol) was added to a red suspension of complex 4 (0.075 g, 0.1223 mmol) and H2PMes* (0.0334 g, 0.1200 mmol) in toluene (20 mL) at room temperature. After 24 h,
31P{1H}

NMR showed full conversion of

H2PMes* into 1. Pentane (5 mL) was added, the resulting dark green suspension was filtered, and after concentration of the solvent, dark green crystals of 1 were obtained at 20 C (0.0471 g, 0.074 mmol, 63%). [(6-C6H6)(IiPr2Me2)RuI2] (4). RuI2-complex 4 was obtained from [(6C6H6)RuCl2]2 and IiPr2Me2 and subsequent anion-exchange using NaI as follows: a freshly distilled suspension of [(6-C6H6)RuCl2]2 (0.275 g, 0.55 mmol) in THF (40 mL) was treated with a solution of IiPr2Me2 (0.194 g, 1.075 mmol) in THF (10 mL) at room temperature. After 3 days, the solvents were evaporated and the dark brown residue was purified by column chromatography (silica, CHCl3) and crystallization from CH2Cl2/pentanes at 20 C yielded light brown micro crystals of [(6-C6H6)(IiPr2Me2)RuCl2] (0.140 g, 0.325 mmol, 30%). Mp 193C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 1.48 (d, 3JHH = 7.1 Hz, 6 H, CH(CH3)2), 1.55 (d, 3JHH = 7.1 Hz, 6 H, CH(CH3)2), 2.27 (s, 6 H, =CCH3), 5.57 (s, 6 H, C6H6), 5.70 (sp, 3JHH = 7.1 Hz, 2 H, CH(CH3)2).
13C{1H}

NMR (62.90 MHz,

CDCl3, 300 K): 10.7 (s, =CCH3), 23.0 (s, CH(CH3)2), 23.2 (s, CH(CH3)2), 53.5 (s, CH(CH3)2), 85.9 (s, C6H6), 127.4 (s, =CCH3), 169.5 (s, N2C). HR FAB-MS: calcd for [C17H26Cl2N2Ru]: 430.0515; found: 430.0512. MS m/z (%): 430 (5) [M+], 395 (55) [M+ Cl], 357 (75), 275 (27), 181 (100) [IiPr2Me2 + H+]. Subsequently, a solution of [(6-C6H6)(IiPr2Me2)RuCl2] (0.057 g, 0.135 mmol) and NaI (0.596 g, 3.97 mmol) in acetone (20 mL) was stirred overnight at room temperature. The resulting dark purple mixture was evaporated and the residue was extracted

61

Chapter 2

with CH2Cl2, filtered and subsequent crystallization from CH2Cl2/pentanes at 20 C gave dark purple micro crystals of 4 (0.079 g, 0.128 mmol, 95%). Mp 198 C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 1.46 (d, 3JHH = 7.0 Hz, 6 H, CH(CH3)2), 1.53 (d, 3JHH = 7.0 Hz, 6 H, CH(CH3)2), 2.27 (s, 6 H, =CCH3), 5.67 (s, 6 H, C6H6), 5.94 (sp, 3JHH = 7.0 Hz, 2 H, CH(CH3)2). 13C{1H} NMR (62.90 MHz, CDCl3, 300 K): 11.1 (s, =CCH3), 23.4 (s, CH(CH3)2), 23.6 (s, CH(CH3)2), 56.4 (s, CH(CH3)2), 86.3 (s, C6H6), 128.0 (s, =CCH3), N2C was not observed. HRMS (EI, 70 eV): calcd for [C17H26I2N2Ru] [HI]: 486.0100; found: 486.0116. MS m/z (%): 486 (100) [M+ HI], 404 [M+ HI 2iPr], 357 (28) [M+ HI I]. [(6-C6H6)(IMes)Ru=PMes] (7a). A solution of IMes (0.459 g, 1.51 mmol) in toluene (5 mL) was added to a brown toluene (5 mL) suspension of complex [(6-C6H6)RuCl2(PH2Mes)] (0.201 g, 0.50 mmol) at 78 C. The mixture was stirred for 1 h and was slowly allowed to warm up to room temperature. The solvent was evaporated in vacuo and the residue was extracted with pentanes (40 mL). After filtration (P4 glass filter, Microfibre Teflon and Celite) and concentration, dark brown crystals of 7a were obtained at room temperature (0.217 g, 0.343 mmol, 68%). Mp 190 C (dec). 1H NMR (250.13 MHz, C6D6, 300 K): 2.07 (s, 6 H, o-CH3MesP), 2.14 (s, 6 H, p-CH3MesN), 2.28 (s, 12 H, o-CH3MesN), 2.34 (s, 3 H, p-CH3MesP), 4.43 (s, 6 H, C6H6), 6.42 (s, 2 H, NCH=), 6.73 (s, 4 H, m-MesN), 6.89 (s, 2 H, m-MesP).
13C{1H}

NMR (100.64 MHz,

C6D6, 300 K): 19.6 (s, o-CH3MesN), 21.0 (s, p-CH3MesN), 21.3 (s, p-CH3MesP), 21.5 (s, o-CH3MesP), 79.4 (s, C6H6), 122.1 (s, =CH), 127.6 (s, m-MesP), 128.7 (s, m-MesN), 131.5 (s, o-MesP), 133.3 (s, p-MesP), 136.2 (s, o-MesN), 138.0 (s, pMesN), 138.6 (s, i-MesN), 168.7 (d, 1JCP = 87.3 Hz, i-MesP), 190.4 (s, N2C).
31P{1H}

NMR (101.3 MHz, C6D6, 300 K): 752.5 (s, Ru=P). HRMS (EI, 70 eV): calcd for C36H41N2PRu: 634.2051; found: 634.2054. MS m/z (%): 634 (20) [M+], 556 (22) [M+ Bz], 484 (40) [M+ PMes], 322 (100). Crystalline 7a contains, according to 31P NMR, 2.7% of the corresponding toluene adduct 7b. [(6-pCy)(IMes)Ru=PMes] and [(6-Tol)(IMes)Ru=PMes] (7b). A solution of IMes (0.481 g, 1.415 mmol) in toluene (10 mL) was added to a toluene (5 mL) suspension of complex [(6-pCy)RuCl2(PH2Mes)] (0.2125 g, 0.464 mmol) at 78 C. The mixture was stirred for 0.5 h and was slowly allowed to warm up to 62

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

room temperature. After two days, the solvent was evaporated in vacuo and the dark brown solid was extracted with pentanes (20 mL), filtered (P4 glass filter, Teflon microfibre, Celite), and after concentration a mixture of [(6-pCy)(IMes)Ru=PMes] and 7b (5:1 ratio) was obtained as a dark brown solid (0.180 g, 0.261 mmol, 56%). [(6-pCy)(IMes)Ru=PMes]. 1H NMR (400.13 MHz, C6D6, 300 K): 0.84 (d, 3JHH = 6.9 Hz, 6 H, pCy-CH(CH3)2), 1.51 (s, 3 H, pCyCH3), 1.64 (sp, 3JHH = 6.9 Hz, 1 H, pCy-CH(CH3)2), 2.12 (s, 6 H, o-CH3MesP), 2.25 (s, 6 H, p-CH3MesN), 2.32 (s, 12 H, o-CH3MesN), 2.37 (s, 3 H, p-CH3MesP), 4.38 (d, 3JHH = 5.9 Hz, 2 H, C6H4), 4.65 (d, 3JHH = 5.9 Hz, 2 H, C6H4), 6.35 (s, 2 H, NCH=), 6.73 (s, 4 H, m-MesN), 6.94 (s, 2 H, m-MesP).
13C{1H}

NMR (100.64 MHz, C6D6,

300 K): 19.5 (s, pCy-CH3), 19.7 (s, o-CH3NMes), 20.9 (s, o-CH3PMes), 21.3 (s, pCH3PMes), 21.5 (s, p-CH3NMes), 25.2 (s, pCy-CH(CH3)2), 32.2 (s, pCy-CH(CH3)2), 81.1, 83.1, 87.9, and 100.7 (s, C6H4), 122.3 (s, =CH), 127.5 (s, m-MesP), 128.9 (s, m-MesN), 132.3 (s, o-MesP), 132.9 (s, p-MesP), 136.3 (s, i-MesN), 138.0 (s, oMesN), 138.9 (s, p-MesN), 168.3 (d, 1JCP = 100.0 Hz, i-MesP), 191.7 (s, N2C).
31P{1H}

NMR (101.3 MHz, C6D6, 300 K): 724.9 (s, Ru=P). HRMS (EI, 70 eV): calcd

for [C40H49N2PRu] [C9H11P]: 540.2072; found: 540.2091. MS m/z (%): 540 (16) [M+ PMes], 302 (10) [M+ PMes 2Mes]. 7b: 1H NMR (400.13 MHz, C6D6, 300 K): 1.55 (s, 3 H, C6H5CH3), 2.08 (s, 6 H, o-CH3PMes), 2.13 (s, 6 H, p-CH3NMes), 2.28 (bs, 12 H, o-CH3NMes), 2.35 (s, 3 H, p-CH3PMes), 4.28 (d, 3JHH = 5.6 Hz, 2 H, C6H5CH3), 4.34 (t, 3JHH = 5.6 Hz, 1 H, C6H5CH3), 4.59 (t, 3JHH = 5.6 Hz, 2 H,

C6H5CH3), 6.40 (s, 2 H, NCH=), 6.73 (s, 4 H, m-MesN), 6.90 (s, 2 H, m-MesP).
13C{1H}

NMR (100.64 MHz, C6D6, 300 K): 19.6 (s, o-CH3NMes), 19.9 (s, p-

CH3PMes), 20.9 (s, CH3), 21.3 (s, p-CH3PMes), 21.5 (s, o-CH3NMes), 74.9, 82.1, 82.2, and 90.5 (s, C6H5), 122.1 (s, =CH), 127.8 (s, m-MesP), 128.9 (s, m-MesN), 131.8 (s, o-MesP), 133.1 (s, p-MesP), 136.1 (s, i-MesN), 138.0 (s, o-MesN), 138.7 (s, p-MesN), 168.8 (d, 1JCP = 80.0 Hz, i-Mes), 191.7 (s, N2C).
31P{1H}

NMR (101.3

MHz, C6D6, 300 K): 736.8 (s, Ru=P). HRMS (EI, 70 eV): calcd for [C37H43N2PRu] [C9H11P]: 498.1603; found: 498.1593. MS m/z (%): 498 (4) [M+ PMes]. [(6-pCy)(IMes)Ru=PMes] and [(6-d3-Tol)(IMes)Ru=PMes] (d3-7b). A

solution of IMes (0.164 g, 0.54 mmol) in d3-toluene (0.8 mL) was added to a d3-toluene (0.2 mL) suspension of complex [(6-pCy)RuCl2(PH2Mes)] (0.0824 g, 63

Chapter 2

0.180 mmol) at 78C. The mixture was allowed to warm up to room temperature overnight and the resulting dark brown reaction mixture was filtered (P4 glass filter), evaporated in vacuo and crystallized from toluene/pentanes (respectively 1 and 5 mL) to obtain a mixture of [(6pCy)(IMes)Ru=PMes] and d3-7b (5:1 ratio) as a dark brown solid at 20 C (0.0966 g, 0.142 mmol, 79%). d3-7b: 1H NMR (400.13 MHz, C6D6, 300 K): 2.08 (s, 6 H, o-CH3PMes), 2.13 (s, 6 H, p-CH3NMes), 2.28 (bs, 12 H, o-CH3NMes), 2.35 (s, 3 H, p-CH3PMes), 4.28 (d, 3JHH = 5.6 Hz, 2 H, C6H5CH3), 4.34 (t, 3JHH = 5.6 Hz, 1 H, C6H5CH3), 4.59 (t, 3JHH = 5.6 Hz, 2 H, C6H5CH3), 6.40 (s, 2 H, NCH=), 6.73 (s, 4 H, m-MesN), 6.90 (s, 2 H, m-MesP).
13C{1H}

NMR (100.64 MHz, C6D6, 300 K):

19.5 (s, o-CH3NMes), 19.6 (s, p-CH3PMes), 21.4 (s, p-CH3PMes), 21.5 (s, oCH3NMes), 74.0, 82.0, 82.2, and 87.9 (s, C6H5), 122.2 (s, =CH), 127.6 (s, m-MesP), 129.3 (s, m-MesN), 131.8 (s, o-MesP), 133.1 (s, p-MesP), 136.3 (s, i-MesN), 138.0 (s, o-MesN), 138.7 (s, p-MesN), 168.8 (d, 1JCP = 80.0 Hz, i-MesP), 191.8 (s, N2C).
31P{1H}

NMR (101.3 MHz, C6D6, 300 K): 736.6 (s, Ru=P). HRMS (EI, 70 eV): calcd

for C37H40D3N2PRu: 651.2398; found: 651.2380. MS m/z (%): MS m/z (%): 651 (4) [M+], 556 (40) [M+ d3-Tol], 436 (24) [M+ d3-Tol Mes], 304 (20). [(6-C6Me6)(IMes)Ru=PMes]. A solution of IMes (0.258 g, 0.85 mmol) in toluene (1 mL) was added under argon to a toluene (2 mL) suspension of complex [(6-C6Me6)RuCl2(PH2Mes)] (0.136 g, 0.28 mmol) at 78C. The mixture was allowed to slowly warm up to room temperature and the resulting dark brown reaction mixture was filtered (P4 glass filter) and evaporated in vacuo. Extraction with pentanes (5 mL), filtration and concentration at room temperature yielded [(6-C6Me6)(IMes)Ru=PMes] as a dark brown solid (0.116 g, 0.162 mmol, 58%). Mp 198 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 1.66 (s, 18 H, C6(CH3)6), 2.14 (s, 6 H, o-CH3PMes), 2.18 (s, 6 H, p-CH3NMes), 2.21 (s, 6 H, o-CH3NMes), 2.36 (s, 3 H, p-CH3PMes), 2.52 (s, 6 H, o-CH3NMes), 6.33 (s, 2 H, =CH), 6.75 (s, 2 H, m-MesN), 6.79 (s, 2 H, m-MesN), 6.90 (s, 2 H, m-MesP).
13C{1H}

NMR (100.64 MHz, C6D6, 300 K): 16.5

(s, C6(CH3)6), 19.9 (s, o-CH3NMes), 21.0 (s, p-CH3NMes), 21.1 (s, o-CH3NMes), 21.4 (s, p-CH3PMes), 21.5, and 21.6 (s, o-CH3PMes), 90.9 (s, C6Me6), 122.8 (s, =CH), 127.6 (s, m-MesP), 128.4, and 129.8 (s, m-MesN), 132.9 (s, p-MesP), 133.6 64

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

(bs, o-MesP), 135.2 (s, o-MesN), 137.3 (s, o-MesN), 137.5 (s, p-MesN), 139.3 (s, iMesN), 165.6 (d, 1JCP = 97.6 Hz, i-MesP), 192.9 (d, 2JCP = 13.0 Hz, N2C).
31P{1H}

NMR (101.3 MHz, C6D6, 300 K): 686.4 (s, Ru=P). MS EI m/z (%): 718 (<1) [M+, correct isotope pattern], 566 (4), 303 (8), 162 (62), 147 (100).

Reactivity of [(6-C6H6)(IiPr2Me2)Ru=PMes*] (1). CH2I2 (0.153 mmol, 12.3 L) was added to a dark green solution of 1 (19.5 mg, 0.0306 mmol) in C6D6 (0.4 mL) at 20 C. Within 1.5 min., a brown precipitate was formed and the quantitative formation of H2C=PMes* 6 was determined by 31P NMR (t = 22 min at 0 C in d8-toluene). The solvent was evaporated in vacuo and the residue was extracted with pentanes (2 x 5 mL), subsequent filtration and evaporation to dryness yielded H2C=PMes*[31] 6 as an off-white solid (15.0 mg, 94%). Subsequent purification by column chromatography (silica, CH2Cl2/pentanes) of the dark brown residue and crystallization from CH2Cl2/pentanes at 20 C yielded dark purple micro crystals of [(6C6H6)(IiPr2Me2)RuI2] (4) (22.7 mg, 67%).

Reactivity of [(6-C6H6)(PPh3)Ru=PMes*] (2). CH2I2 (0.153 mmol, 12.3 L) was added to a green solution of 2 (22 mg, 0.0306 mmol) in C6D6 (0.4 mL) at 20 C. After 5h and 10 min. (t = 60 min; t = 925 min at 0 C in d8-toluene) all phosphinidene reacted to form H2C=PMes* 4 (45.3%), 5 (31P 21.4 (53.5%)), and several unidentified byproducts.
31P

NMR

(101.3 MHz, C6D6, 293 K): 73.3 (s, 22.6%), 38 (s, 3.8%), 35.9 (d, JPP = 56.1 Hz, 7.5%), 25.6 (d, JPP = 56 Hz, 12.6%).

Acknowledgement. This work was partially supported by the Council for


Chemical Sciences of the Netherlands Organization for Scientific Research (NWO/CW). The assistance from Dr. F. J. J. de Kanter (NMR), Dr. M. Smoluch (HR EI-MS) and J. W. H. Peeters (HR FAB-MS; University of Amsterdam) is gratefully acknowledged.

65

Chapter 2

References and Notes


[1] (a) D. Bourissou, O. Guerret, F. P. Gabba, G. Bertrand, Chem. Rev. 2000, 100, 39 92. (b) M. F. Lappert, J. Organomet. Chem. 2005, 690, 54675473. (c) W. A. Herrmann, Angew. Chem. Int. Ed. 2002, 41, 12901309. (d) N. M. Scott, S. P. Nolan, Eur. J. Inorg. Chem. 2005, 18151828. [2] (a) S. Diez-Gonzalez, S. P. Nolan, Coord. Chem. Rev. 2007, 251, 874883. (b) H. Jacobsen, A. Correa, A. Poater, C. Costabile, L. Cavallo, Coord. Chem. Rev. 2009, 253, 687703. (c) S. Wrtz, F. Glorius, Acc. Chem. Res 2008, 41, 15231533. (d) N-Heterocyclic Carbenes in Synthesis; S. P. Nolan, Ed.; WILEY-VCH: Weinheim, 2006. [3] (a) K. Getty, M. U. Delgado-Jaime, P. Kennepohl, J. Am. Chem. Soc. 2007, 129, 1577415776. (b) M. S. Sanford, M. Ulman, R. H. Grubbs, J. Am. Chem. Soc. 2001, 123, 749750. (c) M. S. Sanford, J. A. Love, R. H. Grubbs, J. Am. Chem. Soc. 2001, 123, 65436554. [4] (a) C. Adlhart, P. Chen, Angew. Chem. Int. Ed. 2002, 41, 44844487. (b) B. F. Straub, Angew. Chem. Int. Ed. 2005, 44, 59745978. [5] (a) J. C. Slootweg, K. Lammertsma, In Science of Synthesis; B. M.Trost, F. Mathey, Eds.; Georg Thieme Verlag: Stuttgart, 2009; Vol. 42, pp 1536. (b) F. Mathey, Dalton Trans. 2007, 18611868. [6] A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 22002208. [7] (a) F. Mathey, Angew. Chem. Int. Ed. 2003, 42, 15781604. (b) P. Le Floch, Coord. Chem. Rev. 2006, 250, 627681. [8] (a) K. J. T. Noonan, D. P. Gates, Angew. Chem. Int. Ed. 2006, 45, 72717274. (b) D. P. Gates, Top. Curr. Chem. 2005, 250, 107126. [9] (a) A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2002, 21, 31963202. (b) A. T. Termaten, H. Aktas, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2003, 22, 1827 1834. [10] X-ray crystal structure determination of 1: C35H55N2PRu, Fw = 635.85, black octahedral block, 0.30 x 0.30 x 0.27 mm3, orthorhombic, Pnma (no. 62), a = 18.9145(7), b = 17.2265(5), c = 10.4262(4) , V = 3397.2(2) 3, Z = 4, Dx = 1.243 g/cm3, = 0.53 mm1. 66347 Reflections were measured on a Nonius Kappa CCD diffractometer with rotating anode (graphite monochromator, = 0.71073 ) up to a resolution of (sin /)max = 0.65 1 at a temperature of 150(2) K. Intensities were integrated with EvalCCD.[32] Absorption correction and scaling was

66

N-Heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes

performed with SADABS[33] (0.71-0.87 correction range). 4041 Reflections were unique (Rint = 0.035), of which 3557 were observed [I>2(I)]. The structure was solved with Direct Methods using the program SIR-97.[34] The structure was refined with SHELXL-97[35] against F2 of all reflections. Non hydrogen atoms were refined with anisotropic displacement parameters. Hydrogen atoms were introduced in calculated positions and refined with a riding model. The tert-butyl group was rotationally disordered and refined with two orientations (with 50% occupancy, each). 208 Parameters were refined with 6 restraints for distances and angles in the disordered tert-butyl group. R1/wR2 [I > 2 (I)]: 0.0335 / 0.0817. R1/wR2 [all refl.]: 0.0408 / 0.0848. S = 1.172. Residual electron density between -0.34 and 0.97 e/3. The maximum residual density of 0.97 e/3 has a distance of 1.30 to P1 and is considered an artifact. Geometry calculations and checking for higher symmetry was performed with the PLATON program.[36] Crystallographic

Information File (cif) with crystallographic data for compound 1 is available free of charge via the Internet at http://pubs.acs.org. [11] S. E. Lehman Jr., K. B. Wagener, Organometallics 2005, 24, 14771482. [12] T. L. Breen, D. W. Stephan, J. Am. Chem. Soc. 1995, 117, 1191411921. [13] R. Menye-Biyogo, F. Delpech, A. Castel, H. Gornitzka, P. Rivire, Angew. Chem. Int. Ed. 2003, 42, 56105612. [14] Only 16-electron zirconium phosphinidenes have been detected by
31P

NMR

spectroscopy, see: (a) J. Ho, Z. Hou, R. J. Drake, D. W. Stephan, Organometallics 1993, 12, 31453157. (b) A. Mahieu, A. Igau, J.-P. Majoral, Phosphorus, Sulfur Silicon Relat. Elem. 1995, 104, 235239. [15] For complexes with a MP triple bond, see: (a) B. P. Johnson, G. Balzs, M. Scheer, Top. Curr. Chem. 2004, 232, 123. (b) G. Balzs, L. J. Gregoriades, M. Scheer, Organometallics 2007, 26, 30583075. [16] When the para-cymene derivative [(6-pCy)RuCl2(PH2Mes)] was used, 15% of 7b was obtained (see the Experimental Section). [17] (a) E. L. Muetterties, J. R. Bleeke, E. J. Wucherer, T. A. Albright, Chem. Rev. 1982, 82, 499525. (b) M. A. Bennett, Z. Lu, X. Wang, M. Bown, D. C. R. Hockless, J. Am. Chem. Soc. 1998, 120, 1040910415. [18] ADF2005.01 and ADF2005.01b, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands, http://www.scm.com/. [19] S. H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 1980, 58, 12001211. [20] A. D. Becke, Phys. Rev. A 1988, 38, 30983100.

67

Chapter 2

[21] J. P. Perdew, Phys. Rev. B 1986, 33, 88228824. [22] (a) L. Fan, L. Versluis, T. Ziegler, E. J. Baerends, W. Raveneck, Int. J. Quantum Chem. Quantum Chem. Symp. 1988, S22, 173181. (b) L. Versluis, T. Ziegler, J. Chem. Phys. 1988, 88, 322328. [23] E. van Lenthe, A. W. Ehlers, E. J. Baerends, J. Chem. Phys. 1999, 110, 89438953. [24] (a) K. Morokuma, Acc. Chem. Res. 1977, 10, 294300. (b) T. Ziegler, A. Rauk, Inorg. Chem. 1979, 18, 17551759. (c) T. Ziegler, A. Rauk, Theor. Chim. Acta 1977, 46, 1 10. [25] R. A. Zelonka, M. C. Baird, Can. J. Chem. 1972, 50, 30633072. [26] M. A. Bennett, T. N. Huang, T. W. Matheson, A. K. Smith, Inorg. Synth. 1982, 21, 74 78. [27] N. Kuhn, T. Kratz, Synthesis 1993, 561562. [28] A. J. Arduengo, III, H. V. R. Dias, R. L. Harlov, M. Kline, J. Am. Chem. Soc. 1992, 114, 55305534. [29] A. H. Cowley, J. E. Kilduff, T. H. Newman, M. Pakulski, J. Am. Chem. Soc. 1982, 104, 58205821. [30] Y. van den Winkel, H. M. M. Bastiaans, F. Bickelhaupt, J. Organomet. Chem. 1991, 405, 183194. [31] R. Appel, C. Casser, M. Immenkeppel, F. Knoch, Angew. Chem. 1984, 96, 905 906; Angew. Chem. Int. Ed. Engl. 1984, 23, 895896. [32] A. J. M. Duisenberg, L. M. J. Kroon-Batenburg, A. M. M. Schreurs, J. Appl. Cryst. 2003, 36, 220-229. [33] G. M. Sheldrick, 1999, SADABS: Area-Detector Absorption Correction, v2.10, Universitt Gttingen, Germany. [34] A. Altomare, M. C. Burla, M. Camalli, G. L. Cascarano, C. Giacovazzo, A. Guagliardi, A. G. G. Moliterni, G. Polidori, R. Spagna, J. Appl. Cryst. 1999, 32, 115119. [35] G. M. Sheldrick, Acta Cryst. 2008, A64, 112-122. [36] A. L. Spek, J. Appl. Cryst. 2003, 36, 7-13.

68

Chapter 3
N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes
Halil Aktas, J. Chris Slootweg, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, and Koop Lammertsma*,

Department of Chemistry and Pharmaceutical Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands, and Bijvoet Center for Biomolecular Research, Crystal Structural Chemistry, Utrecht University, Padualaan 8, 3584 CH Utrecht, The Netherlands

Organometallics 2009, 28, 51665172

Abstract:

The

N-heterocyclic

carbene

(NHC)

functionalized

phosphinidene

complexes

[L(NHC)M=PR] (M = Ru, Os, Rh, Ir) of the Group 8-9 transition metals were synthesized. The effect of the NHC ligand on the electronic properties of the phosphinidene complexes

[(Ring)(NHC)M=PH] 816 bearing group 7-9 transition metals and cycloheptatrienyl (Cht+), 16 benzene, cyclopentadienyl (Cp) as ancillary ligands were studied by density functional theory. On going to the right in the Periodic Table, the structures show an increase in MNHC bond energy that concurs with the net charge on the phosphorus atom of the M=P bond. The NHC ligand effects the M=P properties and interacts via the metal with the ancillary ligand L (Cht+, Bz, Cp).

Chapter 3

3.1

Introduction

Transition metal complexed phosphinidenes display carbene-like behavior that has enabled the synthesis of a plethora of organophosphorus compounds.[1] Transition metals with strong -donor ancillary ligands render nucleophilic complexes[2] akin to Schrocktype carbene complexes. Several synthetic approaches to these isolable species are known, such as the condensation of

lithiumphosphides to halogenated transition metal complexes,[3] oxidation of Ni-phosphides followed by deprotonation,[4] and addition of phosphides to alkylidene complexes with a concurrent -Hmigration.[5] We developed a general route for group 8 (Ru and Os)[6] and group 9 (Co, Rh,[7] and Ir[8]) complexes entailing the double dehydrohalogenation of primary phosphine precursors with DBU (1,8biazabicyclo-[5.4.0]-undec-7-ene) in the presence of stabilizing ligands (PR3, P(OR)3 AsR3, CO, RNC, dppe). Reacting these nucleophilic phosphinidenes with gem-dihalogens gives access to phosphaalkenes (P=C),[9] which are attractive building blocks for Pfunctionalized polymers[10] and as catalyst ligands.[11] Stimulated by the N-heterocyclic carbene (NHC) ligated ruthenium alkylidene complexes that are efficient catalysts in olefin metathesis,[12,13] we recently reported the first NHC-functionalized (1; Ru-complexed Mes* = 2,4,6-

phosphinidene

[(6-C6H6)(IiPr2Me2)Ru=PMes*]

tBu3C6H2), which displays enhanced reactivity toward CH2I2 in comparison to the corresponding phosphine analogue 2 (Scheme 1) due to the sterically imposed ligand rotation of the NHC fragment.[14] We also reported the first related iridium complexed phosphinidene (3) using the lithiumphosphide method (Scheme 1).[3a] In these studies we showed that NHCs,[15] which are commonly regarded as strong donor ligands, can also act as -acceptor ligands, a feat that hitherto
70

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

had been related to the magnitude of the charge density on the metal.[16,17] For 1, we found that the relative -donor/-acceptor ability of the NHC ligand can be tuned by changing its substituents, thereby stabilizing either the in-plane arrangement of the NHC with the Ru=P bond for maximum -overlap or the out-of-plane conformation that makes the NHC an effective -donor.[14] This simple substituent-controlled conformational change of the transition metal ligand impacts the reactivity of the phosphinidene complex, as was demonstrated for the synthesis of phosphaalkenes.[14]

Scheme 1. Synthesis of NHC-functionalized phosphinidenes 1 and 3.

In the present study, we report on the synthetic access to NHCfunctionalized phosphinidene complexes of the group 8 and 9 metals and analyze by DFT calculations the intrinsic electronic properties of a comprehensive set of model compounds, (Ring)M(NHC)=PH, with a broader set of transition metals (M = group 7 (Mn, Tc, Re), group 8 (Fe, Ru, Os), group 9 (Co, Rh, Ir)) and ligands (Ring = cycloheptatrienyl (Cht), benzene (Bz), cyclopentadienyl (Cp)).

71

Chapter 3

3.2

Results and Discussion

First, we describe the synthesis and characterization of novel ruthenium phosphinidenes with different 6-arene and NHC ligands and with different phosphine substituents. Next, we show that the osmium, rhodium, and iridium complexed phosphinidenes can also be accessed by the double dehydrohalogenation-ligation sequence using NHCs. Finally, we perform a detailed theoretical analysis to unveil the intrinsic electronic properties, reactivities, and stabilities of these systems using (Ring)M(NHC)=PH as model for a more complete set of transition metals of groups 7-9.

3.2.1 The

Synthesis of NHC-functionalized group 8 phosphinidenes NHCand p-Cymene (pCy)-ligated ruthenium-complexed 4 (IiPr2Me2 = 1,3like

phosphinidene

[(6-pCy)(IiPr2Me2)Ru=PMes*]

diisopropyl-4,5-dimethylimidazol-2-ylidene)

was

obtained,

ruthenium phosphinidene 1,[14] by a double dehydrohalogenationligation sequence of the primary phosphine [(6-pCy)RuCl2(PH2Mes*)] using three equivalents of IiPr2Me2 in toluene at 78 C (Scheme 2). Crystallization from n-hexane at 20 C yielded dark green thermally stable crystals (60%; 208 C, dec.) that are air and moisture sensitive. Indicative for the formation of 4 is the single
31P

NMR resonance at

733.5 ppm (1:[14] 31P 751.7 ppm), identifying a bent phosphinidene complex,[6-8] and the characteristic
13C

NMR resonance for the Ru

carbene at 188.7 ppm (1:[14] 13C 187.9 ppm). In addition to the arene ligand, also the NHC and the P substituent can be varied, as illustrated by the reaction of the less crowded Ru complexes [(6Ar)RuCl2(PH2Mes)] (Ar = pCy (a), C6Me6 (b)), having mesityl instead of supermesityl substituents, with the bulkier 1,3-dimesityl-imidazol-2ylidene (IMes). The isolated products 5a,b (31P 724.9 (a, 56%),[18] 686.4
72

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

ppm (b, 58%)) are dark brown solids with typical carbene resonances (13C 191.7 (a), 192.9 ppm (b)) and the common E configuration for the Ru=P bond, as established by NOE experiments (Scheme 2).

Mes* Ru PH2 Cl Cl

3 IiPr2Me2 - 2 IiPr2Me2.HCl N

Mes* Ru P N 4

L Cl

Mes Ru PH2 Cl

3 IMes - 2 IMes.HCl Mes N

L Ru P

Mes 5 a, L = pCy b, L = C6Me6

N Mes

Scheme 2. Synthesis of NHC-functionalized ruthenium phosphinidenes 4 and 5.

The facile double dehydrohalogenation-ligation sequence, with two equivalents of NHC acting as a Brnsted base and one equivalent as a carbene ligand, which affords 18-electron NHCfunctionalized phosphinidenes, can be readily extended to other transition metals if their primary phosphine precursor complexes are available.[19] For instance, reaction of the phosphine complex [(6pCy)OsCl2(PH2Mes*)] with three equivalents of IiPr2Me2 in toluene resulted in the formation of the osmium phosphinidene complex [(6pCy)(IiPr2Me2)Os=PMes*] (6), which was isolated as orange crystals in 85% yield (Scheme 3). The single 31P NMR resonance of 6 at 557.6 ppm is highly shielded in comparison to its triphenylphosphine analogue (31P 667.5 ppm),[6] in analogy to the corresponding ruthenium complexes,[14] an may be attributed to the enhanced -donor capacity of the NHC.

73

Chapter 3

3.2.2

Synthesis of NHC-functionalized group 9 phosphinidenes

NHC-functionalized group 9 phosphinidene complexes are equally accessible by this protocol. Thus, rhodium phosphinidene [(5Cp*)(IiPr2Me2)Rh=PMes*] (7; 31P 745.9 ppm) was obtained by the reaction of [(5-Cp*)RhCl2(PH2Mes*)] with three equivalents of IiPr2Me2 in toluene and isolated as dark pink crystals from n-pentane (88%; Scheme 3). Likewise, iridium phosphinidene 3[3a] was generated from [(5-Cp*)IrCl2(PH2Mes*)] and IiPr2Me2 (60%), thereby offering a facile alternative to the lithiumphosphide method. Unfortunately, the attempted synthesis of cobalt phosphinidene [(5-

Cp)(IiPr2Me2)Co=PMes*] from the readily available precursor [(5Cp)CoI2(PH2Mes*)][7] only led to unidentifiable products.

Scheme 3. General synthesis of NHC-functionalized phosphinidene complexes using NHCs.

The Cs-symmetrical structure of rhodium phosphinidene 7 was ascertained by single-crystal X-ray analysis (Figure 1).[20] The geometry around rhodium shows a two-legged piano stool with a characteristic acute C18Rh1P1 angle (83.94(7)), a Rh1P1C1 angle (110.19(8)) that typifies a bent phosphinidene complex, and an E configuration for the Rh1P1 double bond (2.1827(7) ). This latter bond is similar to that of the triphenylphosphine analogue [(5-Cp*)(Ph3P)Rh=PMes*] (2.1903(4) ),[7] whereas its Ru1Cp(cg) bond is significantly shorter (1.8878(10) vs 1.9253(9) [7]). The out-of-plane conformation of the

74

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

NHC fragment, like that of 1[14] and 3,[3a] is caused by steric congestion induced by the isopropyl wingtip groups of the ligand that prevent the coplanar arrangement.[14]

Figure 1. Molecular structure of 7, displaying displacement ellipsoids at the 50% probability level. Only one conformation of the disordered p-tert-butyl group is shown. H-atoms and disordered solvent (pentane) have been excluded for clarity. Cp(cg) denotes the centroid position of the pentamethylcyclopentadienyl ring. Selected bond distances () and angles and torsion angles (): Rh1P1 2.1827(7), Rh1C18 2.036(2), Rh1Cp(cg) 1.8878(10), P1C1 1.858(2), N1C18 1.357(2), N1C19 1.393(2), C19C19a 1.342(4); Rh1P1C1 110.19(8), P1Rh1Cp(cg) 147.47(4), P1Rh1 C18 83.94(7), C18Rh1Cp(cg) 128.60(7); C2aC1C2C3 14.9(3). Symmetry

operation (a) x, 0.5 y, z.

3.3 The

Theoretical calculations on [(Ring)M(NHC)=PH] = model compound [(Ring)M(NHC)=PH] was investigated

computationally for the influence of group 79 transition metals (M = Mn (8), Tc (9), Re (10), Fe (11), Ru (12), Os (13), Co (14), Rh (15), Ir (16); Figure 2) and the stabilizing ligand (Ring = cyclopentadiene anion (Cp), neutral benzene (Bz) and cycloheptatriene cation (Cht+)) on the electronic properties of NHC-functionalized phosphinidenes. We
75

Chapter 3

first discuss their structural properties and then present charge decomposition analyses.

3.3.1

Geometries.

For each complex, all four possible geometrical isomers were calculated with the NHC ligand in-plane and out-of-plane with both the E and Z arrangements of the M=P double bond (see the Appendix 2). The lowest energy Cs-symmetric structures of the NHCfunctionalized phosphinidene complexes 816 on which we

concentrate here all have similar two-legged piano-stool geometries with a coplanar, NHC in-plane arrangement with the E conformer of the M=P bond (Figure 2); selected bond lengths and angles are listed in Table 1. All data for the less stable out-of-plane E, inplane Z, and out-of-plane Z isomers are given in the Appendix 2.

M C N N

M C N N

M C N N

8a M=Mn 9a M=Tc 10a M=Re


Figure 2. BP86/TZP optimized

11a M=Fe 12a M=Ru 13a M=Os


in-plane E structures

14a M=Co 15a M=Rh 16a M=Ir


(Cs symmetry) for

[(Ring)(NHC)M=PH] (Ring = Cht, Bz, Cp; M = Mn (8a), Tc (9a), Re (10a), Fe (11a), Ru (12a), Os (13a), Co (14a), Rh (15a), and Ir (16a)).

76

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

The first-row TM phosphinidenes 8a, 11a, and 14a have rather short M=P distances (2.0912.136 ) and acute M=PH (104.8109.8) and (NHC)CM=P (85.990.8) angles (Table 1). The calculated M=P distances follow the size of the TMs (Mn > Fe > Co), as do the (NHC)C M bond lengths (2.0341.884 ). The second- (2.2332.199 ) and third-row TM (2.2552.215 ) phosphinidenes have slightly longer M=P bond distances. As expected, the (NHC)CM bonds and MRing distances are shorter for the third-row complexes 10a, 13a, and 16a than for those of the second row, due to relativistic contraction.

Table 1. Selected BP86/TZP bond lengths () and angles (deg) for model complexes [(Ring)(NHC)M=PH] 8a16a. M 8a 9a 10a 11a 12a 13a 14a 15a 16a Mn Tc Re Fe Ru Os Co Rh Ir M=P 2.136 2.233 2.255 2.115 2.208 2.227 2.091 2.199 2.215 MC(NHC) 2.034 2.141 2.118 1.952 2.068 2.061 1.884 2.022 1.999 MRing 1.484 1.652 1.668 1.574 1.782 1.788 1.717 1.980 1.944 M=PH 109.8 111.2 110.2 106.9 107.7 107.1 104.8 104.3 104.2 (NHC)CM=P 85.9 85.3 85.2 87.5 85.6 85.9 90.8 86.9 87.7

3.3.2

Energy Decomposition Analysis

The metalNHC bond dissociation energies (BDEs) for the model complexes (for details see the Computational Section), tabulated in Table 2, depend on the transition metal and increases on going from left to right and from top to bottom in the Periodic Table.[21] Thus, the Mn phosphinidene 8a has the weakest MNHC bond (34.0 kcalmol1) and Ir complex 16a the strongest bond (60.6 kcalmol1; Figure 3). The same applies to the - (Ea) and -orbital interactions (Ea) of the MNHC bond, which also increase from left to right and from top to
77

Chapter 3

bottom in the Periodic Table (Table 2). The average -contribution is ~14% of the total orbital interaction (E oi) for the MNHC bond of the group 7 TMs, whereas this amounts to ~18% for the group 8 and group 9 TMs, indicating that -backdonation from the metal to the NHC ligand is substantial for 8a16a and that it cannot be neglected.[17c]

65 60 55 BDE (kcal/mol) 50 45 40 35 30

Mn Tc

Re

Fe

Ru

Os

Co

Rh

Ir

Figure 3. MNHC Bond Dissociation Energies (kcal/mol) for first-row (solid line), second-row (dashed line), and third-row (dotted line) phosphinidene complexes.

Table 3 gives the data for the metalphosphinidene bond. The bond dissociation energies for the first-row metal complexes 8a, 11a and 14a amount to ~70 kcalmol1 and increase for the heavier analogues, being more pronounced for the group 7 metals than those for group 8 and 9, e.g., 9a (91.5) > 12a (88.2) > 15a (82.4 kcalmol1). The phosphinidene group is a strong ligand in all [(Ring)M(NHC)=PH] complexes with a high -contribution of 7478% of the total orbital interaction (Eoi), which is in agreement with earlier findings.[2,3a,7,14] Both the - and -contributions follow the trend noted for the BDEs. Evidently, the yet unknown phosphinidene complexes 8 11 have strong M=P double bonds.
78

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

Table 2. BP86/TZP Bond Dissociation Energies (in kcal/mol) and Energy Decomposition Analysis for LnMC(NHC) bonds of the Group 79 complexes. Values in parentheses refer to charge (e) transferred to (positive) and from (negative) the transition metal center. Ea 7.7 (0.09) 8.8 (0.09) 11.8 (0.11) 11.6 (0.16) 11.7 (0.15) 15.1 (0.18) 13.7 (0.18) 12.6 (0.15) 17.3 (0.19) 23.5 73.3 14.5 56.9 9.5 58.1 17.1 69.3 14.5 53.0 10.8 51.6 11.5 66.9 49.9 42.8 45.5 55.7 51.6 49.5 60.6 6.0 10.3 48.1 51.3 34.0 42.2 E steric BDE E tot

Ea

8a 9a

Mn Tc

46.4 (0.55) 52.7 (0.46)

10a

Re

66.6 (0.49)

11a

Fe

50.9 (0.59)

12a

Ru

55.8 (0.47)

13a

Os

71.2 (0.49)

14a

Co

53.8 (0.63)

15a

Rh

58.8 (0.48)

16a

Ir

79.4 (0.49)

Table 3. BP86/TZP Bond Dissociation Energies and Energy Decomposition Analysis for (Ring)(NHC)M=PH double bond. Energies are in kcal/mol. Ea 30.7 41.7 46.5 28.9 36.9 43.0 27.4 31.6 38.2 66.0 50.4 52.6 60.6 60.6 57.0 71.9 55.4 65.6 76.7 97.0 108.0 74.5 91.6 104.5 74.2 84.5 98.6 E steric E tot E prep 7.5 5.5 7.6 4.0 3.5 5.8 2.2 2.1 3.8 BDE 69.2 91.5 100.4 70.5 88.2 98.7 72.0 82.4 94.8

Ea

8a 9a

Mn Tc

101.4 120.9

10a

Re

133.4

11a

Fe

102.6

12a

Ru

115.3

13a

Os

127.5

14a

Co

97.2

15a

Rh

105.5

16a

Ir

121.0

79

Chapter 3

The reactivity of the phosphinidene complexes toward electron-rich or electron-poor reagents is directed by the charge, which is influenced by the transition metal fragment and can be deferred from the energies of the bond-forming orbitals. The energy for the singly occupied p orbital of 3PH amounts to 5.5 eV. The corresponding energies of the [Ring][NHC]M fragments are collected in Table 4. The major trend for the LnM fragment is the decreasing energy of the -orbital on going down (Mn 2.13 > Tc 2.24 > Re 2.44 eV) or to the right (Mn 2.13 > Fe 2.23 > Co 2.55) of the Periodic Table (Figure 4). The -orbital energies follows the same trend on going from the first- to the second-row TM but increase again for the third-row TM, which may be caused by relativistic effects. Consequently, the net charge on phosphorus (Table 4) decreases in the row group 7 > group 8 > group 9. The second- and third-row complexes show an increased charge on P as compared to the first-row congeners. This effect should be reflected in an increased nucleophilicity that indeed matches our experimental observations. The Ru and Rh phosphinidene complexes are the most reactive ones of the group 8 and 9 triads.[3a,7,14] We briefly return to the effect of in-plane and out-of-plane orientation of the NHC ligand on the electronic properties of the complexes. In earlier studies, we found that the relative -donor/acceptor ability of the NHC ligand in ruthenium complexes can be tuned by changing its substituents, thereby stabilizing either the inplane arrangement of the NHC with the Ru=P bond for maximum overlap or the out-of-plane conformation that makes the NHC an more effective -donor.[14] Unfortunately, a direct discrimination of the and orbital contributions is not possible for the out-of plane conformation since both fall into the a representation. However,
80

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

from the amount of charge transferred (see the Appendix 2) it is apparent that in this case -backdonation is reduced compared to the in-plane arrangement for the group 8 and 9 complexes while it is negligible for the NHC-bond to metals for group 7.

Table 4. BP86/TZP occupied SOMO energies (eV) for the [(Ring)(NHC)M] fragments and calculated Hirshfeld charges of M and P. M 8a 9a 10a 11a 12a 13a 14a 15a 16a Mn Tc Re Fe Ru Os Co Rh Ir Ed () 2.13 2.24 2.44 2.23 2.37 2.56 2.55 2.65 2.78 Ed () 2.35 2.45 2.38 2.51 2.62 2.50 2.78 2.95 2.77 M +0.01 +0.04 +0.13 0.05 +0.15 +0.03 0.07 +0.11 0.01 P 0.15 0.17 0.21 0.12 0.18 0.16 0.10 0.14 0.13

-2.0 -2.1 -2.2 Orbital Energy (eV) -2.3 -2.4 -2.5 -2.6 -2.7 -2.8 -2.9 -3.0

Mn

Fe

Co

Tc

Ru

Rh

Re

Os

Ir

Figure 4. - () and - () Orbital energies (eV) for the (Ring)(NHC)M fragment of first-row (solid line), second-row (dashed line), and third-row (dotted line) phosphinidene complexes.

81

Chapter 3

3.4

Conclusion

The scope of the dehydrohalogenationligation sequence using NHCs both as Brnsted base and as stabilizing ligand has been successfully extended by the synthesis of novel ruthenium, osmium, and rhodium phosphinidene complexes. The solid-state structure of rhodium phosphinidene 7 shows an out-of-plane conformation of the NHC fragment with the (E)-M=P bond and is similar to that reported for Ir and Ru complexes 1 and 3. The DFT calculations present insightful details. The orbital energies of the metalNHC bond are dominated by a -interaction, but the -interaction is also substantial (~20%) and thus not negligible. These interactions are influenced by the ancillary ligand, which was illustrated by a slight change in - and -capacity due to the Cht+ and Cp ligands. For the group 79 metals, the BDEs and strength of the - and -interactions follow the order of the transition metals in the Periodic Table. According to these theoretical calculations, the rhenium, ruthenium, and rhodium phosphinidene complexes are expected to be the most reactive ones of the group 79 triad metals.

3.5

Computational Section

All DFT calculations have been performed with the parallelized Amsterdam density functional (ADF) package (version 2006.01).[22] The KohnSham MOs were expanded in a triple Slatertype basis set with polarization functions (TZP). The 1s core shell of carbon and nitrogen, the 1s2s2p core shells of phosphorus and a small core of the TM were treated by the frozen core approximation. All calculations were performed at the nonlocal exchange selfconsistent field (NLSCF) level, using the local density approximation (LDA) in the VoskoWilkNusair parametrization[23] with nonlocal corrections for exchange (Becke88)[24] and correlation (Perdew86).[25] All geometries were optimized using the analytical gradient method implemented by 82

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

Versluis and Ziegler,[26] including relativistic effects by the Zero Order Regular Approximation (ZORA).[27] Bonding Energy Analysis. The bonding interactions of the transition metal to ligand bonds (MP, MC(NHC), and MRing) were analyzed with ADF's established energy decomposition[28] into an exchange (or Pauli) repulsion (E Pauli) between the electrons on the two fragments plus electrostatic interaction energy part (E elst) and an orbital interaction energy (charge transfer, polarization) part (E oi). The energy necessary to convert fragments from their ground state equilibrium geometries to the geometry and electronic state they acquire in the complex is represented by a preparation energy term (E prep). The overall bond energy (E tot) is formulated as: E tot = E Pauli+E elst+E oi+E prep Note that E tot is defined as the negative of the bond dissociation energy (BDE), i.e., E tot=E (molecule) E (fragments), thereby giving negative values for stable bonds. The orbital interaction term E oi accounts for interactions between occupied orbitals on one fragment with unoccupied orbitals on the other fragment, including HOMO-LUMO interactions and polarization (empty/occupied orbital mixing on the same fragment). The extended transition state (ETS) method developed by Ziegler and Rauk was used to decompose E oi into contributions from each irreducible

representation of the interacting system.[28] In systems with a clear , separation (a' and a), this symmetry partitioning proves to be most informative.

3.6

Experimental Section

General. All experiments and manipulations were performed under an atmosphere of dry nitrogen with rigorous exclusion of air and moisture using flamedried glassware using Schlenk techniques. NMR spectra were recorded on a Bruker Advance 250 (1H, Advance 400 (1H,
13C, 31P; 13C, 31P;

85% H3PO4) or a Bruker

85% H3PO4) and referenced internally to residual


13C{1H}:

solvent resonances (CDCl3: 1H: 7.26,

77.16; C6D6: 1H: 7.16, 83

Chapter 3

13C{1H}:

128.06). IR Spectra were recorded on a Mattson6030 Galaxy FTIR

spectrophotometer, highresolution mass spectra (HRMS) on a Finnigan Mat 900 spectrometer operating at an ionization potential of 70 eV. Melting points were measured on samples in sealed capillaries on a Stuart Scientific SMP3 melting point apparatus and are uncorrected. Reagents. Solvents were distilled from sodium (toluene), P2O5 (CH2Cl2), or LiAlH4 (pentanes, diethyl ether) and kept under an atmosphere of dry nitrogen. Deuterated solvents were dried over 4 molecular sieves (CDCl3, C6D6). All solid starting materials were dried in vacuo. [(6-C6Me6)RuCl2]2,[29] [(6-pCy)RuCl2(PH2Mes*)],[6] [(6-pCy)OsCl2(PH2Mes*)],[6] [(5-

Cp)CoI2(PH2Mes*)],[7] [(5-Cp*)RhCl2(PH2Mes*)],[7] [(5-Cp*)IrCl2(PH2Mes*)],[8] 1,3-diisopropyl-4,5-dimethylimidazol-2-ylidene (IiPr2Me2),[30] 1,3-dimesityl-

imidazol-2-ylidene (IMes),[31] and Mes*PH2[32] were prepared according to literature procedures. MesPH2 was prepared, analogously to IsPH2,[33] by LiAlH4 reduction of MesPCl2. [(6-pCy)RuCl2]2 was purchased from Strem and used as received. [(6-pCy)RuCl2(PH2Mes)]. MesPH2 (0.219 g, 1.437 mmol) was added to a dark red suspension of [(6-pCy)RuCl2]2 (0.400 g, 0.653 mmol) in CH2Cl2 (25 mL), which was stirred for 1 h at room temperature and filtered to remove insoluble material. After concentrating the solution to a few mL, pentanes (30 mL) was added slowly, which resulted in precipitation of an orange powder that was isolated by filtration, washed with pentanes (25 mL), and dried in vacuo to yield [(6-pCy)RuCl2(PH2Mes)] (0.561 g, 1.22 mmol, 94%). Mp 163 C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 1.19 (d, 3JHH = 7.0 Hz, 6 H, CH(CH3)2), 1.99 (s, 3 H, pCy-CH3), 2.34 (s, 3 H, p-CH3), 2.50 (s, 6 H, o-CH3), 2.70 (sp, 3JHH = 7.0 Hz, 1 H, CH(CH3)2), 5.13 (d, 3JHH = 5.7 Hz, 2 H, C6H4), 5.34 (d,
3JHH

= 5.7 Hz, 2 H, C6H4), 5.52 (d, 1JHP = 389.8 Hz, 2 H, PH2); 7.01 (d, 4JHP = 2.3 Hz,
13C{1H}

2 H, m-Mes).

NMR (62.90 MHz, CDCl3, 300 K): 18.4 (s, pCy-CH3), 21.6

(d, 5JCP = 1.1 Hz, p-CH3), 22.5 (s, CH(CH3)2), 22.9 (d, 3JCP = 8.2 Hz, o-CH3), 31.0 (s, CH(CH3)3), 85.5 (d, 2JCP = 5.7 Hz, C6H4), 89.1 (d, 2JCP = 5.3 Hz, C6H4), 97.6 (s, C6H4), 107.7 (d, 2JCP = 1.0 Hz, C6H4), 122.3 (d, 1JCP = 42.3 Hz, i-Mes), 130.1 (d,

84

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

3JCP

= 7.8 Hz, m-Mes), 141.4 (d, 2JCP = 7.7 Hz, o-Mes), 141.5 (d, 4JCP = 2.9 Hz, p31P

Mes).

NMR (101.3 MHz, CDCl3, 300 K): 48.0 (t, 1JPH = 389.9 Hz, PH2). IR

(KBr, cm1): v = 3043.0 (w), 2967.2 (m), 2913.3 (w), 2864.7 (w), 2387.6 (m, PH), 2353.7 (m, PH), 1601.4 (m), 1437.9 (s), 1382.3 (s). HR FAB-MS: calcd for C19H27PCl2Ru: 458.0271; found: 458.0270. MS m/z (%): 458 (70) [M+], 423 (100) [M+ Cl], 388 (100) [M+ 2Cl], 271 (35) [M+ Cl H2PMes], 154 (90) [H2PMes]. [(6-C6Me6)RuCl2(PH2Mes)]. Analogous to that described for [(6-

pCy)RuCl2(PH2Mes)], [(6-C6Me6)RuCl2]2 (1.01 g, 1.50 mmol) and MesPH2 (0.464 g, 3.05 mmol) were used to give [(6-C6Me6)RuCl2(PH2Mes)] as a yellow powder (1.29 g, 2.65 mmol, 88%). Mp 257 C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 1.92 (s, 18 H, C6(CH3)6), 2.29 (s, 3 H, p-CH3), 2.39 (s, 6 H, o-CH3), 5.45 (d, 1JHP = 379.7 Hz, 2 H, PH2), 6.95 (d, 4JHP = 2.1 Hz, 2 H, m-Mes).
13C{1H}

NMR (62.90 MHz, CDCl3, 300 K): 15.7 (s, C6(CH3)6), 21.6 (s, p-CH3),

23.4 (d, 3JCP = 8.0 Hz, o-CH3), 96.1 (d, 3JCP = 3.5 Hz, C6(CH3)6), 119.3 (d, 1JCP = 38.3 Hz, i-Mes), 129.9 (d, 3JCP = 7.6 Hz, m-Mes), 141.1 (d, 4JCP = 2.5 Hz, p-Mes), 142.7 (d, 2JCP = 7.4 Hz, o-Mes).
1JPH 31P

NMR (101.3 MHz, CDCl3, 300 K): 53.7 (t,

= 379.8 Hz, PH2). IR (KBr, cm1): v = 3019.1 (w), 2953.2 (m), 2914.4 (m),

2397.3 (m, PH), 2364.6 (m, PH), 1601.3 (m), 1447.67 (s), 1383.7 (m). HR FABMS: calcd for C21H31PCl2Ru: 486.0584; found: 486.0588. MS m/z (%): 486 (65) [M+], 451 (50) [M+ Cl], 413 (50) [M+ 2HCl], 299 (60) [M+ 2HCl H2PMes], 154 (100) [H2PMes]. [(6-pCy)(IiPr2Me2)Ru=PMes*] (4). A cold solution of IiPr2Me2 (0.119 g, 0.66 mmol) in toluene (4 mL) was added to an orange toluene (5 mL) suspension of complex [(6-pCy)RuCl2(PH2Mes*)] (0.117 g, 0.20 mmol) at 78 C, which immediately resulted in a color change to green. After the mixture was stirred for 1 h at 78 C and warmed up to room temperature, solvents were removed in vacuo. The precipitates were extracted into n-hexane (10 mL) and the dark green suspension was filtered. After concentration, green crystals of 4 were obtained at 20 C (83 mg, 0.12 mmol, 60%). Mp 208 C (dec). 1H NMR (250.13 MHz, C6D6, 300 K): 0.94 (d, 3JHH = 6.9 Hz, 6 H, pCyCH(CH3)2), 1.30 (d, 3JHH = 7.1 Hz, 6 H, CH(CH3)2), 1.37 (d, 3JHH = 7.1 Hz, 6 H, CH(CH3)2), 1.57 (s, 9 H, p-C(CH3)3), 1.64 (s, 3 H, pCy-CH3), 1.82 (s, 18 H, o85

Chapter 3

C(CH3)3), 1.89 (s, 6 H, CH3), 2.37 (sp, 3JHH = 6.9 Hz, 1 H, pCy-CH(CH3)2), 4.63 (d,
3JHH

= 5.7 Hz, 2 H, C6H4), 4.85 (d, 3JHH = 5.7 Hz, 2 H, C6H4), 5.91 (sp, 3JHH = 7.1 Hz,
13C{1H}

2 H, NCH(CH3)2), 7.55 (s, 2 H, m-Mes*).

NMR (100.64 MHz, C6D6, 300 K):

10.5 (s, =CCH3), 18.9 (s, pCy-CH3), 22.0 (s, NCH(CH3)2), 22.2 (s, NCH(CH3)2), 23.9 (s, pCy-CH(CH3)2), 32.2 (s, p-C(CH3)3), 32.2 (s, CH(CH3)2), 32.7 (d, 4JCP = 7.5 Hz, o-C(CH3)3), 34.9 (s, p-C(CH3)3), 38.8 (s, o-C(CH3)3), 54.9 (s, NCH(CH3)2), 76.2, 81.5, 91.7, and 105.9 (s, C6H4), 119.9 (s, m-Mes*), 124.6 (s, =CCH3), 144.9 (s, p-Mes*), 146.3 (s, o-Mes*), 174.1 (d, 1JCP = 111.3 Hz, i-Mes*), 188.7 (s, N2C).
31P{1H}

NMR (101.3 MHz, C6D6, 300 K): 733.5 (s, Ru=P). HRMS (EI, 70 eV):

calcd for C39H63N2PRu: 692.3766; found: 692.3707. MS m/z (%): 692 (8) [M+], 416 (20) [M+ PMes*], 283 (8). [(6-pCy)(IMes)Ru=PMes] (5a). A solution of IMes (0.481 g, 1.415 mmol) in toluene (10 mL) was added to a toluene (5 mL) suspension of complex [(6pCy)RuCl2(PH2Mes)] (0.2125 g, 0.464 mmol) at 78 C. The mixture was stirred for 0.5 h and was slowly allowed to warm up to room temperature. After two days, the solvent was evaporated in vacuo and the dark brown solid was extracted with pentanes (20 mL), filtered (P4 glass filter, Teflon microfibre, Celite), and after concentration a mixture of 5a and 6-toluene adduct [(6Tol)(IMes)Ru=PMes]14 (5:1 ratio) was obtained as a dark brown solid (0.180 g, 0.261 mmol, 56%). 5a: 1H NMR (400.13 MHz, C6D6, 300 K): 0.84 (d, 3JHH = 6.9 Hz, 6 H, pCy-CH(CH3)2), 1.51 (s, 3 H, pCy-CH3), 1.64 (sp, 3JHH = 6.9 Hz, 1 H, pCyCH(CH3)2), 2.12 (s, 6 H, o-CH3MesP), 2.25 (s, 6 H, p-CH3MesN), 2.32 (s, 12 H, oCH3MesN), 2.37 (s, 3 H, p-CH3MesP), 4.38 (d, 3JHH = 5.9 Hz, 2 H, C6H4), 4.65 (d,
3JHH

= 5.9 Hz, 2 H, C6H4), 6.35 (s, 2 H, NCH=), 6.73 (s, 4 H, m-MesN), 6.94 (s, 2 H,

m-MesP). 13C{1H} NMR (100.64 MHz, C6D6, 300 K): 19.5 (s, pCy-CH3), 19.7 (s, o-CH3NMes), 20.9 (s, o-CH3PMes), 21.3 (s, p-CH3PMes), 21.5 (s, p-CH3NMes), 25.2 (s, pCy-CH(CH3)2), 32.2 (s, pCy-CH(CH3)2), 81.1, 83.1, 87.9, and 100.7 (s, C6H4), 122.3 (s, =CH), 127.5 (s, m-MesP), 128.9 (s, m-MesN), 132.3 (s, o-MesP), 132.9 (s, p-MesP), 136.3 (s, i-MesN), 138.0 (s, o-MesN), 138.9 (s, p-MesN), 168.3 (d, 1JCP = 100.0 Hz, i-MesP), 191.7 (s, N2C).
31P{1H}

NMR (101.3 MHz, C6D6, 300

K): 724.9 (s, Ru=P). HRMS (EI, 70 eV): calcd for [C40H49N2PRu] [C9H11P]:

86

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

540.2072; found: 540.2091. MS m/z (%): 540 (16) [M+ PMes], 302 (10) [M+ PMes 2Mes]. [(6-C6Me6)(IMes)Ru=PMes] (5b). A solution of IMes (0.258 g, 0.85 mmol) in toluene (1 mL) was added under argon to a toluene (2 mL) suspension of complex [(6-C6Me6)RuCl2(PH2Mes)] (0.136 g, 0.28 mmol) at 78C. The mixture was allowed to slowly warm up to room temperature and the resulting dark brown reaction mixture was filtered (P4 glass filter) and evaporated in vacuo. Extraction with pentanes (5 mL), filtration and concentration at room temperature yielded 5b as a dark brown solid (0.116 g, 0.162 mmol, 58%). Mp 198 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 1.66 (s, 18 H, C6(CH3)6), 2.14 (s, 6 H, o-CH3PMes), 2.18 (s, 6 H, p-CH3NMes), 2.21 (s, 6 H, o-CH3NMes), 2.36 (s, 3 H, p-CH3PMes), 2.52 (s, 6 H, o-CH3NMes), 6.33 (s, 2 H, =CH), 6.75 (s, 2 H, m-MesN), 6.79 (s, 2 H, m-MesN), 6.90 (s, 2 H, m-MesP).
13C{1H}

NMR (100.64 MHz, C6D6, 300 K): 16.5 (s, C6(CH3)6), 19.9 (s, o-

CH3NMes), 21.0 (s, p-CH3NMes), 21.1 (s, o-CH3NMes), 21.4 (s, p-CH3PMes), 21.5, and 21.6 (s, o-CH3PMes), 90.9 (s, C6Me6), 122.8 (s, =CH), 127.6 (s, mMesP), 128.4, and 129.8 (s, m-MesN), 132.9 (s, p-MesP), 133.6 (bs, o-MesP), 135.2 (s, o-MesN), 137.3 (s, o-MesN), 137.5 (s, p-MesN), 139.3 (s, i-MesN), 165.6 (d, 1JCP = 97.6 Hz, i-MesP), 192.9 (d, 2JCP = 13.0 Hz, N2C).
31P{1H}

NMR (101.3

MHz, C6D6, 300 K): 686.4 (s, Ru=P). MS EI m/z (%): 718 (<1) [M+], 566 (4), 303 (8), 162 (62), 147 (100). [(6-pCy)(IiPr2Me2)Os=PMes*] (6): A solution of IiPr2Me2 (0.079 g, 0.438 mmol) in toluene (2 mL) was added under vigorous stirring to a yellow solution of [(6-pCy)OsCl2(PH2Mes*)] (0.098 g, 0.146 mmol) in toluene (4 mL), which resulted in a immediate color change to dark orange. After 16 h at room temperature, all volatiles were removed in vacuo and the product was extracted into pentane (15 mL) and filtered through a short Celite column. After concentration of the pentane solution, orange crystals of 6 were obtained at room temperature (0.097 g, 0.124 mmol, 85%). Mp 230 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 0.92 (d, 3JHH = 6.9 Hz, 6 H, pCyCH(CH3)2), 1.25 (d, 3JHH = 7.2 Hz, 6 H, NCH(CH3)2), 1.40 (d, 3JHH = 7.2 Hz, 6 H,

87

Chapter 3

NCH(CH3)2), 1.54 (s, 9 H, p-C(CH3)3), 1.70 (s, 3 H, pCy-CH3), 1.87 (s, 6 H, =CCH3), 1.89 (s, 18 H, o-C(CH3)3), 2.28 (sept, 3JHH = 6.9 Hz, 1 H, pCy-

CH(CH3)2), 4.86 (d, 3JHH = 5.5 Hz, 2 H, C6H4), 5.10 (d, 3JHH = 5.5 Hz, 2 H, C6H4), 5.89 (sept, 3JHH = 7.2 Hz, 2 H, NCH(CH3)2), 7.51 (s, 2 H, m-Mes*).
13C{1H}

NMR

(100.64 MHz, C6D6, 300 K): 10.4 (s, =CCH3), 19.1 (s, pCy-CH3), 21.7 (s, NCH(CH3)2), 21.8 (s, NCH(CH3)2), 24.2 (s, pCy-CH(CH3)2), 32.2 (s, p-C(CH3)3), 32.4 (s, pCy-CH(CH3)2), 33.0 (d, 4JCP = 6.8 Hz, o-C(CH3)3), 34.6 (s, p-C(CH3)3), 38.6 (s, o-C(CH3)3), 56.4 (s, NCH(CH3)2), 68.4, 73.6, 85.5, and 101.1 (s, C6H4), 119.8 (s, m-Mes*), 123.9 (s, =CCH3), 144.5 (s, p-Mes*), 147.0 (s, o-Mes*), 174.3 (s, N2C), 175.3 (d, 1JCP = 108.0 Hz, i-Mes*).
31P

NMR (101.3 MHz, C6D6, 300 K):

557.6 (s, P=Os). IR (KBr, ): v 3055 (w), 2960 (vs), 2900 (s), 2863 (s), 1782 (w), 1712 (w), 1644 (m), 1574 (s), 1476, 1462, and 1444 (m), 1388 (m), 1376 (m), 1354 (s), 1291 (s), 1261 (m), 1238 (m), 1211 (m), 1104 (w), 1084 (w), 1070 (m), 1026 (m), and 1019 (m), 864 (s), 800 (m), 785 (s), 739 (s), 684 (w), 642 (w), 537 cm1 (w). HRMS (EI, 70 eV): calcd for C39H63N2OsP: 782.4344; found: 782.4364. MS m/z (%): 782 (<1) [M+]. [(5-Cp*)(IiPr2Me2)Rh=PMes*] (7): A red solution of [(Cp*)RhCl2(PH2Mes*)] (0.269 g, 0.457 mmol) in toluene (15 mL) was added dropwise under vigorous stirring to a solution of IiPr2Me2 (0.247 g, 1.372 mmol) in toluene (5 mL), which resulted in a immediate color change to dark pink. After 24 h at room temperature, all volatiles were removed in vacuo and the product was extracted into pentane (45 mL) and filtered through a short Celite column. After concentration to few mL, large dark pink crystals of 7C5H12 were obtained at 4 C (0.309 g, 0.403 mmol, 88%). Mp 204.8 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 0.88 (t, 3JHH = 7.3 Hz, pentane), 1.24 (m, 3JHH = 7.3 Hz, pentane), 1.38 (d, 3JHH = 7.1 Hz, 6 H, NCH(CH3)2), 1.39 (d, 3JHH = 7.1 Hz, 6 H, NCH(CH3)2), 1.56 (s, 9 H, p-C(CH3)3), 1.61 (s, 15 H, Cp*), 1.77 (s, 18 H, oC(CH3)3), 1.88 (s, 6 H, =CCH3), 5.77 (sept, 3JHH = 7.1 Hz, 2 H, NCH(CH3)2), 7.57 (s, 2 H, m-Mes*). 13C{1H} NMR (100.64 MHz, C6D6, 300 K): 10.2 (s, =CCH3), 10.4 (s, C5(CH3)5), 14.2 (s, CH3, pentane), 22.5 (s, NCH(CH3)2), 22.7 (s, CH2, pentane), 23.0 (s, NCH(CH3)2), 31.6 (bs, o-C(CH3)3), 32.1 (s, p-C(CH3)3), 34.4 (s, CH2, pentane), 34.7 (s, o-C(CH3)3), 38.7 (s, p-C(CH3)3), 54.1 (s, NCH(CH3)2), 88

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

93.5 (d, 2JCP = 4.2 Hz, C5(CH3)5), 120.6 (s, m-Mes*), 125.0 (bs, =CCH3), 144.0 (s, o-Mes*), 144.4 (s, p-Mes*), 170.0 (d, 1JCP = 120.1 Hz, i-Mes*), 185.0 (d, 1JCRh = 63.7 Hz, N2CRh).
31P

NMR (101.3 MHz, C6D6, 300 K): 745.9 (d, 1JPRh = 70.4 Hz,

P=Rh). IR (KBr): v 3076 (w), 2955 (s), 2900 (s), 2868 (s), 1738 (w), 1643 (m), 1586 (s), 1462 (s), 1380 (m), 1358 (s), 1290 (s), 1239 (m), 1210 (m), 1189 (m), 1130 (w), 1104 (w), 1068 (w), 1022 (m), 870 (m), 806 (bw), 746 (s), 700 (w), 642 (w), 590 (w), 518 cm1 (w). HRMS (EI, 70 eV): calcd for C39H64N2PRh: 694.3862; found: 694.3867. MS m/z (%): 694 (8) [M+], 418 (40) [M+ PMes*], 376 (100) [M+ PMes* C3H6]. [(5-Cp*)(IiPr2Me2)Ir=PMes*] (3): Iridium complex [(5-Cp*)IrCl2(PH2Mes*)] (0.223 g, 0.33 mmol) was added to a solution of IiPr2Me2 (0.184 g, 1.02 mmol) in toluene (4 mL) at 78C. The resulting reaction mixture was allowed to warm up to room temperature and, after 18 h, the mixture was filtered (P4 glass filter) and evaporated in vacuo. Extraction of the product into pentanes (5 mL), filtration and removal of all volatiles in vacuo followed by crystallization from toluene at 20C yielded 3 as red crystals (0.155 g, 0.198 mmol, 60%) with identical spectroscopic data as previously reported.3a

Acknowledgement.

This

work

was

supported

by

The

Netherlands

Foundation for Chemical Sciences (CW) with financial aid from the Netherlands Organization for Scientific Research (NWO). Dr. M. Smoluch and J. W. H. Peeters (University of Amsterdam) are acknowledged for measuring respectively HR EIMS and HR FABMS.

Supporting information Available.

1H

NMR spectral data of all novel

compounds, cartesian coordinates () and energies (a.u.) of all stationary points, and the cif file with crystallographic data for compound 7. This material is available free of charge on the Internet at http://pubs.acs.org.

References and Notes


[1] (a) J. C. Slootweg, K. Lammertsma, In Science of Synthesis; B. M. Trost, F. Mathey,

89

Chapter 3

Eds.; Georg Thieme Verlag: Stuttgart, 2009; Vol. 42, pp 1536. (b) F. Mathey, Dalton Trans. 2007, 18611868. (c) K. Lammertsma, Top. Curr. Chem. 2003, 229, 95119. (d) K. Lammertsma, M. J. M. Vlaar, Eur. J. Org. Chem. 2002, 11271138. (e) F. Mathey, N. T. Tran Huy, A. Marinetti, Helv. Chim. Acta 2001, 84, 29382957. (f) D. W. Stephan, Angew. Chem. 2000, 112, 322338; Angew. Chem. Int. Ed. 2000, 39, 314329. (g) A. H. Cowley, Acc. Chem. Res. 1997, 30, 445451. [2] A. W. Ehlers, E. J. Baerends, K. Lammertsma, J. Am. Chem. Soc. 2002, 124, 2831 2838. [3] (a) A. T. Termaten, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 35773582. (b) C. C. Cummins, R. R. Schrock, W. M. Davis, Angew. Chem. 1993, 105, 758761; Angew. Chem. Int. Ed. 1993, 32, 756759. (c) Z. Hou, D. W. Stephan, J. Am. Chem. Soc. 1992, 114, 1008810089. [4] R. Melenkivitz, D. J. Mindiola, G. L. Hillhouse, J. Am. Chem. Soc. 2002, 124, 3846 3847. [5] (a) F. Basuli, J. Tomaszewski, J. C. Huffman, D. J. Mindiola, J. Am. Chem. Soc. 2003, 125, 1017010171. (b) F. Basuli, B. C. Bailey, J. C. Huffman, M.-H. Baik, D. J. Mindiola, J. Am. Chem. Soc. 2004, 126, 19241925. (c) G. Zhao, F. Basuli, U. J. Kilgore, H. Fan, H. Aneetha, J. C. Huffman, G. Wu, D. J. Mindiola, J. Am. Chem. Soc. 2006, 128, 1357513585; and references therein. [6] A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 22002208. [7] A. T. Termaten, H. Aktas, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2003, 22, 18271834. [8] A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2002, 21, 31963202. [9] (a) T. L. Breen, D. W. Stephan, J. Am. Chem. Soc. 1995, 117, 1191411921. For reviews on phosphaalkenes, see: (b) P. Le Floch, Coord. Chem. Rev. 2006, 250, 627681. (c) F. Mathey, Angew. Chem. 2003, 115, 16161643; Angew. Chem. Int. Ed. 2003, 42, 15781604. (d) L. Weber, Eur. J. Inorg. Chem. 2000, 24252441. (e) M. J. Yoshifuji, J. Chem. Soc. Dalton Trans. 1998, 33433349. (f) L. Weber, Angew. Chem. 1996, 108, 292310; Angew. Chem. Int. Ed. Engl. 1996, 35, 271288. (g) A. C. Gaumont, J. M. Denis, Chem. Rev. 1994, 94, 14131439. (h) F. Mathey, Acc. Chem. Res. 1992, 25, 9096. (i) R. Appel, Multiple Bonds and Low Coordination in Phosphorus Chemistry (Eds.: M. Regitz, O. J. Scherer), Thieme, Stuttgart, 1990.

90

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

[10] (a) C.-W. Tsang, M. Yam, D. P. Gates, J. Am. Chem. Soc. 2003, 125, 14801481. (b) D. P. Gates, Top. Curr. Chem. 2005, 250, 107126. (c) K. J. T. Noonan, D. P. Gates, Angew. Chem. 2006, 118, 74297432; Angew. Chem. Int. Ed. 2006, 45, 72717274. [11] F. Ozawa, M. Yoshifuji, Dalton Trans. 2006, 49874995. [12] (a) T. Weskamp, W. C. Schattenmann, M. Spiegler, W. A. Herrmann, Angew. Chem. 1998, 110, 26312633; Angew. Chem. Int. Ed. 1998, 37, 24902493. (b) M. Scholl, T. M. Trnka, J. P. Morgan, R. H. Grubbs, Tetrahedron Lett. 1999, 40, 2247 2250. (c) J. Huang, E. D. Stevens, S. P. Nolan, J. L. Petersen, J. Am. Chem. Soc. 1999, 121, 26742678. (d) J. Huang, H.J. Schanz, E. D. Stevens, S. P. Nolan, Organometallics 1999, 18, 53755380. (e) C. W. Bielawski, R. H. Grubbs, Angew. Chem. 2000, 112, 30253028; Angew. Chem. Int. Ed. 2000, 39, 29032906. [13] R. H. Grubbs, Angew. Chem. 2006, 118, 38453850; Angew. Chem. Int. Ed. 2006, 45, 37603765. [14] H. Aktas, J. C. Slootweg, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, J. Am. Chem. Soc., 2009, 131, 66666667. [15] M. F. Lappert, J. Organomet. Chem. 2005, 690, 54675473. [16] L. Mercs, G. Labat, A. Neels, A. Ehlers, M. Albrecht, Organometallics 2006, 25, 56485656. [17] (a) C. Boehme, G. Frenking, Organometallics 1998, 17, 58015809. (b) S. DiezGonzalez, S. P. Nolan, Coord. Chem. Rev. 2007, 251, 874883. (c) H. Jacobsen, A. Correa, A. Poater, C. Costabile, L. Cavallo, Coord. Chem. Rev. 2009, 253, 687 703. [18] 5a contains 15% of toluene adduct [(6-Tol)(IMes)Ru=PMes] (31P 736.8), which is formed by arene exchange with the solvent; see for more details ref.14. [19] The primary phosphine complexes are unavailable for the Group 7 transition metals as well as for the iron complex [(6-Ar)FeCl2 (PH2Mes*)]. [20] X-ray crystal structure determination of 7. C39H64N2PRh 0.66C5H12, Fw = 742.42, black block, 0.54 x 0.48 x 0.18 mm3, monoclinic, P21/m (no. 11), a = 10.4727(3), b = 15.1067(4), c = 13.7281(2) , = 95.432(2), V = 2162.14(9) 3, Z = 2, Dx = 1.14 g/cm3, = 0.46 mm1. 28979 Reflections were measured on a Nonius Kappa CCD diffractometer with rotating anode (graphite monochromator, = 0.71073 ) up to a resolution of (sin /)max = 0.65 1 at a temperature of 150(2) K. Intensity integration was performed with EvalCCD.[34] The SADABS[35] program was used for absorption correction based on multiple measured reflections (0.62-0.92 transmission range). 5158 Reflections were unique (Rint = 0.021), of which 4627

91

Chapter 3

were observed [I>2(I)]. The structure was solved with Direct Methods using the program SHELXS-97[36]. The structure was refined with SHELXL-97[36] against F2 of all reflections. Non hydrogen atoms were refined with anisotropic displacement parameters. All hydrogen atoms were introduced in calculated positions and refined with a riding model. The p-tert-butyl group was refined with a disorder model with respect to the mirror plane. The pentane solvent molecule was also refined with a disorder model and with partial occupancy. 270 Parameters were refined with 72 restraints concerning the disorder and the solvent molecule. R1/wR2 [I > 2(I)]: 0.0258 / 0.0861. R1/wR2 [all refl.]: 0.0351 / 0.0906. S = 1.053. Residual electron density between -0.26 and 0.79 e/3. Geometry calculations and checking for higher symmetry was performed with the PLATON program.[37] [21] G. Frenking, I. Antes, M. Boehme, S. Dapprich, A. Ehlers, V. Jonas, A. Neuhaus, M. Otto, R. Stegmann, A. Veldkamp, S. F. Vyboyshchikov, in Reviews in

Computational Chemistry; Lipkowitz, K. B.; Boyd, D. B., Eds.; VCH: New York, 1996; Vol. 8, p 63. [22] ADF2006.01, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands, http://www.scm.com/. [23] S. H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 1980, 58, 12001211. [24] A. D. Becke, Phys. Rev. A 1988, 38, 30983100. [25] J. P. Perdew, Phys. Rev. B 1986, 33, 88228824. [26] (a) L. Fan, L. Versluis, T. Ziegler, E. J. Baerends, W. Raveneck, Int. J. Quantum Chem. Quantum Chem. Symp. 1988, S22, 173181. (b) L. Versluis, T. Ziegler, J. Chem. Phys. 1988, 88, 322328. [27] E. van Lenthe, A. W. Ehlers, E. J. Baerends, J. Chem. Phys. 1999, 110, 89438953. [28] (a) K. Morokuma, Acc. Chem. Res. 1977, 10, 294300. (b) T. Ziegler, A. Rauk, Inorg. Chem. 1979, 18, 17551759. (c) T. Ziegler, A. Rauk, Theor. Chim. Acta 1977, 46, 1 10. [29] M. A. Bennett, T. N. Huang, T. W. Matheson, A. K. Smith, Inorg. Synth. 1982, 21, 74 78. [30] N. Kuhn, T. Kratz, Synthesis 1993, 561562. [31] Arduengo A. J. Arduengo, III, H. V. R. Dias, R. L. Harlov, M. Kline, J. Am. Chem. Soc. 1992, 114, 55305534. [32] A. H. Cowley, J. E. Kilduff, T. H. Newman, M. Pakulski, J. Am. Chem. Soc. 1982, 104, 58205821.

92

N-Heterocyclic Carbene Functionalized Group 79 Transition Metal Phosphinidene Complexes

[33] Y. van den Winkel, H. M. M. Bastiaans, F. Bickelhaupt, J. Organomet. Chem. 1991, 405, 183194. [34] A. J. M. Duisenberg, L. M. J. Kroon-Batenburg, A. M. M. Schreurs, J. Appl. Cryst. 2003, 36, 220229. [35] G. M. Sheldrick, 1999, SADABS: Area-Detector Absorption Correction, v2.10, Universitt Gttingen, Germany. [36] G. M. Sheldrick, Acta Cryst. 2008, A64, 112122. [37] A. L. Spek, J. Appl. Cryst. 2003, 36, 713.

93

Chapter 3

94

Chapter 4
Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides
Halil Aktas, Jos Mulder, Frans J. J. de Kanter, J. Chris Slootweg, Marius Schakel, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, and Koop Lammertsma*,
Department of Chemistry and Pharmaceutical Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands, and Bijvoet Center for Biomolecular Research, Crystal Structural Chemistry, Utrecht University, Padualaan 8, 3584 CH Utrecht, The Netherlands J. Am. Chem. Soc. 2009, 131, 1353113537

Abstract: 18-Electron nucleophilic, Schrock-type phosphinidene complexes 3 [Cp*(XyNC)Ir=PAr] (Ar = Mes*, Dmp, Mes) are capable of unprecedented [1+2]cycloadditions with one equivalent of isocyanide RNC (R = Xy, Ph) to give novel

iridaphosphirane

complexes

[Cp*(XyNC)IrPArC=NR].

Their

structures

were

ascertained by Xray diffraction. Density functional theory investigations on model structures revealed that the iridaphosphirane complexes are formed from the addition of the isocyanide to 16-electron species [Cp*Ir=PAr] forming first complex 3 that subsequently reacts with another isocyanide to give the products following a different pathway than its nitrogen analogue [Cp*IrNt-Bu] 1.

Chapter 4

4.1 Introduction Ever since stable nucleophilic phosphinidenes were discovered by Lappert et al. in 1987,[1] the structural properties and intriguing reactivity of these Schrock-type species continue to fascinate.[2] Exemplary valuable transfer reactions are those with oxo- and halophilic transition metal complexes,[3] such as the phospha-Wittig reaction with carbonyl compounds that yields phosphaalkenes[4,5] and the P/O-exchange with epoxides that gives the threemembered phosphiranes.[4] Phosphaalkenes also result on reaction with geminal dihalides,[4,6,7] while other dihalides give phosphorus heterocycles.[4] Four-membered phosphametallocycles are

accessible by [2+2]-cycloadditions with alkenes[8] and alkynes,[9] thereby mimicking the behavior of Schrock-type carbenes.[10] In the present study we compare the behavior of phosphinidene and imido complexes. Inspired by the reported reaction of stable, 18-electron imido complex [(5-Cp*)IrNt-Bu] 1 with isocyanides that gives

2-

coordinated carbodiimide complex 2,[11] we targeted this reaction for the corresponding phosphinidenes.[13] At the outset there is an intriguing difference between imido complex 1 and its phosphorus analogue, namely 16-electron complex [(5-Cp*)Ir=PR] is not a stable species[12] and whereas many 18-electron [(5-Cp*)(L)Ir=P Mes*] (L = PR3, AsR3, dppe, RNC, CO, NHC) have been reported,[6,14] it is not known to be capable of [1+2]-cycloadditions like its wellestablished Fischer-type electrophilic counterpart.[15] Isocyanide-stabilized 18-electron [(5-Cp*)(XyNC)Ir=PMes*] 3a (Mes* = 2,4,6-tBu3C6H2),[6] generated by double dehydrohalogenation[6,16] of [(5-Cp*)IrCl2(PH2Mes*)] and concomitant ligation with 2,6xylyl isocyanide, is an ideal starting point for the present study in
96

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

which we examine the Ir=P reactivity toward isocyanides and explore the scope of reaction for the transient, in-situ generated 16-electron complex [(5-Cp*)Ir=PR]. We use DFT calculations to address differences between the imido and phosphinidene complexes.

4.2

Results and Discussion of deep pink iridium phosphinidene [(5-Cp*)(Xy

Reaction

NC)Ir=PMes*] 3a ( 31P 757.1)[6] with a tenfold excess of phenyl or 2,6xylyl isocyanide at room temperature resulted in the formation of yellow crystalline [1+2]-adducts 4a and 5 [(Cp*)(XyNC)IrPMes*C=NR] (R = 4a, Xy; 5, Ph) as sole products in respectively 77% and 87% yield after crystallization (Scheme 1). The highly shielded resonances in the
31P

NMR spectrum at = 190.4 (4a) and 190.2 ppm (5) are

diagnostic for three-membered P-rings and indicate the formation of the desired iridaphosphiranes in analogy to iridaaziridine 2. Singlecrystal X-ray analysis of 4a and 5 established unequivocally the unique 2-(P,C)-phosphaazaallene moiety[17] and the linear 2,6-xylyl isocyanide, both coordinated to iridium (Figure 1).[18] The ca. 2.41 Ir P bond of the IrPC ring is elongated from the reported 2.172.21 Ir=P double bond of phosphinidene complexes like 3a (carrying PPh3, CO, and NHC ligands instead of an isocyanide),[6,14] whereas the ca. 2.03 IrC bond compares well with the 2.017(9) reported for the IrNC ring in carbodiimide complex 2;[11c] the ca. 1.80 PC bond length is normal for three-membered P-rings.

97

Chapter 4

Scheme 1. Synthesis of iridaphosphiranes 4a and 5.

Figure 1. Displacement ellipsoid plot (50% probability) of 4a and 5. Hydrogen atoms are omitted for clarity. Selected bond lengths [] and angles [] for 4a and 5 (in square brackets): Ir1P1 2.4147(5) [2.4082(6)], Ir1C19 2.036(2) [2.029(2)], Ir1C38[C36] 1.900(2) [1.895(2)], P1C1 1.8766(19) [1.877(2)], P1C19 1.805(2) [1.799(2)], N1C19 1.267(2) [1.265(3)], N1C20 1.424(3) [1.424(3)], N2C38[C36] 1.169(3) [1.168(3)], N2 C39[C37] 1.402(3) [1.396(3)]; Ir1P1C19 55.46(6) [55.40(7)], P1Ir1C19 46.91(6) [46.89(7)], Ir1C38[C36]N2 177.27(19) [178.3(2)], C19N1C20 125.99(18) [123.2(2)], C38[C36]N2C39[C37] 171.7(2) [175.9(2)].

The described reaction illustrates that different isocyanides can be embedded in the product, something that appears not feasible for the imido complexes.[10] The question is than whether the reaction mechanisms are the same for the two systems. Does the distinction lie in the stability of the 16-electron intermediate or the lack thereof? By reducing the stability of the 18-electron phosphinidene complex and
98

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

attempting to synthesize iridaphosphiranes directly via in situ generated 16-electron iridium phosphinidene complexes, we

examined whether the reaction with isocyanides would mimic more closely that of imido complex 1. The first step was to reproduce the formation of 4a. Double dehydrohalogenation[6,16] of orange colored primary phosphine complex [(5-Cp*)IrCl2(PH2Mes*)] (6a) with two equivalents of 1,8-diazabicyclo-[5.4.0]undec-7-ene (DBU) at 78 C in the presence of ten equivalents of 2,6-xylyl isocyanide resulted in an immediate color change to deep purple, indicative for the formation of 18-electron 3a, and upon warming to room temperature to yellow to give after crystallization indeed iridaphosphirane 4a in 86% isolated yield (Scheme 2). This protocol was extended to other phosphine complexes as illustrated in Scheme 2 for [(5-Cp*)IrCl2(PH2R)] 6b (R = Mes) and 6c (R = Dmp = 2,6-dimesitylphenyl).[19]

Scheme 2. Synthesis of iridaphosphiranes 4 via in situ-generated phosphinidenes 3.

Mesityl-substituted

6b

undergoes

facile

double

dehydrohalogenation-ligation with two equivalents of DBU and an excess of 2,6-xylyl isocyanide to afford iridacycle 4b (
31P

218.7) as

the sole product, which was isolated by crystallization (60%; Scheme 2) and structurally characterized by a single-crystal X-ray structure determination (Figure 2). In this case, the sterically less shielded
99

Chapter 4

phosphinidene intermediate [(5-Cp*)(XyNC)Ir=PMes] 3b could not be detected by 31P NMR spectroscopy under the reaction conditions; the accordingly obtained more encumbered Dmp derivative 3c (deep purple) was observed (
31P

768.1) and converted at room

temperature to yellow complex 4c (64 %; 31P 216.7).

Figure 2. Displacement ellipsoid plot (50% probability) of 4b. Hydrogen atoms are omitted for clarity. Selected bond lengths [] and angles [] for 4b: Ir1P1 2.3952(6), Ir1C10 2.039(2), Ir1C29 1.888(2), P1C1 1.837(2), P1C10 1.799(3), N1C10 1.267(3), N1C11 1.429(3), N2C29 1.172(3), N2C30 1.398(3); Ir1P1C10 56.04(8), P1Ir1C10 47.02(7), Ir1C10P1 76.94(9), Ir1C29N2 179.4(3), C10N1C11 123.8(2), C29N2C30 161.9(2).

4.3

Transient species

Is the suggested 16-electron [(5-Cp*)Ir=PR] indeed formed on double dehydrohalogenation of 6 in the absence of isocyanides (7; Scheme 2)? So far, only 16-electron zirconium phosphinidenes have been observed in arene solvents by ppm)[20] and as
31P 31P

NMR spectroscopy ( Cp*2Zr=PMes*-LiCl

31P

438526 in

unstable

adduct

dimethoxyethane (

537 ppm).[21] In contrast to imido analogue 1,

attempted access to Mes- and Mes*-substituted 16-electron [(5Cp*)Ir=PR] (7a,b) only led to dimers.[12] DFT calculations have shown
100

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

the 16-electron imido and phosphinidene complexes to differ with the first having a linear IrNH arrangement with an IrN triple bond[22] and [(5-Cp)Ir=PH] having a bent conformation (<IrPH 126.4) with an Ir=P double bond,[12] a difference with consequences for their reactivity. By introducing the sterically demanding Dmp substituent on phosphorus, we envisioned to stabilize the 16-electron species by shielding the Ir=P double bond. Double dehydrohalogenation of primary phosphine complex [(5-Cp*)IrCl2(PH2Dmp)] (6c) with two equivalents of DBU in CD2Cl2 at 10 C showed after warm up to room temperature a low-field
31P

NMR resonance at 672 ppm, suggesting

indeed the formation of a bent phosphinidene. To confirm its identity the


31P

NMR chemical shifts were calculated for the E and Z

conformers of [(5-Cp*)Ir=PDmp] 7c and their solvent adducts (Figure 3).

Figure 3. Intermediate E- and Z-[7c(CH2Cl2)] calculated at the BP86/TZP level of theory. Hydrogen atoms are omitted for clarity. Selected bond lengths [] and angles [] for E-[7c(CH2Cl2)] and Z-[7c(CH2Cl2)] (in square brackets): IrP 2.192 [2.206], IrCl1 2.567 [2.368]; IrPDmp 125.2 [114.6], PIrCl1 87.5 [102.9].

101

Chapter 4

Those for solvent-free E-7c and Z-7c (178.9 and 418.2 ppm, respectively) are significantly shielded from the experimental one, but those for the solvent adducts are in the expected range (E[7c(CH2Cl2)] 457.2, Z-[7c(CH2Cl2)] 683.6 ppm; Figure 3) with an excellent match for the more stable Z-conformer (E = 0.3 kcalmol1). Therefore, we conclude that the Z-conformer is also a likely intermediate when applying smaller donor ligands.[6,16a,23] 4.4 Mechanism

The experimental work leads us to conclude that the in situ formed 16-electron phosphinidene coordinates with an isocyanide to the observable and isolable 18-electron complex 3, which gives a [1+2]cycloaddition with another isocyanide molecule to form

iridaphosphirane 4. The reaction with imido complex 1 is different in that this species is stable and that there are no indications for an observable isocyanide coordinated 18-electron imido complex. The question then arises whether this distinction is due to the relative stabilities of the reactants, thus whether the first or second isocyanide addition is rate-determining, or to different mechanistic pathways. To answer this question we resorted to density functional theory at BP86/TZP to compare the reaction of the 16-electron imido and phosphinidene complexes 1' and 7' with isocyanide HNC using model structures incorporating only H substituents.

4.4.1

Imido complex

We begin with the imido complex, [(5-Cp)IrNH] 1'. The reaction starts with the addition of the isocyanide to give bent 18-electron imido complex 8' (E = 26.1 kcalmol1; Figure 4),[24] followed by ring closure to 16-electron 2-carbodiimide complex 9', requiring 18.4

102

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

kcalmol1 (E = 10.9 kcalmol1), and subsequent addition of a second isocyanide to afford the 48.3 kcalmol1 more stable product 2'.

Figure 4. Relative BP86/TZP energies (in kcalmol1) for the reaction of [(5-Cp)IrNH] 1' with HNC. Selected bond lengths [] and angles [] for 8': Ir1N1 1.876, Ir1C1 1.859, C1N2 1.210, Ir1N1H 105.5; TS8'9': Ir1N1 1.883, Ir1C1 1.926, C1N1 1.900, C1N2 1.216, Ir1N1H 120.1; 9': Ir1N1 1.913, Ir1C1 2.040, C1N1 1.352, C1N2 1.275, Ir1N1 H 153.1; 2': Ir1N1 2.124, Ir1C1 2.076, C1N1 1.337, C1N2 1.264, Ir1C2 1.843, C2N3 1.218, Ir1N1H 115.5.

4.4.2

Phosphinidene complex

The mechanism for formation of iridaphosphirane 4 is different. Namely, addition of HNC to [(5-Cp)Ir=PH] 7' is far more exothermic in giving the bent 18-electron species ([(5-Cp)(HNC)Ir=PH] 3' (E = 58.2 kcalmol1; Figure 5) that, moreover, is energetically prohibited to undergo ring closure to 10 (E = 27.1 kcalmol1) and instead is susceptible for attack of a second isocyanide to the phosphorus center to give 1-(P,C)-phosphaazaallene complex 11' (E = 17.6
103

Chapter 4

kcalmol1), which rearranges barrierless to the more favorable 2coordinated product 4' (E = 20.3 kcalmol1).

Figure 5. Relative BP86/TZP energies (in kcalmol1) for the reaction of [(5-Cp)Ir=PH] 7' with HNC. Selected bond lengths [] and angles [] for 3': Ir1P1 2.223, Ir1C1 1.858, C1N2 1.213, Ir1P1H 98.9; 10': Ir1P1 2.307, Ir1C1 2.025, P1C1 1.818, C1N2 1.268, Ir1P1H 110.7; 11': Ir1P1 2.241, Ir1C1 1.834, Ir1C2 3.403, P1C2 1.726, C1N1 1.229, C2N2 1.217, Ir1P1H 122.9; 4': Ir1P1 2.439, Ir1C1 1.850, Ir1C2 2.066, P1C2 1.813, C1N1 1.216, C2N2 1.258, Ir1P1H 101.2, C2Ir1P1 46.6.

The higher nucleophilicity of the imido complex originates from the lower energies of the ligand orbitals (NH 6.79 eV; PH 5.46 eV) and is reflected in the calculated charges (0.27 on N in 8'; 0.04 on P in 3').[25] Two differences that define the dissimilar chemistries of the phosphinidene and imido complexes are revealed by the calculated reaction pathways. First, as noted, the initial ligand addition to the 16electron complex is far more exothermic for the phosphinidene than for imido complex 8', which suffers from distortion of the IrNR unit
104

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

from linearity. The second difference lies in the electronic structure of the ring-closed structures of 9' and 10'. An allenic, -conjugated NC N moiety is formed in Cs-symmetric 9' that binds in a bidentate 2fashion to the metal, but in phosphorus analogue 10' -conjugation is less favorable due to the smaller overlap between the C(2p) and P(3p) orbitals so that the allenic NCP unit binds instead to the metal in an 2-fashion, acting as a 2-electron and not a 4-electron donor as is the case for 9'. Therefore, nucleophilic attack of the phosphorus of 3' at the isocyanide is prohibited, while a modest barrier is observed for the corresponding nitrogen attack of 8'. As a consequence, stable 18-electron phosphinidene complexes can be observed experimentally whereas the imido complexes are prone to

rearrangements. Formation of the final product 4' occurs by direct attack of a second isocyanide to the phosphorus, but this process can be hindered by steric congestion using bulky substituents, which explains why 3a can be observed at low temperatures by NMR spectroscopy even in an excess of isocyanide.

4.5

Conclusions

Imido and phosphinidene iridium complexes differ in their properties and reactivities as established for the reaction with isocyanides. Both isolated and in-situ generated 18-electron iridium phosphinidene [(5Cp*)(XyNC)Ir=PMes*] 3 afford iridaphosphirane 4 as sole product. Despite this apparent resemblance with the stable 16-electron imido complex 1 that gives iridaazirane 2, the course of events is entirely different for the two reactions as elucidated by DFT calculations. The imido complex uses the first isocyanide molecule to construct a 16electron 2-carbodiimide complex, whereas for the phosphorus analogue it is the second isocyanide molecule that induces the ring
105

Chapter 4

closure, thereby giving the unique 2-phosphaazaallene complex. It is further established that the sterically demanding dimesitylphenyl substituent enables the detection of the solvent-stabilized 16-electron phosphinidene intermediate [(5-Cp*)Ir=PDmp] 7c, generated by double dehydrogenation of phosphine precursor 6, prior to the reaction with isocyanides.

4.6

Computational Section

All density functional theory calculations have been performed with the parallelized Amsterdam density functional (ADF) package (version 2005.01b and 2006.01).[26] The KohnSham MOs were expanded in a large, uncontracted basis set of Slatertype orbitals (STOs), of a triple basis set with polarization functions quality, corresponding to basis set TZP in the ADF package. The 1s core shell of carbon and nitrogen and the 1s2s2p core shells of phosphorus were treated by the frozen core approximation. The transition metal centers were described by a triple basis set for the outer ns, np, nd and (n+1)s orbitals, whereas the shells of lower energy were treated by the frozen core approximation using a small core. All calculations were performed at the nonlocal exchange selfconsistent field (NLSCF) level, using the local density approximation (LDA) in the VoskoWilkNusair parameterization[27] with nonlocal corrections for exchange (Becke88)[28] and correlation (Perdew86).[29] All geometries were optimized using the analytical gradient method implemented by Versluis and Ziegler,[30] including relativistic effects by the Zeroth Order Regular Approximation (ZORA).[31] The
31P

NMR chemical shift tensors were calculated with ADFs

NMR program,[32] using singlepoint calculations with an allelectron basis for P within the ZORAapproximation on the optimized frozen core structures (vide supra), using the E-isomer of [Cp*(Me3P)Ir=PMes*][6] as reference ( 455.2 ppm) for the total isotropic shielding tensors ( +629.3 ppm with respect to 85% H3PO4). All model complexes were calculated without molecular symmetry.

106

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

4.7

Experimental Section

General. All experiments and manipulations were performed under an atmosphere of dry nitrogen with rigorous exclusion of air and moisture using flamedried glassware using Schlenk techniques. Solvents were distilled from sodium (toluene), CaCl2 (CH2Cl2), or LiAlH4 (pentanes, diethyl ether) and kept under an atmosphere of dry nitrogen. Deuterated solvents were purchased from Aldrich and dried over 4 molecular sieves (CD2Cl2, CDCl3, C6D6). All solid starting materials were dried in vacuo. NMR spectra were recorded on a Bruker Advance 250 (1H,
13C, 31P; 13C, 31P;

85% H3PO4) or a Bruker Advance 400 (1H,

85% H3PO4) and referenced internally to residual solvent resonances 7.26,


13C{1H}:

(CDCl3: 1H: 7.16,

77.16; CD2Cl2: 1H: 5.25, 13C{1H}: 54.00; C6D6: 1H:

13C{1H}:

128.06). IR Spectra were recorded on a Mattson6030

Galaxy FTIR spectrophotometer, highresolution mass spectra (HRMS) on a Finnigan Mat 900 spectrometer operating at an ionization potential of 70 eV, and fast atom bombardment (FAB) mass spectrometry was carried out using a JEOL JMS SX/SX 102A foursector mass spectrometer (70 eV). Melting points were measured on samples in sealed capillaries on a Stuart Scientific SMP3 melting point apparatus and are uncorrected. [(5-Cp*)IrCl2(PH2Mes*)] (6a),[6] [(5-Cp*)IrCl2(PH2Mes)] (6b),[6] [(5-Cp*)(XyNC)Ir=PMes*] (3a),[6] and Mes*PH2,[33] were prepared according to literature procedures. PhPH2, MesPH2, and DmpPH2[34] were prepared analogously to IsPH2,[35] by LiAlH4 reduction of respectively PhPCl2, MesPCl2, and DmpPCl2.[36] PhNC[37] was prepared by dehydration of the corresponding formamide with phosphoryl chloride. 2,6-Xylyl isocyanide (XyNC) was purchased from Fluka and used as received. [(5-Cp*)IrCl2(PH2Dmp)] (6c). A mixture of freshly prepared DmpPH2 (1.00 g, 2.89 mmol) and [(5-Cp*)IrCl2]2 (0.59 g, 0.74 mmol) in CH2Cl2 (50 mL) was stirred for 30 min. at room temperature. Evaporation to dryness and chromatography of the residue over silica with CHCl3, followed by CHCl3/diethyl ether 4:1, as eluent and subsequent crystallization from CH2Cl2/pentane at 20 C yielded [(5-Cp*)IrCl2(PH2Dmp)] (6c) (1.08 g, 1.45 mmol, 98%) as orange crystals. Mp: 258 C (dec). 1H NMR (400.13 MHz, CDCl3, 107

Chapter 4

300 K): 1.32 (d, 4J(H,P) = 3.3 Hz, 15H; C5(CH3)5), 2.15 (s, 12H; o-CH3), 2.34 (s, 6H; p-CH3), 5.49 (d, 1J(H,P) = 378.6 Hz, 2H; PH2), 6.95 (s, 4H; m-MesH), 7.05 (dd,
3J(H,H)

= 7.5 Hz, 4J(H,P) = 2.9 Hz, 2H; m-PhH), 7.48 (t, 3J(H,H) = 7.5 Hz, 1H; pNMR (100.64 MHz, CDCl3, 300 K): 7.9 (s; C5(CH3)5), 21.2 (s; p-

PhH).

13C{1H}

CH3), 21.5 (s; o-CH3), 92.1 (d, 2J(C,P) = 3.0 Hz; C5(CH3)5), 121.1 (d, 1J(C,P) = 52.0 Hz; ipso-Ph), 128.9 (s; m-Mes), 129.7 (d, 3J(C,P) = 7.8 Hz; m-Ph), 131.1 (d, 4J(C,P) = 2.1 Hz; p-Ph), 136.7 (s; p-Mes), 137.3 (s; o-Mes), 138.0 (d, 3J(C,P) = 4.3 Hz; ipso-Mes), 146.9 (d, 2J(C,P) = 8.8 Hz; o-Ph).
31P

NMR (101.3 MHz, CDCl3, 300 K):

82.8 (t, 1J(P,H) = 378.6 Hz, PH2). IR (KBr): 3023.9 (w), 2988.2 (m), 2962.1 (m), 2915.9 (s), 2854.1 (w), 2383.6 (m, PH), 2370.1 (m, PH), 1610.3 (s), 1565.0 (s), 1448.3 (s), 1376.9 (s), 1027.9 (s), 921.8 (s), 845.6 (s), 806.1 (s), 745.4 (s), 460.9 cm
1

(s). HR FABMS: calcd for C34H42Cl2PIr: 744.2018, found 744.2024. m/z (%): 744

(5) [M]+, 709 (100) [M Cl]+, 673 (12) [M Cl HCl]+, 363 (95) [M Cl DmpPH2]+. [(5-Cp*)IrCl2(PH2Ph)] (6d). A freshly prepared solution of PhPH2 (0.6 M in Et2O, 1.05 mL, 0.63 mmol) was added to an orange solution of [(5-Cp*)IrCl2]2 (0.250 g, 0.314 mmol) in CH2Cl2 (20 mL) at room temperature. After 30 min., the resulting mixture was filtered over a short silica column and eluted with CHCl3, after which the orange fractions were combined and evaporated to dryness. Subsequent crystallization from CH2Cl2/pentane at 20 C yielded [(5-Cp*)IrCl2(PH2Ph)] (6d) as yellow microcrystals, which were washed with pentane and dried in vacuo (0.268 g, 0.528 mmol, 84%). Mp: 222 C (dec). 1H NMR (250.13 MHz, CDCl3, 300 K): 1.63 (s, 15H; C5(CH3)5), 5.88 (d, 1J(H,P) = 394.1 Hz, 2H; PH2), 7.447.47 (m, 3H; m- and p-PhH), 7.787.85 (m, 2H; o-PhH).
13C{1H}

NMR (62.90 MHz, CDCl3, 300 K): 8.6 (s; C5(CH3)5), 91.9 (d, 2J(C,P) = 3.3

Hz; C5(CH3)5), 122.6 (d, 1J(C,P) = 53.3 Hz; ipso-Ph), 128.8 (d, 2J(C,P) = 10.7 Hz; oPh), 131.5 (d, 4J(C,P) = 2.7 Hz; p-Ph), 133.5 (d, 3J(C,P) = 8.8 Hz; m-Ph).
31P

NMR

(101.3 MHz, CDCl3, 300 K): 56.3 (t, 1J(P,H) = 394.1 Hz, PH2). IR (KBr): 3051.8 (s), 2973.7 (s), 2914.9 (s), 2870.5 (m), 2412.5 (w, PH), 2389.4 (m, PH), 1451.2 (s), 1435.8(s), 1378.9 (s), 1156.1 (w), 1070.3 (m), 1031.7 (s), 899.6 (s), 745.4 (s), 697.1 cm1 (s). HR FABMS: calcd for C16H22Cl2PIr: 508.0451, found 508.0459. m/z (%): 508 (8) [M]+, 473 (29) [M Cl]+, 363 (20) [M Cl PhPH2]+.

108

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

[(Cp*)(XyNC)IrPMes*C=NXy]

(4a).

An

orange

solution

of

[(5-

Cp*)IrCl2(PH2Mes*)] (6a; 0.135 g, 0.20 mmol) in CH2Cl2 (2 mL) was added to a mixture of DBU (59.8 L, 0.40 mmol) and XyNC (0.262 g, 2.0 mmol) in CH2Cl2 (3 mL) at 78 C, which resulted in a immediate color change to deep purple. After 1 h, the reaction mixture was allowed to warm up to room temperature and stirred for an additional hour. After evaporation to dryness, the yellow residue was washed with pentane (2 x 1 mL) and extracted into diethyl ether (4 x 15 mL), and the solution was filtered. After concentration of the solution to a few mL, 4a (0.148 g, 0.171 mmol, 86%) was obtained as yellow crystals by crystallization at 20 C. Mp: 143 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 1.11 (s, 9H; p-C(CH3)3), 1.60 (s, 15H; C5(CH3)5), 1.92 (s, 9H; o-C(CH3)3), 1.93 (s, 6H; o-C=NXyCH3), 2.02 (s, 9H; o-C(CH3)3), 2.39 (s, 3H; oCNXyCH3), 2.45 (s, 3H; o-CNXyCH3), 6.69 (d, 3J(H,H) = 7.5 Hz, 2H; m-C=NXy), 6.75 (m, 3J(H,H) = 7.5 Hz, 1H; p-C=NXy), 6.92 (m, 3J(H,H) = 7.0 Hz, 1H; m-CNXy), 6.95 (m, 3J(H,H) = 7.0 Hz, 1H; p-CNXy), 7.04 (m, 3J(H,H) = 7.0 Hz, 1H; m-CNXy), 7.06 (s, 1H; m-Mes*), 7.29 (s, 1H; m-Mes*).
13C{1H}

NMR (100.64 MHz, C6D6, 300

K): 9.4 (s; C5(CH3)5), 19.5 (s; o-C=NXyCH3), 20.0 and 22.1 (s; o-CNXyCH3), 31.2 (s; p-C(CH3)3), 33.9 (d, 4J(C,P) = 11.2 Hz; o-C(CH3)3), 34.3 (s; p-C(CH3)3), 34.6 (d, 4J(C,P) = 6.3 Hz; o-C(CH3)3), 39.8 and 41.0 (s; o-C(CH3)3), 97.1 (s; C5(CH3)5), 121.9 (s; m-Mes*), 122.7 (s; p-CNXy), 123.8 (s; m-Mes*), 125.9 (s; pC=NXy), 126.7 (s; o-CNXy), 127.8 (s; m-CNXy), 127.9 (s; m-C=NXy), 128.8 (s; m-CNXy), 129.7 (s; o-CNXy), 130.3 (s; o-C=NXy), 132.1 (d, 1J(C,P) = 92.1 Hz; C=NXy), 134.1 (s; ipso-C=NXy), 141.5 (s; CNXy), 146.1 (s; p-Mes*), 150.6 (d,
4J(C,P)

= 12.9 Hz; ipso-CNXy), 156.9 (d, 2J(C,P) = 8.4 Hz; o-Mes*), 158.4 (s; o31P

Mes*), 180.6 (d, 1J(C,P) = 103.4 Hz; ipso-Mes*).

NMR (101.3 MHz, C6D6, 300

K): 190.4 (s; PMes*). IR (KBr): v = 3062.4 (w), 2960.2 (s), 2948.6 (s), 2903.3 (s), 2863.8 (s), 2272.7 (very broad w, PC=N), 2072.2 and 2020.1 (s, CN), 1637.3 and 1586.2 (s, C=N), 1463.7 (s, C=C), 1390.4, 1381.8 and 1359.6 (s, PAr), 1240.0 (s), 1195.7 and 1185.1 (s, P=C), 1124.3 (w), 1090.6 (w), 1025.0 (m, PAr), 922.8 (w), 873.6 (w), 773.3 and 745.4 (s, PC), 708.7 (w), 697.8 (m), 522.6 cm1 (m). HR EIMS: calcd for C37H53IrNP (M CNXy) 735.3548, found 735.3541. m/z (%): 866 (2) [M]+, 735 (100) [M CNXy]+, 590 (16) [M PMes*]+, 455 (24) [M 109

Chapter 4

PMes* Cp*]+. Alternatively, 4a can be prepared from 3a as follows: XyNC (0.197 g, 1.50 mmol) was added to a dark purple solution of 3a (0.110 mg, 0.15 mmol) in pentanes (20 mL) at room temperature. After 15 h., the yellow reaction mixture was evaporated to dryness and the residue was washed with pentane (2 x 1 mL) and extracted into diethyl ether. After filtration and concentration to a few mL, yellow crystals of 4a (0.113 g, 0.131 mmol, 87%) were obtained at 20 C.

[(Cp*)(XyNC)IrPMes*C=NPh] (5). A freshly prepared solution of PhNC (0.143 M, 10.5 mL, 1.50 mmol) in CH2Cl2 was added to a dark purple solution of 3a (0.110 mg, 0.15 mmol) in pentane (10 mL) at room temperature. After 15 h., the yellow reaction mixture was evaporated to dryness and the residue was extracted into pentane. After removal of a black residue by filtration, all volatiles were removed in vacuo. The residual yellow solid was extracted into diethyl ether and after concentration to a few mL, yellow crystals of 5 (0.097 g, 0.116 mmol, 77%) were obtained at 20 C. Mp: 159 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 0.93 (s, 9H; p-C(CH3)3), 1.62 (s, 15H; C5(CH3)5), 1.94 (s, 15H; o-C(CH3)3 and o-CNXyCH3), 2.05 (s, 9H; o-C(CH3)3), 6.63 (d, 3J(H,H) = 7.4 Hz, 2H; m-CNXy), 6.70 (t, 3J(H,H) = 7.4 Hz, 1H; p-CNXy), 7.00 (t, 3J(H,H) = 7.3 Hz, 1H; p-C=NPh), 7.03 (d, 4J(C,P) = 7.8 Hz, 1H; m-Mes*), 7.18 (d, 3J(H,H) = 7.9 Hz, 2H, m-C=NPh), 7.45 (d, 4J(C,P) = 7.8 Hz, 1H; m-Mes*), 7.82 (d, 3J(H,H) = 7.9 Hz, 2H; o-C=NPh). 13C{1H} NMR (100.64 MHz, C6D6, 300 K): 9.4 (s; C5(CH3)5), 19.0 (s; o-CNXyCH3), 31.0 (s; p-C(CH3)3), 33.6 (d, 4J(C,P) = 11.3 Hz; o-C(CH3)3), 34.1 (s; p-C(CH3)3), 34.6 (d, 4J(C,P) = 6.8 Hz; o-C(CH3)3), 39.9 and 40.9 (s; oC(CH3)3), 96.7 (s; C5(CH3)5), 119.6 (s; m-Mes*), 122.2 (s; m-Mes*), 122.6 (s; oC=NPh), 123.8 (s; p-C=NPh), 125.8 (s; p-CNXy), 127.9 (s; m-CNXy), 128.8 (s; m-C=NPh), 129.9 (s; ipso-CNXy), 131.6 (d, 1J(C,P) = 80.0 Hz; C=NPh), 133.9 (s; o-CNXy), 140.6 (bs, CNXy), 146.2 (s; p-Mes*), 152.9 (d, 3J(C,P) = 13.1 Hz; ipso-C=NPh), 156.9 (d, 2J(C,P) = 8.2 Hz; o-Mes*), 157.7 (s; o-Mes*), 187.6 (d,
1J(C,P)

= 101.4 Hz; ipso-Mes*).

31P

NMR (101.3 MHz, C6D6, 300 K): 190.2 (s;

PMes*). IR (KBr): v = 3069.2 (m), 3047.0 (w), 2966.0 (m), 2945.6 (s), 2913.9 (m), 2898.5 (w), 2866.7 (w), 2851.3 (w), 2283.3 (very broad w, PC=N), 2084.7 and 2037.4 (s, CN), 1677.8, 1644.0, 1583.3, and 1544.7 (s, C=N), 1480.1, 1462.8, and 110

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

1441.5 (s, C=C), 1390.4, 1375.0, and 1359.6 (m, PAr), 1315.2 (m), 1261.2 (m), 1198.5 (m, P=C), 1066.4 (m), 1026.0 (m, PAr), 898.7 (m), 875.5 (m), 789.7, 763.7, 747.3, and 735.7 (s, PC), 692.3 and 683.6 (s), 523.6 (m), 496.6 cm1 (w). HR EI MS: calcd for C37H53IrNP (M CNPh) 735.3548, found 735.3517. m/z (%): 838 (2) [M]+, 735 (5) [M CNPh]+, 707 (5) [M CNXy]+, 562 (6) [M PMes*]+.

[(5-Cp*)(XyNC)IrPMesC=NXy] (4b). DBU (59.8 L, 0.40 mmol) was added to a yellow solution of [(5-Cp*)IrCl2(PH2Mes)] (6b; 0.110 g, 0.20 mmol) and XyNC (0.262 g, 2.0 mmol) in CH2Cl2 (6 mL) at 78 C, and the mixture was allowed to warm up to room temperature. After evaporation to dryness, the residue was washed with pentane (3 mL) and extracted into diethyl ether (2 x 10 mL), and the solution was filtered. After concentration of the solution to a few mL a yellow solid was obtained, which was recrystallized from diethyl ether at 20C to yield 4b (0.089 g, 0.120 mmol, 60%) as yellow crystals. Mp: 122 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 1.74 (s, 15H; C5(CH3)5), 1.81 (s, 6H; o-C=NXyCH3), 1.85 (s, 3H; p-MesCH3), 2.46 (s, 6H; o-MesCH3), 2.84 (s, 6H; o-CNXyCH3), 6.44 (bs, 2H; m-Mes), 6.64 (d, 3J(H,H) = 7.3 Hz, 2H; m-C=NXy), 6.71 (m, 3J(H,H) = 7.3 Hz, 1H; p-C=NXy), 6.99 (m, 3J(H,H) = 7.0 Hz, 1H; p-CNXy), 7.06 (d, 3J(H,H) = 7.0 Hz, 2H; m-CNXy).
13C{1H}

NMR (100.64 MHz, C6D6, 300 K):

9.5 (s; C5(CH3)5), 18.8 (s; o-C=NXyCH3), 20.6 (d, 3J(C,P) = 16.5 Hz, o-MesCH3), 20.8 (s; p-MesCH3), 23.8 and 23.9 (s; o-CNXyCH3), 96.7 (s; C5(CH3)5), 122.7 (s; p-CNXy), 125.9 (s; p-C=NXy), 127.4 (s; m-C=NXy), 127.5 (s; o-CNXy), 128.0 (s; p-Mes), 128.1 (s; m-CNXy), 128.7 (s; m-Mes), 130.2 (s; ipso-C=NXy), 133.7 (s; oC=NXy), 135.4 (s; o-Mes), 143.5 (d, 2J(C,P) = 9.7 Hz, CNXy), 151.9 (d, 1J(C,P) = 70 Hz, ipso-Mes), C=NXy and ipso-CNXy could not be detected.
31P

NMR

(101.3 MHz, C6D6, 300 K): 218.7 (s; PMes). IR (KBr): v = 3034.6 (w), 2972.8 (w), 2940.0 (w), 2911.0 (m), 2886.0 (m), 2871.5 (m), 2842.6 (w), 2281.4 (very broad w, PC=N), 2057.7 and 2019.1 (s, CN), 1638.2 and 1587.1 (s, C=N), 1459.9 and 1437.7 (s, C=C), 1376.0 (s, PAr), 1240.0 (s), 1187.9 (m, P=C), 1025.0 (m, PAr), 984.5 (m), 840.8 (s), 774.3 and 761.7 (s, PC), 678.8 (s), 517.8 cm1 (s). HR EIMS: calcd for C37H44IrN2P 740.2871, found 740.2896. m/z (%): 740 (<1) [M]+, 609 (1) [M CNXy]+, 590 (20) [M PMes]+.

111

Chapter 4

[(Cp*)(XyNC)IrPDmpC=NXy]

(4c).

An

orange

solution

of

[(5-

Cp*)IrCl2(PH2Dmp)] (6c; 0.075 g, 0.10 mmol) in CH2Cl2 (1.5 mL) was added to a mixture of DBU (29.9 L, 0.20 mmol) and XyNC (0.131 g, 1.0 mmol) in CH2Cl2 (1 mL) at 78 C and the reaction mixture was allowed to warm up to room temperature. After evaporation to dryness, the residue was washed with pentane (2 x 1 mL) and extracted into diethyl ether (2 x 10 mL), and the solution was filtered. After concentration of the solution to a few mL, 4c (0.060 g, 0.064 mmol, 64%) was obtained as yellow crystals by crystallization at 20 C. Mp: 175 C (dec). 1H NMR (400.13 MHz, C6D6, 346 K): 1.41 (s, 15H; C5(CH3)5), 1.97 (s, 6H; o-C=NXyCH3), 2.15 (s, 6H; o-MesCH3), 2.20 (s, 6H; oMesCH3), 2.23 (s, 3H; p-MesCH3), 2.50 (s, 3H; p-MesCH3), 2.58 (s, 6H; oCNXyCH3), 6.77 (m, 5H; m-C=NXy, m-Mes, and p-C=NXy), 6.84 (m, 3J(H,H) = 7.4 Hz, 3H; m-PhP and p-CNXy), 6.96 (d, 3J(H,H) = 7.4 Hz, 2H; m-CNXy), 6.98 (s, 2H; m-Mes), 7.08 (t, 3J(H,H) = 7.4 Hz, 1H; p-PhP).
13C{1H}

NMR (100.64 MHz,

C6D6, 300 K): 9.1 (d, 3J(C,P) = 2.9 Hz; C5(CH3)5), 19.2 (s; o-C=NXyCH3), 19.8 (s; p-MesCH3), 21.0 (s; o-MesCH3), 22.022.2 (bs; o-MesCH3, o-CNXyCH3, oCNXyCH3, and p-MesCH3), 96.9 (s; C5(CH3)5), 122.4 (s; p-CNXy), 126.0 (s, oCNXy), 126.1 (s; o-CNXy), 126.2 (s; p-C=NXy), 127.1 (s; m-CNXy), 127.2 (s; pPhP), 128.0 (s; ipso-Mes, and m-Mes), 128.5 (s; m-C=NXy), 128.6 (s; m-CNXy), 129.0 (s; m-Mes), 129.8 (s; p-Mes), 129.9 (bs; m-PhP), 130.6 (s; ipso-C=NXy), 134.5 (s; o-C=NXy), 136.1 and 136.5 (s; o-Mes), 138.0 (d, 1J(C,P) = 69.5 Hz; ipsoPhP), 140.8 (s; ipso-Mes), 144.0 (s; CNXy), 148.0 (s; ipso-CNXy), 148.2 (d,
2J(C,P)

= 12.9 Hz; o-Ph), 173.3 (d, 1J(C,P) = 93.5 Hz; P-C=NXy).

31P

NMR (101.3

MHz, C6D6, 300 K): 216.7 (s; PDmp). IR (KBr): v = 3036.4 (m), 2977.6 (m), 2949.6 (s), 2909.1 (m), 2851.3 (m), 2268.9 (very broad w, PC=N), 2054.8 and 2012.4 (s, CN), 1645.0 and 1586.2 (s, C=N), 1457.0 and 1443.5 (s, C=C), 1377.0 (s, PAr), 1188.9 (m, P=C), 1088.6 (m), 1030.8 (m, PAr), 845.6 (m), 800.3 (m), 761.7 and 745.4 (s, PC), 672.1 (s), 546.7 (w), 520.0 cm1 (w). HR EIMS: calcd for C52H58IrN2P 934.3967, found 934.3930. m/z (%): 934 (14) [M]+, 803 (100) [M CNXy]+, 589 (60) [M DmpPH]+, 454 (16) [M DmpPH Cp*]+. Double dehydrohalogenation of [(5-Cp*)IrCl2(PH2Dmp)] (6c). Two

equivalents DBU (8.4 L, 5.6 mol) were added to an orange solution of [(5112

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

Cp*)IrCl2(PH2Dmp)] (6c; 20.9 mg, 2.8 mol) in CD2Cl2 (0.4 mL) at 10 C. After 5 min. at room temperature, the reaction mixture turned deep red of which
31P

NMR spectroscopy at 263 K showed unreacted [(5-Cp*)IrCl2(PH2Dmp)] at

81.5 (9.2%) together with phosphinidene [(5-Cp*)(CD2Cl2)Ir=PDmp)] (7c(DCM)) at 672.1 (11.5%), the mono-dehydrohalogenated product [(5Cp*)(Cl)Ir=PHDmp] at 105.8 (d, 1J(P,H) = 369.6 Hz, 50.7%), and two minor products at 17.4 (d, 1J(P,H) = 185.3 Hz, 19.3%), and 67.9 (d, 1J(P,H) = 212.6 Hz, 9.3%). Attempted synthesis of [(5-Cp*)(XyNC)Ir=PDmp] (3c). An orange = solution of [(5-Cp*)IrCl2(PH2Dmp)] (6c; 75 mg, 0.10 mmol) in CH2Cl2 (1 mL) was added to a clear, colorless solution of XyNC (13 mg, 0.10 mmol) and DBU (29.8 L, 0.20 mmol) in toluene (2 mL) at room temperature. After 3 h., a color changed to dark purple was observed and the phosphinidene [(5Cp*)(XyNC)Ir=PDmp] (3c) could be detected by
31P

NMR spectroscopy at

768.1 (50%) together with another product at 88.6 (d, 1J(P,H) = 202 Hz, 50%).

Acknowledgement. This work was supported by The Netherlands


Foundation for Chemical Sciences (CW) with financial aid from the Netherlands Organization for Scientific Research (NWO). Dr. M. Smoluch and J. W. H. Peeters (HR FAB-MS; University of Amsterdam) are acknowledged for measuring respectively HR EIMS and HR FABMS, and Mr. A.K. Mohamud is acknowledged for his contribution at the early stage of the study.

Supporting Information. Cartesian coordinates () and energies (au) of all


stationary points. Cif files with crystallographic data and copies of the NMR spectra of all novel compounds. This material is available free of charge via the Internet at http://pubs.acs.org.

References and Notes


[1] (a) P. B. Hitchcock, M. F. Lappert, W.P. Leung, J. Chem. Soc., Chem. Commun. 1987, 12821283. (b) R. Bohra, P. B. Hitchcock, M. F. Lappert, Leung, W.P.,

113

Chapter 4

Polyhedron 1989, 8, 1884. [2] (a) J. C. Slootweg, K. Lammertsma, In Science of Synthesis; B. M. Trost, F. Mathey, Eds.; Georg Thieme Verlag: Stuttgart, 2009; Vol. 42, pp 1536. (b) F. Mathey, Dalton Trans. 2007, 18611868. [3] D. W. Stephan, Angew. Chem. 2000, 112, 322338; Angew. Chem. Int. Ed. 2000, 39, 314329. [4] T. L. Breen, D. W. Stephan, J. Am. Chem. Soc. 1995, 117, 1191411921. [5] C. C. Cummins, R. R. Schrock, W. M. Davis, Angew. Chem. 1993, 105, 758761; Angew. Chem. Int. Ed. Engl. 1993, 32, 756759. [6] A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2002, 21, 31963202. [7] H. Aktas, J. C. Slootweg, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, J. Am. Chem. Soc. 2009, 131, 66666667. [8] R. Waterman, G. L. Hillhouse, J. Am. Chem. Soc. 2003, 125, 1335013351. [9] (a) T. L. Breen, D. W. Stephan, J. Am. Chem. Soc. 1996, 118, 42044205. (b) T. L. Breen, D. W. Stephan, Organometallics 1996, 15, 57295737. (c) G. Zhao, F. Basuli, U. J. Kilgore, H. Fan, H. Aneetha, J. C. Huffman, G. Wu, D. J. Mindiola, J. Am. Chem. Soc. 2006, 128, 1357513585. [10] R. R. Schrock, Angew. Chem. Int. Ed. 2006, 45, 37483759. [11] (a) D. S. Glueck, F. J. Hollander, R. G. Bergman, J. Am. Chem. Soc. 1989, 111, 27192721. (b) D. S. Glueck, J. Wu, F. J. Hollander, R. G. Bergman, J. Am. Chem. Soc. 1991, 113, 20412054. (c) A. A. Danopoulos, G. Wilkinson, T. K. N. Sweet, M. B. Hursthouse, J. Chem. Soc., Dalton Trans. 1996, 37713778. [12] A. T. Termaten, T. Nijbacker, A. W. Ehlers, M. Schakel, M. Lutz, A. L. Spek, M. L. McKee, K. Lammertsma, Chem. Eur. J. 2004, 10, 40634072. [13] One example is reported were a titanium phosphinidene reacts with three equivalents of isocyanide to afford a 2-(N,C)-phosphaazaallene, see: F. Basuli, L. A. Watson, J. C. Huffman, D. J. Mindiola, Dalton Trans. 2003, 42284229. [14] A. T. Termaten, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 35773582. [15] (a) K. Lammertsma, Top. Curr. Chem. 2003, 229, 95119. (b) K. Lammertsma, M. J. M. Vlaar, Eur. J. Org. Chem. 2002, 11271138. (c) F. Mathey, N. H. Tran Huy, A. Marinetti, Helv. Chim. Acta 2001, 84, 29382957. [16] (a) A. T. Termaten, H. Aktas, M. Schakel, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2003, 22, 18271834. (b) H. Aktas, J. C. Slootweg,

114

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Angew. Chem. 2009, 121, 3154 3157; Angew. Chem. Int. Ed. 2009, 48, 31083111. [17] Three platinum phosphaazaallene complexes with the general formula PtL2-[2(P,C)-Mes*PCNPh] were characterized spectroscopically, see: M.-A. David, J. B. Alexander, D. S. Glueck, G. P. A. Yap, L. M. Liable-Sands, A. L. Rheingold, Organometallics 1997, 16, 378383. [18] X-ray crystal structure determination of complexes 4a, 4b, 5. Intensities were measured at 150(2) K on a Nonius KappaCCD diffractometer with rotating anode (graphite monochromator, = 0.71073 ) up to a resolution of (sin /)max = 0.65 1. Integration was performed with EvalCCD[38] (compounds 4a and 4b) or HKL2000[39] (compound 5). The program SADABS[40] was used for absorption correction and scaling. The structures were solved with automated Patterson methods using the program DIRDIF-99[41] and refined with SHELXL-97[42] against F2 of all reflections. Non hydrogen atoms were refined with anisotropic

displacement parameters. All hydrogen atoms were located in difference Fourier maps and refined with a riding model. Geometry calculations and checking for higher symmetry was performed with the PLATON program.[43] Further details are given in Table 1. [19] Base-induced double dehydrohalogenation of [(5-Cp*)IrCl2(PH2Ph)] (6d) in the presence of ten equivalents of 2,6-xylyl isocyanide gave a mixture of unidentifiable products. [20] (a) J. Ho, Z. Hou, R. J. Drake, D. W. Stephan, Organometallics 1993, 12, 31453157. (b) A. Mahieu, A. Igau, J.-P. Majoral, Phosphorus, Sulfur Silicon 1995, 104, 235239. [21] (a) Z. Hou, D. W. Stephan, J. Am. Chem. Soc. 1992, 114, 1008810089. (b) Z. Hou, T. L. Breen, D. W. Stephan, Organometallics 1993, 12, 31583167. [22] E. W. Jandciu, P. Legzdins, W. S. McNeil, B. O. Patrick, K. M. Smith, Chem. Commun. 2000, 18091810; and references therein. [23] A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 22002208. [24] The, alternative, direct [1+2]-cycloaddition of HNC to [(5-Cp)IrNH] 1' to afford

2-carbodiimide complex 9' was not investigated in detail.


[25] F. L. Hirshfeld, Theoret. Chim. Acta 1977, 44, 129138. [26] ADF2005.01b and ADF2006.01, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands, http://www.scm.com/. [27] S. H. Vosko, L. Wilk, M. Nusair, Can. J. Phys. 1980, 58, 12001211. [28] A. D. Becke, Phys. Rev. A 1988, 38, 30983100.

115

Chapter 4

[29] J. P. Perdew, Phys. Rev. B 1986, 33, 88228824. [30] (a) L. Fan, L. Versluis, T. Ziegler, E. J. Baerends, W. Raveneck, Int. J. Quantum Chem. Quantum Chem. Symp. 1988, S22, 173181. (b) L. Versluis, T. Ziegler, J. Chem. Phys. 1988, 88, 322328. [31] E. van Lenthe, A. W. Ehlers, E. J. Baerends, J. Chem. Phys. 1999, 110, 89438953. [32] (a) G. Schreckenbach, T. Ziegler, J. Phys. Chem. 1995, 99, 606611. (b) G. Schreckenbach, T. Ziegler, J. Quantum Chem. 1997, 61, 899918. (c) S. K. Wolff, T. Ziegler, J. Chem. Phys. 1998, 109, 895905. (d) S. K. Wolff, T. Ziegler, E. van Lenthe, E. J. Baerends, J. Chem. Phys. 1999, 110, 76897698. [33] A. H. Cowley, J. E. Kilduff, T. H. Newman, M. Pakulski, M. J. Am. Chem. Soc. 1982, 104, 58205821. [34] E. Urnezius, J. D. Protasiewicz, Main Group Chem. 1996, 1, 369372. [35] Y. van den Winkel, H. M. M. Bastiaans, F. Bickelhaupt, J. Organomet. Chem. 1991, 405, 183194. [36] C. Overlnder, J. J. Tirre, M. Nieger, E. Niecke, C. Moser, S. Spirk, R. Pietschnig, Appl. Organometal. Chem. 2007, 21, 4648. [37] (a) I. Ugi, R. Meyr, Chem. Ber. 1960, 93, 239248. (b) R. Obrecht, R. Hermann, I. Ugi, Synthesis 1985, 400402. [38] A. J. M. Duisenberg, L. M. J. Kroon-Batenburg, A. M. M. Schreurs, J. Appl. Cryst. 2003, 36, 220229. [39] Z. Otwinowski, W. Minor, In Methods in Enzymology; C.W. Carter, Jr., R.M. Sweet, Eds.; Academic Press, 1997, Vol. 276, pp 307326. [40] Sheldrick, G. M. (1999). SADABS: Area-Detector Absorption Correction, v2.10, Universitt Gttingen, Germany. [41] P.T. Beurskens, G. Admiraal, G. Beurskens, W.P. Bosman, S. Garcia-Granda, R.O. Gould, J.M.M. Smits, C. Smykalla, (1999) The DIRDIF99 program system, Technical Report of the Crystallography Laboratory, University of Nijmegen, The

Netherlands. [42] G. M. Sheldrick, Acta Cryst. 2008, A64, 112122. [43] A. L. Spek, J. Appl. Cryst. 2003, 36, 713.

116

Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides

Table 1: Details of the X-ray crystal structure determinations. 4a formula Fw crystal colour crystal size [mm3] crystal system space group a [] b [] c [] [] V Z Dx [g/cm3] [mm1] abs. corr. abs. corr. range refl. meas. / unique param. / restraints R1/wR2 [I>2(I)] R1/wR2 [all refl.] S min/max [e/3] [3] C46H62IrN2P 866.15 yellow 0.24x0.21x0.10 monoclinic P21/c (no. 14) 14.3239(4) 16.0921(4) 21.5079(7) 122.067(1) 4201.2(2) 4 1.369 3.248 multi-scan[40] 0.45 0.72 69676 / 9645 469 / 0 0.0187 / 0.0358 0.0313 / 0.0391 1.043 -0.54 / 1.05 4b C37H44IrN2P 739.91 yellow 0.40x0.15x0.12 monoclinic P21/c (no. 14) 11.8147(2) 14.3753(4) 20.8906(4) 113.837(2) 3245.39(12) 4 1.514 4.191 multi-scan[40] 0.21 0.61 68819 / 7452 382 / 0 0.0193 / 0.0389 0.0277 / 0.0412 1.053 -0.92 / 1.35 5 C44H58IrN2P 838.09 yellow 0.27x0.04x0.04 monoclinic P21/c (no. 14) 11.8029(1) 16.4042(1) 20.8175(2) 96.7501(3) 4002.68(6) 4 1.391 3.407 multi-scan[40] 0.47 0.87 82096 / 9159 449 / 0 0.0222 / 0.0426 0.0334 / 0.0455 1.043 -0.56 / 0.99

117

Chapter 4

118

Chapter 5
3-Diphosphavinylcarbene:
A P2 Analogue of the Dtz Intermediate

Halil Aktas, J. Chris Slootweg, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, and Koop Lammertsma*,

Department of Chemistry and Pharmaceutical Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands, and Bijvoet Center for Biomolecular Research, Crystal Structural Chemistry, Utrecht University, Padualaan 8, 3584 CH Utrecht, The Netherlands

Angew. Chem. 2009, 121, 31543157; Angew. Chem. Int. Ed. 2009, 48, 31083111

Abstract: The reaction of in situ generated phosphinidenes with phosphaalkynes is a facile route to the new 3-diphosphavinylcarbene 5, which shows facile ligand exchange reactions or undergoes an unprecedented rearrangement that involves phosphinidene complex 8 and

3-

phosphaalkenylphosphinidene 9, the 1,3-isomer of 5.

Chapter 5

5.1

Introduction

Olefin metathesis[1] constitutes a powerful tool for the construction of a plethora of unsaturated building blocks, pharmaceuticals and advanced materials. The principle steps involve a transition-metal carbene complex (A) that undergoes a [2+2] cycloaddition/cycloreversion protocol with alkenes according to the Chauvin

mechanism.[2] For the widely used Grubbs catalysts, however, support for the key intermediate of this process, the four-membered metallacyclobutane, relies on theoretical studies[3] and low-

temperature NMR spectroscopy.[4,5] The unsaturated analogues provide more insight. For example, reaction of alkynes with the second-generation Grubbs catalyst gives stable 3-vinylcarbene complexes (1, Mes = 2,4,6-Me3C6H2) that form after rearrangement of the initial ruthenium cyclobutenes (B).[6,7] These puckered metal 3vinylcarbene complexes[8] (C) are highly potent and represent the key intermediate in enyne metathesis,[9] alkyne polymerization (via D),[10] and the versatile Fischer carbene mediated Dtz

(benzannulation) reaction (which involves complexes such as 2).[11,12]

In our quest for novel reactivity patterns and in light of the diagonal relationship between carbon and phosphorus,[13] we here report on a diphosphorus analogue of 3-vinylcarbene complex C (P2-C).[14] Our approach involves the reaction of in situ generated phosphinidene
120

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

[M=PMes*] (Mes* = (M = 2,4,6-tBu3C6H2;[15] M = (6-pCy)Ru (pCy = para-cymene), (5-Cp*)Ir (Cp* = C5Me5)), the phosphorus analogues of carbenes, with phosphaalkynes (RCP; R = Mes*, tBu).

5.2

Synthesis of 3-diphosphavinylcarbenes

Treatment of primary phosphine complexes [Cl2M(PH2Mes*)] 3 with two equivalents of 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) in the presence of one equivalent Mes*CP afforded

3-

diphosphavinylcarbenes 5a (P2-C) as the sole products in 89% (Ru) and 76% (Ir) yield after crystallization (Scheme 1). The synthesis of 5a is remarkably selective, as evidenced by the single AB spin system in the
31P

NMR spectrum (Ru-5a: 31P 52.1 and 4.8 ppm, 1JP,P = 431.4 Hz;

Ir-5a: 31P 55.7 and 21.1 ppm, 1JP,P = 394.0 Hz); the large P,P coupling indicates the presence of a PP bond. The highly deshielded resonances in the ppm (Ir-5a) are
13C

NMR spectrum at = 306.9 (Ru-5a) and 260.3 for metal alkylidenes.[6,16] They

diagnostic

unambiguously support the carbenoid nature of the ring carbon atom and exclude the diphosphametallacyclobutene (P2-B)

conformation.

Scheme 1. Synthesis and reactivity of 3-diphosphavinylcarbene 5.

121

Chapter 5

X-ray crystal structure determination[17] of dark-green Ru-5a (Figure 1) and deep red crystals of Ir-5a (Figure 2) revealed unequivocally the puckered structure of the unique 3-diphosphavinylcarbene ligand. Complex Ru-5a displays short metal-alkylidene and P2C19 bonds (1.928(2) and 1.879(2) , respectively) and typical single metalP

(2.4339(5) and 2.4478(6) ), and PP bonds 2.1771(8) ).

Figure 1. Displacement ellipsoid plot of Ru-5a with ellipsoids drawn at the 50% probability level. Hydrogen atoms and toluene solvent are omitted for clarity. Selected bond lengths [], angles [] and torsion angles []: Ru1P1 2.4339(5), Ru1P2 2.4478(6), Ru1C19 1.928(2), Ru1pCy(cg) 1.7473(7), P1C1 1.879(2), P1P2 2.1771(8), P2C19 1.759(2), C19C20 1.480(3); P2P1Ru1 63.84(2), P1P2C19 88.26(7), Ru1C19 P2 83.06(8); P1P2Ru1C19 111.06(9).

Whereas no intermediates could be detected by variabletemperature NMR spectroscopy during the formation of Ru-5a, selective monodehydrohalogenation of iridium precursor 3 occurred at 80 C to give syn/anti-Ir-4 (31P NMR = 119.5 (1JP,H = 398.9 Hz) and 128.3 ppm (1JP,H = 369.6 Hz)) in a 5:1 ratio. At this temperature, Ir-4 is unreactive towards Mes*CP, suggesting that the second

dehydrohalogenation to afford the transient [M=PMes*] takes place prior to phosphaalkyne addition. At 60 C the gradual formation of Ir-5a was observed.
122

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

Figure 2. Displacement ellipsoid plot of Ir-5a with ellipsoids drawn at the 50% probability level. Only the major disorder component is drawn. Hydrogen atoms are omitted for clarity. Selected bond lengths [], angles [] and torsion angles [] for the major disorder component of Ir-5a: Ir1P1 2.4995(10), Ir1P2 2.3960(10), Ir1C1 1.943(4), P1C1 1.756(4), P1P2 2.1961(13), P2C20 1.920(3), C1C2 1.461(5); Ir1P2P1 65.79(4), P2P1C1 88.62(15), Ir1C1P1 84.86(18); Ir1C1P1P2 52.59(12), P2P1Ir1C1 114.73(18).

5.3

Synthesis of 1,3-diphospha-3H-indenes 3 with the less sterically congested

Surprisingly, reaction of

phosphaalkyne tBuCP using two equivalents of DBU resulted, at room temperature, in yellow crystalline 6b (80% (Ru), 87% (Ir); Scheme 1) instead of the expected carbene 5b, as indicated by a different set of signals in the
31P

NMR spectrum. The smaller P,P couplings are

characteristic (Ru-6b: = 1.5 and 92.1 ppm, 2JP,P = 20.7 Hz; Ir-6b: = 32.7 and 113.0 ppm,
2J P,P

= 24.5 Hz), as is the absence of a


13C

carbenoid resonance in the

NMR spectrum. The molecular

structure of Ru-6b, established by single-crystal X-ray structure determination,[17] exhibits a number of fascinating features. First and foremost, 6b contains an unprecedented 1,3-diphospha-3H-indene moiety,[18] bearing a dearomatized Mes*-substituent (Figure 3). The striking resemblance of this stable P2-entity with the benzannulation (Wheland) intermediate of the Dtz reaction is evident.[12c] Secondly,
123

Chapter 5

4-coordination of the P2C2 moiety to the ruthenium center is favored


over coordination of the all-carbon butadiene fragment.[19] As a result, the P2C2 unit displays delocalization, as evidenced by the similar P C bond lengths (1.7787(16)1.7926(15) ) and an alternating bond pattern for the dearomatized Mes* ring.

Figure 3. Displacement ellipsoid plot of Ru-6b with ellipsoids drawn at the 30% probability level. Hydrogen atoms are omitted for clarity. Selected bond lengths [] and torsion angles []: Ru1P1 2.3664(4), Ru1P2 2.3863(4), Ru1C1 2.2624(14), Ru1 C19 2.2055(14), Ru1pCy(cg) 1.7473(7), P1C1 1.7926(15), P1C19 1.7802(15), P2C6 1.8777(15), P2C19 1.7787(16), C1C2 1.486(2), C1C6 1.547(2), C2C3 1.351(2), C3 C4 1.461(2), C4C5 1.341(2), C5C6 1.518(2); C1P1C19P2 2.11(9).

5.4

Mechanism

Monitoring the reaction of ruthenium precursor 3 with tBuCP by variable-temperature NMR spectroscopy showed full conversion at 80 C to dark-green Ru-5b (31P NMR: = 2.0 and 35.7 ppm, 1JP,P = 400.3 Hz) and above 40 C its subsequent rearrangement to yellow Ru-6b. Monitoring the corresponding reaction of the iridium congener of 3 at 80 C showed, besides the monodehydrohalogenation to syn/anti-Ir-4 (ca. 3:1 ratio), the formation of two new P2 species (10:1

124

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

ratio) with sizable P,P and P,H couplings (major isomer,

31P

NMR =

358.7 and 138.9 ppm, 1JP,P = 164.0 Hz, 1JP,H = 387.8 Hz). We ascribe these products to the syn and anti adducts of Ir-7b (Figure 4), which suggests that in this case the addition of the less congested phosphaalkyne tBuCP to Ir-4 is plausible.

Figure 4. Intermediate Ir-7b and model structure Ir-7, calculated at the B3PW91/6 31G(d,p) (LANL2DZ for Ir) level of theory. Selected bond lengths [] and a torsion angle [] of Ir-7: IrP1 2.304, IrC1 1.992, P1P2 2.214, P2C2 1.707; P1P2C1Ir 4.86.

The intermediacy of Ir-7b is supported by the calculated NMR parameters[20] of the unsubstituted model structure Ir-7 (31P NMR = 352.6 and 110.8 ppm, 1JP,P = 256.9 Hz; Figure 4). At 40 C, Ir-7b underwent the second base-induced dehydrohalogenation to yield Ir-5b (31P NMR = 3.2 and 24.4 ppm, 1JP,P = 371.6 Hz), which upon warming to room temperature rearranged into Wheland product Ir6b, as was evident by the color change from red to yellow. The remarkable conversion of 5b into 6b was examined by B3PW91/631G(d,p) calculations on model structures containing H instead of tBu substituents (labeled 5, 6, etc.; Figure 5). No simple direct pathway was found, but instead one that involves

phosphinidene complex 8 and 3-phosphaalkenyl phosphinidene 9, the 1,3-isomer of 5. Interestingly, the planar diphosphametallacyclobutene conformer (P2-B) is not an energy minimum, but corresponds to the transition-state structure for inversion of puckered 5 (E = 24.4 (Ru), 22.1 (Ir) kcalmol1). Isomerization of carbene 5 to
125

Chapter 5

the less stable phosphinidene 8 proceeds by PP bond cleavage (Ru: E = 19.8, E = 20.6 kcalmol1). This step is then followed by facile rotation of the 2-coordinated phosphaalkyne ligand and PC bond formation to give regioisomer 9 (Ru: E = 22.5, E = 4.8 kcalmol1). Finally, the subsequent electrophilic attack of P2 affords the unique Wheland product 6 (Ru: E = 0.6, E = 17.7 kcalmol1), of which the experimental analogues, Ru-6b and Ir-6b, can be isolated. 5.5 Reactivity

For an initial assessment of the reactivity of the novel 3diphosphavinylcarbenes 5, we treated the ruthenacycle Ru-5a with one equivalent of tBuCP and observed the fast and selective formation of bicyclic Ru-6b (t = 18 min at 0 C; 93% yield according to
31P

NMR spectroscopy) together with elimination of Mes*CP

(Scheme 2).

Scheme 2. Reactivity of Ru-5a towards phosphines.

Addition of PPh3 afforded instead terminal phosphinidene complex [(6-pCy)(Ph3P)Ru=Mes*] Ru-10[14a] (t = 100 min at 0 C; quantitative yield). Both conversions follow an unprecedented pathway for 3vinylcarbenes for which an associative ligand exchange at the phosphinidene stage (8) is proposed, as the dissociative pathway (Ru-8 C6H6Ru=PPh + HCP) corresponds to an uphill process (E = 28.8 kcalmol1).
126

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

Figure 5. Relative B3PW91/631G(d,p) (LANL2DZ for Ru and Ir) energies (ZPE corrected, in kcalmol1) for the rearrangement of 5

into 6. The relative energies for the 5-CpIr derivatives are given in parentheses. Selected bond lengths [] for Ru-5: RuP1 2.443,

RuP2 2.571, RuC7 1.894, P1P2 2.180, P2C7 1.752; TS1: P1P2 2.876, P2C7 1.631; Ru-8: RuP1 2.177, RuP2 2.509, RuC7 2.105,

P2C7 1.617; TS2: P1C7 2.439, P2C7 1.630; Ru-9: RuP1 2.432, RuP2 2.194, RuC7 2.244, P1C7 1.784, P2C7 1.772; TS3: RuP1

2.404, P2C6 2.106; Ru-6: RuP1 2.398, RuP2 2.437, RuC1 2.214, RuC7 2.199, P1C1 1.804, P1C7 1.787, P2C6 1.887, P2C7

1.783.

127

Chapter 5

5.6

Conclusion

Phosphorus substitution of metal-bound 3-vinylcarbenes has a marked influence on their reactivity. Depending on the substituent pattern, the readily obtained 3-diphosphavinylcarbene complexes 5 show facile ligand exchange reactions or undergo an

unprecedented rearrangement to the stable Wheland product 6. 5.7 Computational Section

Density functional theory calculations (B3PW91) were performed with the Gaussian03 (Revision C.02) suite of programs,[20] using the LANL2DZ basis set and pseudopotentials for ruthenium and iridium and the 631G(d,p) basis set for all other atoms. The nature of each structure was confirmed by frequency calculations. Intrinsic reaction coordinate (IRC) calculations were performed to ascertain the connection between reactant and product. NMR Calculations. Calculated coordinates were used to calculate the
31P

chemical shielding at the B3PW91/6311++G(2d,p)//B3PW91/631G(d,p level.[20] These values are relative to a bare phosphorus nucleus and can be converted to chemical shifts relative to an appropriate reference system, for which we used trimethyl phosphine (calcd = 367.8, exp = 62.0 ppm) to obtain the relationship: calcd(intermediate) = 305.8 ppm

calcd(intermediate). All Cartesian coordinates in Angstroms, energies in


kcal.mol1 and chemical shielding for calculated compounds are available free of charge via the Internet at htttp://www. angewandte.org.

5.8

Experimental Section

General. All experiments and manipulations were performed under an atmosphere of dry nitrogen or argon with rigorous exclusion of air and moisture using flamedried glassware. All Variable Temperature NMR measurements were performed under an atmosphere of dry nitrogen in Wilmad Screw Cap NMR Tubes (Aldrich). 1H,
13C

and

31P

NMR spectra were

recorded on a Bruker Avance 250 (respectively 250.13, 62.90 and 101.25 128

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

MHz) or on a Bruker Avance 400 (respectively 400.13, 100.64 and 162.06 MHz) spectrometer. 1H NMR spectra were internally referenced to CDHCl2 ( 5.25), or C6D5H ( 7.16),
13C

NMR spectra to C6D6 ( 128.06) and

31P

NMR spectra

externally to 85% H3PO4. IR Spectra were recorded on a Mattson6030 Galaxy FTIR spectrophotometer, highresolution mass spectra (HRMS) on a Finnigan Mat 900 spectrometer operating at an ionization potential of 70 eV, and fast atom bombardment (FAB) mass spectrometry was carried out using a JEOL JMS SX/SX 102A foursector mass spectrometer (70 eV). Melting points were measured on samples in sealed capillaries on a Stuart Scientific SMP3 melting point apparatus and are uncorrected. UV spectra were recorded with a Varian CARY 300 1 Bio UV/Visible spectrophotometer. Reagents. Solvents were distilled from sodium (toluene), CaCl2 (CH2Cl2), or LiAlH4 (pentanes, diethyl ether) and kept under an atmosphere of dry nitrogen. Deuterated solvents were purchased from Aldrich, dried over 4 molecular sieves (CD2Cl2, C6D6), and kept under an atmosphere of dry nitrogen. All solid starting materials were dried in vacuo. [(6-

pCy)RuCl2(PH2Mes*)] Ru-3,[14a] [(5-Cp*)IrCl2(PH2Mes*)] Ir-3,[22] tBuCP,[23] and Mes*CP,[24] were prepared according to literature procedures.

Experimental Procedures. Ru-5a. An orange solution of [(6-pCy)RuCl2(PH2Mes*)] (Ru-3; 0.140 g, 0.239 mmol) in CH2Cl2 (2 mL) was added dropwise to a clear mixture of DBU (71.4

L, 0.477 mmol) and Mes*CP (0.070 g, 0.24 mmol) in toluene (5 mL) at


78 C, which immediately changed the colour of the reaction mixture to deep green. The mixture was allowed to warm up to room temperature and was stirred for an additional 30 min. All solvents were evaporated in vacuo and the residue was extracted into pentanes (3x10 mL). After removal of the pentanes, Ru-5a was obtained as a dark green solid after crystallization from diethyl ether at 20 C (0.170 g, 89%). Suitable dark green crystals for X-ray crystallography were obtained from a saturated toluene solution at 20 C. M.p. 157 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 0.67 (d, 3J(H,H) = 7.5 Hz, 3H; pCy-CH(CH3)2), 0.76 (d, 3J(H,H) = 7.5 Hz, 3H; pCy-CH(CH3)2), 1.04 (s, 129

Chapter 5

9H; o-C(CH3)3), 1.34 (s, 3H; pCy-CH3), 1.39 (s, 9H; p-C(CH3)3), 1.40 (s, 9H; pC(CH3)3), 1.79 (s, 9H; o-C(CH3)3), 1.79 (m, 1H; CH(CH3)2), 1.98 (s, 18H; oC(CH3)3), 4.71, 4.73, 4.86, 4.88 (AB type, 3J(H,H) = 5.9 Hz, 2H; pCyH), 4.82, 4.83, 4.93, 4.94 (AB type, 3J(H,H) = 5.9 Hz, 2H; pCyH), 7.36 (d, 4J(H,P) = 1.8 Hz, 1H; mMes*H), 7.44 (bs, 3H; m-Mes*H).
13C{1H}

NMR (100.64 MHz, C6D6, 300 K): 18.8

(s; pCy-CH3), 23.1 (s; pCy-CH(CH3)2), 23.9 (s; pCy-CH(CH3)2), 30.3 (s; pCyCH(CH3)2), 31.6 (s; p-C(CH3)3), 31.7 (s; p-C(CH3)3), 31.9 (d, 4J(C,P) = 2.0 Hz; oC(CH3)3), 33.3 (s; o-C(CH3)3), 33.4 (s; o-C(CH3)3), 34.3 (br t; o-C(CH3)3), 34.7 (s; p-C(CH3)3), 34.8 (s; p-C(CH3)3), 37.3 (s; o-C(CH3)3), 38.4 (s; o-C(CH3)3), 39.3 (s; o-C(CH3)3), 82.8 (s; C6H4), 85.9 (s; C6H4), 92.4 (s; C6H4), 93.1 (s; C6H4), 102.3 (s; CH3-C6H4), 112.6 (s; (CH3)2CH-C6H4), 120.2 (s; m-Mes*), 121.3 (s; m-Mes*), 121.7 (s; m-Mes*), 137.2 (s; o-Mes*), 140.0 (s; o-Mes*), 146.6 (s; p-Mes*), 146.8 (s; pMes*), 148.2 (d, 1J(C,P) = 113.4 Hz; ipso-Mes*), 154.8 (s; C-ipso-Mes*), 160.0 (d,
3J(C,P)

= 5.1 Hz, o-Mes*), 306.9 (CD2Cl2; dd, 1J(C,P) = 97.9 Hz, 2J(C,P) = 16.6 Hz,

RuC). 31P NMR (101.25 MHz, C6D6, 300 K): 52.1 (d, 1J(P,P) = 431.4 Hz), 4.8 (d,
1J(P,P)

= 431.4 Hz). IR (KBr, cm1): v = 3098 (w), 2963 (s), 2948 (s), 2907 (s), 2865

(s), 1593 (s), 1476, 1462, and 1445 (s), 1391, and 1361 (s; PAr), 1238 (s), 1209 (s), 1117 (w), 1091 (w), 1029 (m; PAr), 951 (w), 925 (w), 874 (m), 841 (w), 800 (w), 779 and 745 (s; PC), 685 (w), 662 (w), 583 (w). HR FAB-MS: calcd for C47H73P2Ru (M + H): 801.4245, found: 801.4221; m/z (%): 801 (8) [M + H]+, 743 (18) [M tBu]+, 609 (45) [M tBu pCy]+, 493 (60) [M P2Mes*]+. UV/Vis (pentanes): max() = 211 (59783), 236 (38414), 271 (23674), 319 (17428), 344 (13606), 360 nm (8961). Ir-5a. Ir-5a was prepared in a similar fashion to that described for Ru-5a by adding [(5-Cp*)IrCl2(PH2Mes*)] (Ir-3; 0.280 g, 0.414 mmol) to a toluene solution of Mes*CP (0.122 g, 0.424 mmol) and DBU (126 L, 0.828 mmol) to afford Ir-5a as deep red blocks (0.2814 g, 76%). M.p. 172174 C, 179 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 1.06 (s, 9H; o-C(CH3)3), 1.23 (s, 15H; C5(CH3)5), 1.37 (s, 9H; p-C(CH3)3), 1.40 (s, 9H; p-C(CH3)3), 1.75 (s, 9H; oC(CH3)3), 1.80 (s, 9H; o-C(CH3)3), 2.12 (s, 9H; o-C(CH3)3), 7.38, 7.38, 7.39, 7.40 (AB type, 4J(H,P) = 1.9 Hz, 2H; m-Mes*H), 7.45 (bs, 1H; m-Mes*H), 7.55 (bs, 1H; m-Mes*H). 130
13C{1H}

NMR (100.64 MHz, C6D6, 300 K): 9.5 (s; C5(CH3)5), 31.7 (s;

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

p-C(CH3)3), 31.8 (s; p-C(CH3)3), 32.3 (d, 5J(C,P) = 2.5 Hz; o-C(CH3)3), 32.4 (d,
5J(C,P)

= 4.1 Hz; o-C(CH3)3), 32.8 (s; o-C(CH3)3), 33.0 (s; o-C(CH3)3), 34.7 (s; p-

C(CH3)3), 34.8 (s; p-C(CH3)3), 35.2 (s; o-C(CH3)3), 35.5 (s; o-C(CH3)3), 37.7 (s, oC(CH3)3), 38.0 (s, o-C(CH3)3), 38.2 (br s, o-C(CH3)3), 40.1 (br d, 3J(C,P) = 3.9 Hz; o-C(CH3)3), 95.7 (s; C5(CH3)5), 119.8 (s; m-Mes*), 121.5 (s; m-Mes*), 122.2 (s; mMes*), 134.0 (s; o-Mes*), 141.4 (d, 3J(C,P) = 4.5 Hz; o-Mes*), 146.4 (s; p-Mes*), 146.9 (dd, 1J(C,P) = 121.7 Hz, 2J(C,P) = 8.8 Hz; P-ipso-Mes*), 147.3 (s; p-Mes*), 153.1 (br s; o-Mes*), 156.0 (br s; o-Mes*), 156.3 (br d, 3J(C,P) = 25.2 Hz; ipsoMes*), 260.3 (dd, 1J(C,P) = 90.3 Hz, 2J(C,P) = 40.8 Hz; IrC). 31P NMR (101.25 MHz, C6D6, 300 K): 55.7 (d, 1J(P,P) = 394.0 Hz), 21.1 (d, 1J(P,P) = 394.0 Hz). IR (KBr, cm1): v = 3081 (w), 2953 (s), 2901 (s), 2864 (s), 1580 (s), 1477, and 1462 (s), 1382, and 1360 (s; PAr), 1237, and 1200 (s), 1119 (w), 1070 (w), 1024 (m; PAr), 957 (w), 924 (w), 899 (w), 871 (s), 804 (w), 767 and 743 (s; PC), 706 (w), 654 (w), 640 (w), 604 (w), 581 (w). HR EI-MS (70 eV): calcd for C47H73IrP2: 892.4817, found: 892.4819; m/z (%): 892 (20) [M]+, 835 (12) [M tBu]+, 585 (44) [M P2Mes*]+. UV/Vis (pentanes): max() = 211 (90450), 257 (52785), 344 (20165), 434 nm (6960). Ru-6b. Ru-6b was prepared in a similar fashion to that described for Ru-5a by adding [(6-pCy)RuCl2(PH2Mes*)] (Ru-3; 0.117 g, 0.200 mmol) to a toluene solution of tBuCP (250 L, 0.250 mmol, 1M solution in toluene) and DBU (60.0 L, 0.400 mmol) to afford Ru-6b as large yellow blocks after crystallization from a saturated toluene solution at 20 C (0.098 g, 80%). M.p. 149 C (dec). 1H NMR (400.13 MHz, C6D6, 300 K): 0.92 (s, 9H; PC-C(CH3)3), 1.18 (d, 3J(H,H) = 6.9 Hz, 3H; CH(CH3)2), 1.28 (s, 9H; p-C(CH3)3), 1.32 (s, 9H; P2CC(CH3)3), 1.33 (d, 3J(H,H) = 6.9 Hz, 3H; CH(CH3)2), 1.45 (s, 9 H; o-C(CH3)3), 2.38 (s, 3H; pCy-CH3), 2.77 (septet, 3J(H,H) = 6.9 Hz, 1H; CH(CH3)2), 4.30, 4.32, 4.37, 4.38 (AB type, 3J(H,H) = 5.8 Hz, 2H; C6H4), 4.74, 4.76, 4.79, 4.80 (AB type,
3J(H,H) 13C{1H}

= 5.8 Hz, 2H; C6H4), 5.875.90 (m, 3J(H,P) = 10.3 Hz, 2H; m-Mes*H). NMR (100.64 MHz, C6D6, 300 K): 21.6 (d, 3J(C,P) = 4.9 Hz, pCy-CH3),

24.1 (d, 4J(C,P) = 3.8 Hz; CH(CH3)2), 24.4 (d, 4J(C,P) = 2.2 Hz; CH(CH3)2), 26.6 (dd, 3J(C,P) = 6.7 Hz, 4J(C,P) = 2.5 Hz; PCC(CH3)3), 30.4 (s; p-C(CH3)3), 32.2 (d, J(C,P) = 12.5 Hz; o-C(CH3)3), 32.8 (s; CH(CH3)2), 34.4 (t, 3J(C,P) = 7.8 Hz; 131

Chapter 5

P2CC(CH3)2), 34.6 (s; p-C(CH3)3), 36.9 (dd, 2J(C,P) = 11.0 Hz, 2J(C,P) = 14.3 Hz; P2CC(CH3)2), 37.7 (d, 3J(C,P) = 2.2 Hz; o-C(CH3)3), 42.2 (d, 3J(C,P) = 5.4 Hz; PCC(CH3)3), 61.4 (dd, 1J(C,P) = 63.3 Hz, 2J(C,P) = 2.3 Hz; PC-C(CH3)3), 75.9 (dd,
1J(C,P)

= 18.5 Hz, 2J(C,P) = 4.6 Hz; PCC(t-Bu)P), 84.1 (s; C6H4), 84.1 (s; C6H4),

84.4 (s; C6H4), 88.8 (s; C6H4), 104.0 (s, CH3C6H4), 116.5 (s, (CH3)2CHC6H4), 117.0 (dd, 3J(C,P) = 8.3 Hz, 4J(C,P) = 2.5 Hz; m-Mes*), 118.0 (dd, 1J(C,P) = 85.7 Hz, 1J(C,P) = 73.8 Hz; P2C), 121.8 (d, 2J(C,P) = 21.2 Hz; m-Mes*), 144.4 (d,
3J(C,P)

= 10.0 Hz, p-Mes*), 154.2 (dd, 2J(C,P) = 14.5 Hz, 3J(C,P) = 2.9 Hz, o-

Mes*). 31P NMR (101.25 MHz, C6D6, 300 K): 1.5 (d, 2J(P,P) = 20.7 Hz; CPCtBu), 92.1 (dd, 2J(P,P) = 20.7 Hz, 3J(P,H) = 11.6 Hz; tBuCPCtBu). IR (KBr, cm1): v = 2956 (s), 2896 (m), 2862 (w), 1474, 1453, and 1441 (s), 1384, and 1356 (s; PAr), 1239 (m), 1192 (m), 1149 (w), 1079 (w), 1049, and 1029 (m; PAr), 829 (m), 801 (m), 684 (w). HR EI-MS (70 eV): calcd for C43H63P2Ru (M tBu): 555.1877, found: 555.1889; m/z (%): 555 (100) [M tBu]+. UV/Vis (pentanes): max() = 210 (34273), 244 (30424), 379 (3224). Ir-6b. Ir-6b was prepared in a similar fashion to that described for Ru-5a by adding [(5-Cp*)IrCl2(PH2Mes*)] (Ir-3; 0.169 g, 0.250 mmol) to a toluene solution of tBuCP (312 L, 0.312 mmol, 1M solution in toluene) and DBU (74.8 L, 0.500 mmol) to afford Ir-6b as yellow needles after crystallization from toluene (0.153 g, 87%). M.p. 173174 C. 1H NMR (400.13 MHz, C6D6, 300 K): 1.02 (s, 9H; PCC(CH3)3), 1.22 (s, 9H; p-C(CH3)3), 1.38 (s, 9H; P2CC(CH3)3), 1.38 (s, 9H; o-C(CH3)3), 1.17 (s, 15H; C5(CH3)5), 5.57 (dd, 3J(H,P) = 11.3 Hz,
4J(H,H)

= 1.3 Hz, 1H; Mes*H), 6.35 (d, 4J(H,H) = 1.3 Hz, 1H; Mes*H).

13C{1H}

NMR

(100.64 MHz, C6D6, 300 K): 10.3 (s; C5(CH3)5), 26.8 (dd, 3J(C,P) = 6.1 Hz,
4J(C,P)

= 2.0 Hz; PCC(CH3)3), 29.9 (s; p-C(CH3)3), 31.5 (d, 4J(C,P) = 12.4 Hz; o-

C(CH3)3), 32.9 (dd, 3J(C,P) = 8.3 Hz, 3J(C,P) = 6.7 Hz; P2CC(CH3)2), 34.8 (s; pC(CH3)3), 35.9 (dd, 2J(C,P) = 13.1 Hz, 2J(C,P) = 10.7 Hz; P2CC(CH3)2), 37.1 (s; oC(CH3)3), 46.7 (d, 2J(C,P) = 4.1 Hz; PC-C(CH3)3), 47.9 (d, 1J(C,P) = 61.6 Hz, PCC(CH3)3), 75.8 (dd, 1J(C,P) = 14.8 Hz, 2J(C,P) = 4.7 Hz; PCC(t-Bu)P), 92.9 (s; C5(CH3)5), 101.8 (dd, 1J(C,P) = 81.9 Hz, 1J(C,P) = 72.6 Hz; P2C), 117.1 (dd,
3J(C,P)

= 8.0 Hz, 4J(C,P) = 1.9 Hz; m-Mes*), 122.8 (d, 2J(C,P) = 20.0 Hz; m-Mes*),

144.1 (d, 3J(C,P) = 11.0 Hz, p-Mes*), 154.2 (dd, 2J(C,P) = 13.2 Hz, 2J(C,P) = 3.6 132

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

Hz, o-Mes*).

31P

NMR (101.25 MHz, C6D6, 300 K): 32.7 (d, 2J(P,P) = 24.5 Hz;

CPCtBu), 113.0 (dd, 2J(P,P) = 24.5 Hz, 3J(P,H) = 11.2 Hz; tBuCPCtBu). IR (KBr, cm1): v = 2950 (vs), 2897 (s), 2859 (s), 1640 (w), and 1622 (w), 1462 (s), 1379, 1366, 1356 (s; PAr), 1239 (m), 1196 (m), 1114 (w), 1066 (w), 1019 (s), and 990 (w; PAr), 835 (m), 797 (m), 689 (w), 633 (w), 588 (w). HR EI-MS (70 eV): calcd for C29H44P2Ir (M tBu): 647.2551, found: 647.2573; m/z (%): 647 (100) [M tBu]+. UV/Vis (pentanes): max() = 211 (60539), 251 (32109), 284 (17354), 379 (4027).

VT-NMR investigations of the formation of 5. Ru-5a. 2 eq. of DBU (5.6 L, 37.4 mol) was slowly added to an orange mol) and Mes*

solution of [(6-pCy)RuCl2(PH2Mes*)] (Ru-3; 11.0 mg, 18.7 CP (5.4 mg, 18.8

mol) in CD2Cl2 (0.5 mL) at 80 C, which immediately

changed the colour of the reaction mixture to deep green. At this temperature, the quantitative conversion into Ru-5a was observed by NMR spectroscopy and no intermediates could be detected. Ir-5a. 2 eq. of DBU (4.8 L, 32.0 mol) was slowly added to an orange mol) and Mes*CP
31P

solution of [(5-Cp*)IrCl2(PH2Mes*)] (Ir-3; 10.8 mg, 16.0

(4.6 mg, 16.0 mol) in CD2Cl2 (0.5 mL) at 80 C, which immediately changed the colour of the reaction mixture to deep red. At this temperature, selective (mono-)dehydrohalogenation of Ir-3 was observed by 31P NMR spectroscopy to give two isomers of intermediate Ir-4 quantitatively (5:1 ratio). This assignment was confirmed by repeating the experiment using only 1 eq. of DBU with or without 1 eq. of phosphaalkyne present.
31P

NMR (101.25 MHz,

CD2Cl2, 193 K): 128.3 (d, 1J(P,H) = 369.6 Hz, 17%), 119.5 (d, 1J(P,H) = 398.9 Hz, 83%). 1H NMR (400.13 MHz, CD2Cl2, 193 K): 8.57 (d, 1J(H,P) = 369.6 Hz, 17%), 9.60 (d, 1J(H,P) = 399.2 Hz, 83%). Warming up the reaction mixture to 60 C showed the gradual formation of Ir-5a that became the sole product at room temperature.

VT-NMR investigations of the formation of 6. Ru-6b. 2 eq. of DBU (6.0 L, 40.0 mol) was slowly added to an orange

solution of [(6-pCy)RuCl2(PH2Mes*)] (Ru-3; 11.7 mg, 20.0 mol) and tBuCP 133

Chapter 5

(20.0 mol, 20.0 L of a 1M solution in toluene) in CD2Cl2 (0.5 mL) at 80 C, which immediately changed the colour of the reaction mixture to deep green. At this temperature, the quantitative conversion into Ru-5b was observed by
31P

NMR spectroscopy and no intermediates could be

detected. 31P NMR (101.25 MHz, CD2Cl2, 193 K): 2.0 (d, 1J(P,P) = 400.3 Hz), 35.7 (d, 1J(P,P) = 400.3 Hz). Warming up the reaction mixture to 40 C showed the gradual formation of Ru-6b that became the main product above 0 C (> 81%). Ir-6b. 2 eq. of DBU (6.2 L, 41.4 mol) was slowly added to an orange mol) and tBuCP

solution of [(5-Cp*)IrCl2(PH2Mes*)] (Ir-3; 14.0 mg, 20.7

(20.7 mol, 20.7 L of a 1M solution in toluene) in CD2Cl2 (0.5 mL) at 80 C, which immediately changed the colour of the reaction mixture to deep red. At this temperature, the two isomers of intermediate Ir-4 were observed by
31P

NMR spectroscopy as well as the two isomers of intermediate Ir-7b (10:1

ratio), which are generated via the subsequent reaction with tBuCP, along with unreacted starting material.
31P{1H}

NMR (101.25 MHz, CD2Cl2, 193 K):

358.7 (d, 1J(P,P) = 164.0 Hz, 28.5%; Ir-7b-major), 342.4 (d, 1J(P,P) = 185.4 Hz, 2.8%; Ir-7b-minor), 128.7 (s, 1.5%; Ir-4-minor), 119.5 (s, 4.8%; Ir-4-major), 54.0 (t,
1J(P,H)

= 392.8 Hz, 15.6%; Ir-3), 70.8 (s, 15.5%; tBuCP), 138.9 (d, 1J(P,P) =

164.0 Hz (1J(P,H) = 387.8 Hz), 28.5%; Ir-7b-major), 163.4 (d, 2J(P,P) = 185.4 Hz (1J(P,H) = 405 Hz), 2.8%; Ir-7b-minor). Warming up the reaction mixture to 40 C showed the formation of product Ir-5b (31P: 3.2 (d, 1J(P,P) = 371.6 Hz), 24.4 (d, 1J(P,P) = 371.6 Hz)), which upon warming up to room temperature rearranged quantitatively into Ir-6b.

Ligand Exchange Reactions. Reaction of Ru-5a with PPh3. To a cooled, green solution of Ru-5a (12.0 mg, 15.0 mol) in CD2Cl2 (0.5 mL) was added 1 eq. of PPh3 (3.9 mg, 15.0 mol) at 78 C, which left the colour of the reaction mixture unchanged. Subsequently, the reaction was monitored by
31P

NMR spectroscopy at 0 C

to afford [(6-pCy)(Ph3P)Ru=PMes*] (Ru-10) and Mes*CP quantitatively (t (0 C, CD2Cl2) = 100 min.). 134

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

Reaction of Ir-5b with PPh3. To a deep red solution of Ir-5a (13.4 mg, 15.0 mol) in C6D6 (0.5 mL) was added 1 eq. of PPh3 (3.9 mg, 15.0 mol).

Subsequently, the reaction was monitored by 31P NMR spectroscopy at 50 C for 210 h to afford as main products [(5-Cp*)(Ph3P)Ir=PMes*] (Ir-10; 60.3%) and Mes*CP (t (50 C, C6D6) = 28 h and 15 min.), along with unidentified by-products.
31P{1H}

NMR (101.25 MHz, C6D6, 273 K): 686.6 (d, 2J(P,P) = 102

Hz, 60.3%; Ph3PIr=PMes*), 153.7 (d, J(P,P) = 18.2 Hz, 5.5%), 34.1 (s, 22.2%; Mes* CP), 24.8 (s, 9.7%), 23.6 (d, 2J(P,P) = 102 Hz, 60.3%; Ph3PIr=PMes*), 14.2 (d, J(P,P) = 62.7 Hz, 6.6%), 24.9 (d, J(P,P) = 18.2 Hz, 5.7%), 38.9 (s, 4.9%), 79.5 (s, 4.6%), 96.8 (s, 2.7%). The reaction in CD2Cl2 at 20C gave after 68 hours only 0.8% of Ir-10, whereas in refluxing CD2Cl2 a variety of products were obtained of which D2C=PMes* could be identified. Reaction of Ru-5a with tBuCP. To a cooled, green solution of Ru-5a (12.0 mg, 15.0 mol) in CD2Cl2 (0.5 mL) was added 1 eq. of tBuCP (15.0 mol, 15.0 L of a 1M solution in toluene) at 0 C. Subsequently, the reaction was monitored by
31P

NMR spectroscopy at 0 C for 2.5 h, during which the

colour of the reaction mixture turned yellow, to afford Ru-6b ( 93%) and Mes*CP (t (0 C, CD2Cl2) = 17 min. and 48 sec.).

Acknowledgement. This work was partially supported by the Council for Chemical Sciences of the Netherlands Organization for Scientific Research (NWO/CW). We thank Dr. M. Smoluch (Vrije Universiteit, Amsterdam) and J. W. H. Peeters (University of Amsterdam) for measuring high-resolution mass spectra.

References and Notes


[1] (a) R. R. Schrock, Angew. Chem. 2006, 118, 38323844; Angew. Chem. Int. Ed. 2006, 45, 37483759. (b) R. H. Grubbs, Angew. Chem. 2006, 118, 38453850; Angew. Chem. Int. Ed. 2006, 45, 37603765. (c) Handbook of Metathesis (Ed: R. H. Grubbs), WILEY-VCH, Weinheim, 2003. (d) A. H. Hoveyda, A. R. Zhugralin, Nature 2007, 450, 243251.

135

Chapter 5

[2] Y. Chauvin, Angew. Chem. 2006, 118, 38243831; Angew. Chem. Int. Ed. 2006, 45, 37413747. [3] (a) C. Adlhart, P. Chen, J. Am. Chem. Soc. 2004, 126, 34963510. (b) L. Cavallo, J. Am. Chem. Soc. 2002, 124, 89658973. (c) B. F. Straub, Angew. Chem. 2005, 117, 61296132; Angew. Chem. Int. Ed. 2005, 44, 59745978. [4] (a) P. E. Romero, W. E. Piers, J. Am. Chem. Soc. 2005, 127, 50325033. (b) P. E. Romero, W. E. Piers, J. Am. Chem. Soc. 2007, 129, 16981704. (c) E. F. van der Eide, P. E. Romero, W. E. Piers, J. Am. Chem. Soc. 2008, 130, 44854491. (d) A. G. Wenzel, R. H. Grubbs, J. Am. Chem. Soc. 2006, 128, 1604816049. [5] With the appropriate choice of olefinic substrate, the ruthenium-olefin adduct can be isolated: J. A. Tallarico, P. J. Bonitatebus, Jr., M. L. Snapper, J. Am. Chem. Soc. 1997, 119, 71577158. [6] T. M. Trnka, M. W. Day, R. H. Grubbs, Organometallics 2001, 20, 38453847; and references therein. [7] For the all-carbon cyclobutene to 1,3-butadiene rearrangement, see: J. C. Slootweg, A. W. Ehlers, K. Lammertsma, J. Mol. Model. 2006, 12, 531536; and references therein. [8] T. Mitsudo, Bull. Chem. Soc. Jpn. 1998, 71, 15251538. [9] S. T. Diver, Coord. Chem. Rev. 2007, 251, 671701. [10](a) T. J. Katz, S. J. Lee, J. Am. Chem. Soc. 1980, 102, 422424. (b) D. E. Schuehler, J. E. Williams, M. B. Sponsler, Macromolecules 2004, 37, 62556257. (c) M. G. Mayershofer, O. Nuyken, J. Polym. Sci. A: Polym. Chem. 2005, 43, 57235747. [11] (a) J. Barluenga, F. Aznar, A. Martn, S. Garca-Granda, E. Prez-Carreo, J. Am. Chem. Soc. 1994, 116, 1119111192. (b) J. Barluenga, F. Aznar, I. Gutirrez, A. Martn, S. Garca-Granda, M. A. Llorca-Baragao, J. Am. Chem. Soc. 2000, 122, 13141324. [12] (a) M. M. Gleichmann, K. H. Dtz, B. A. Hess, J. Am. Chem. Soc. 1996, 118, 10551 10560. (b) K. H. Dtz, Angew. Chem. 1984, 96, 573594; Angew. Chem. Int. Ed. Engl. 1984, 23, 587608. (c) F. Hohmann, S. Siemoneit, M. Nieger, S. Kotila, K. H. Dtz, Chem. Eur. J. 1997, 3, 853859. [13] For example: (a) E. Niecke, A. Fuchs, M. Nieger, Angew. Chem. 1999, 111, 3213 3216; Angew. Chem. Int. Ed. 1999, 38, 30283031. (b) Y. Canac, D. Bourissou, A. Baceiredo, H. Gornitzka, W. W. Schoeller, G. Bertrand, Science 1998, 279, 2080 2082.

136

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

[14] Previously, only mono-phosphametallacyclobutenes (P1-B) have been observed or postulated, Ru: (a) A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Chem. Eur. J. 2003, 9, 22002208. (b) R. Menye-Biyogo, F. Delpech, A. Castel, V. Pimienta, H. Gornitzka, P. Rivire, Organometallics 2007, 26, 50915101. Zr: (c) T. L. Breen, D. W. Stephan, J. Am. Chem. Soc. 1996, 118, 4204 4205. (d) T. L. Breen, D. W. Stephan, Organometallics 1996, 15, 57295737. Ti: (e) G. Zhao, F. Basuli, U. J. Kilgore, H. Fan, H. Aneetha, J. C. Huffman, G. Wu, D. J. Mindiola, J. Am. Chem. Soc. 2006, 128, 1357513585. [15] (a) J. C. Slootweg, K. Lammertsma, In Science of Synthesis; B. M. Trost, F. Mathey, Eds.; Georg Thieme Verlag: Stuttgart, 2009; Vol. 42, pp 1536. (b) K. Lammertsma, Top. Curr. Chem. 2003, 229, 95119. [16] M. Gandelman, K. M. Naing, B. Rybtchinski, E. Poverenov, Y. Ben-David, N. Ashkenazi, R. M. Gauvin, D. Milstein, J. Am. Chem. Soc. 2005, 127, 1526515272. [17] X-ray crystal structure determination of Ru-5a: C47H72P2Ru C7H8, Fw = 892.19, brown block, 0.21 x 0.21 x 0.12 mm3, monoclinic, P21/c (no. 14), a = 16.9441(7), b = 16.2602(3), c = 17.7916(4) , = 93.347(1), V = 4893.5(2) 3, Z = 4, Dx = 1.211 g/cm3, = 0.42 mm1. 79778 Reflections were measured on a Nonius Kappa CCD diffractometer with rotating anode (graphite monochromator, = 0.71073 ) up to a resolution of (sin /)max = 0.65 1 at a temperature of 150(2) K. Intensities were integrated with EvalCC[25] using an accurate description of the experimental setup for the prediction of the reflection contours. The reflections were scaled and corrected for absorption using the program SADABS[26] (correction range 0.74-0.95). 11252 Reflections were unique (Rint = 0.0524), of which 8589 were observed [I>2(I)]. The structure was solved with automated Patterson methods using the program DIRDIF.[27] The structure was refined with SHELXL-97[28] against F2 of all reflections. Non hydrogen atoms were refined with anisotropic displacement parameters. All hydrogen atoms were introduced in calculated positions. Hydrogen atoms H39, H40, H42, and H43 were refined freely with isotropic displacement parameters; all other hydrogen atoms were refined with a riding model. 552 Parameters were refined with no restraints. R1/wR2 [I > 2(I)]: 0.0356 / 0.0674. R1/wR2 [all refl.]: 0.0613 / 0.0755. S = 1.045. Residual electron density between -0.69 and 1.17 e/3. Geometry calculations and checking for higher symmetry was performed with the PLATON program.[29] Ir-5a: C47H73IrP2, Fw = 892.19, red block, 0.30 x 0.24 x 0.12 mm3, monoclinic, P21/c (no. 14), a = 15.2053(1), b = 18.8033(1), c = 17.5203(1) , = 117.9352(4), V =

137

Chapter 5

4425.54(5) 3, Z = 4, Dx = 1.339 g/cm3, = 3.12 mm1. 95880 Reflections were measured on a Nonius Kappa CCD diffractometer with rotating anode (graphite monochromator, = 0.71073 ) up to a resolution of (sin /)max = 0.65 1 at a temperature of 150(2) K. Intensities were integrated with HKL2000.[30] The reflections were scaled and corrected for absorption using the program SADABS[26] (correction range 0.37-0.69). 10141 Reflections were unique (Rint = 0.0479), of which 9064 were observed [I>2(I)]. The structure was solved with automated Patterson methods using the program DIRDIF.[27] The structure was refined with SHELXL-97[28] against F2 of all reflections. Non hydrogen atoms were refined with anisotropic displacement parameters. All hydrogen atoms were introduced in calculated positions and refined with a riding model. The central IrP-P-C part of the molecule was refined with a disorder model corresponding to a twofold rotation about this unit and two positions for Ir (occupancy 79.2:20.8%). 511 Parameters were refined with 4 restraints. R1/wR2 [I > 2(I)]: 0.0288 / 0.0571. R1/wR2 [all refl.]: 0.0355 / 0.0586. S = 1.213. Residual electron density between -0.51 and 0.80 e/3. Geometry calculations and checking for higher symmetry was performed with the PLATON program.[29] Ru-6b: C33H52P2Ru, Fw = 611.76, yellow plate, 0.42 x 0.30 x 0.15 mm3, triclinic, P 1 (no. 2), a = 10.4153(1), b = 10.7589(1), c = 15.1059(2) , = 92.3448(5), = 109.5700(5), = 94.1699(5), V = 1586.88(3) 3, Z = 2, Dx = 1.280 g/cm3, = 0.61 mm1. 32748 Reflections were measured on a Nonius Kappa CCD diffractometer with rotating anode (graphite monochromator, = 0.71073 ) up to a resolution of (sin /)max = 0.65 1 at a temperature of 150(2) K. Intensities were integrated with HKL2000.[30] The reflections were scaled and corrected for absorption using the program SADABS[26] (correction range 0.68-0.91). 7266 Reflections were unique (Rint = 0.0310), of which 6612 were observed [I>2(I)]. The structure was solved with Direct Methods using the program SIR-97.[31] The structure was refined with SHELXL-97[28] against F2 of all reflections. Non hydrogen atoms were refined with anisotropic displacement parameters. All hydrogen atoms were located in difference Fourier maps. Hydrogen atoms H3, H5, H25, H26, H28, and H29 were refined freely with isotropic displacement parameters; all other hydrogen atoms were refined with a riding model. 364 Parameters were refined with no restraints. R1/wR2 [I > 2(I)]: 0.0226 / 0.0526. R1/wR2 [all refl.]: 0.0272 / 0.0544. S = 1.061. Residual electron density between -0.39 and 0.60 e/3. Geometry calculations and checking for higher symmetry was performed with the PLATON program.[29]

138

3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate

CCDC 687878 (Ru-5a), 687879 (Ru-6b), and 687880 (Ir-5a) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif. [18] This coordination mode resembles the known diphospha-cyclobutadiene complexes, e.g. Ru: (a) D. Himmel, M. Seitz, M. Scheer, Z. Anorg. Allg. Chem. 2004, 630, 12201228. Ir: (b) P. B. Hitchcock, M. J. Maah, J. F. Nixon, J. Chem. Soc., Chem. Commun. 1986, 737738. Co: (c) R. Wolf, A. W. Ehlers, J. C. Slootweg, M. Lutz, D. Gudat, M. Hunger, A. L. Spek, K. Lammertsma, Angew. Chem. 2008, 120, 46604663; Angew. Chem. Int. Ed. 2008, 47, 45844587. [19] R. Gleiter, I. Hyla-Kryspin, P. Binger, M. Regitz, Organometallics 1992, 11, 177181. [20] Gaussian 03, Revision C.02, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J. A. Pople, Gaussian, Inc., Wallingford CT, 2004. [21] The 1,3-isomer of Ir-7 (Ir-P=C-P) shows distinct resonances: 31P 652.8 and 93.0,
2JP,P

= 53.4 Hz.

[22] A. T. Termaten, T. Nijbacker, M. Schakel, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2002, 21, 31963202. [23] W. Rsch, T. Allspach, U. Bergstrer, M. Regitz in W. A. Herrmann (ed.), Synthetic Methods of Organometallic and Inorganic Chemistry (Herrmann/Brauer), Vol. 3, Thieme, Stuttgart, 1996, p. 1114. [24] H. Sugiyama, S. Ito, M. Yoshifuji, Angew. Chem. 2003, 115, 3932-3934; Angew. Chem. Int. Ed. 2003, 42, 38023804.

139

Chapter 5

[25] A. J. M. Duisenberg, L. M. J. Kroon-Batenburg, A. M. M. Schreurs, J. Appl. Cryst. 2003, 36, 220229. [26] G. M. Sheldrick, 1999. SADABS: Area-Detector Absorption Correction, v2.10, Universitt Gttingen, Germany. [27] P.T. Beurskens, G. Admiraal, G. Beurskens, W.P. Bosman, S. Garcia-Granda, R.O. Gould, J.M.M. Smits, C. Smykalla, (1999) The DIRDIF99 program system, Technical Report of the Crystallography Laboratory, University of Nijmegen, The

Netherlands. [28] G. M. Sheldrick, Acta Cryst. 2008, A64, 112122. [29] A. L. Spek, J. Appl. Cryst. 2003, 36, 713. [30] Z. Otwinowski, W. Minor, In Methods in Enzymology; C.W. Carter, Jr., R.M. Sweet, Eds.; Academic Press, 1997, Vol. 276, pp 307326. [31] A. Altomare, M. C. Burla, M. Camalli, G. L. Cascarano, C. Giacovazzo, A. Guagliardi, A. G. G. Moliterni, G. Polidori, R. Spagna, J. Appl. Cryst. 1999, 32, 115 119.

140

Appendix 1
BP86/TZP calculated bond dissociation energies and energy decomposition analysis for L(Bz)Ru=PPh (L = NHC, PMe3) belonging to Chapter 2. Table 3. Bond dissociation energies and energy decomposition analysis for (L1)(PhP)RuL2 (L1 = NHC, PMe3; L2 = Bz). Energies are given in kcal/mol. complex 1- 1- 2 Ea 96.2 98.1 94.6 Ea 67.9 64.6 61.2 E steric 107.3 108.5 98.3 E tot 56.8 54.3 57.5 E prep 12.8 21.3 11.7 BDE 44.0 33.0 45.8

Table 4. Bond dissociation energies and energy decomposition analysis for (C6H6)(L1)RuL2 (L1=NHC, PMe3; L2=PPh). Energies are given in kcal/mol. complex 1- 1- 2 Ea 114.8 108.9 111.6 Ea 38.6 39.1 38.8 E steric 67.9 67.6 64.8 E tot 85.5 80.4 85.7 E prep 5.2 10.6 4.8 BDE 80.3 69.8 80.9

141

142

Appendix 2
BP86/TZP optimized structures and calculated bond dissociation energies and energy decomposition analysis for L(Ring)M=PH belonging to Chapter 3.
M M C N N N P C N N P M C N N P C N M P

8 10 a b c d Figure 5. Optimized structures (Cs symmetry) for Group 7 transition metals (M = Mn, 8; Tc = 9; and Re = 10).

M M C N N P C

P M C N N N N P

C N

11 13 a b c d Figure 6. Optimized structures (Cs symmetry) for Group 8 transition metals (M = Fe, 11; Ru = 12; and Os = 13).

M M C N N P C

P M C N N P

C N N

14 16 a b c d Figure 7. Optimized structures (Cs symmetry) for Group 9 transition metals (M = Co, 14; Rh = 15; and Ir = 16).

143

Table 5. Calculated relative energies for complexes 816. Energies are relative to geometry a in kcal/mol. complex Geometry a b c d 8 0.00 4.14 2.15 4.01 9 0.00 4.70 1.05 3.15 10 0.00 4.77 1.14 0.06 11 0.00 3.17 2.42 2.53 12 0.00 2.84 2.14 2.90 13 0.00 5.30 1.42 0.04 14 0.00 1.60 1.91 0.93 15 0.00 1.34 1.66 1.20 16 0.00 1.66 2.15 4.01

Table 6. Bond dissociation energies and energy decomposition analysis for LnM C(NHC) bonds in Group 79 complexes. Energies are in kcal/mol. Values in parentheses refer to charge transferred (eV) to (positive) and from the metal center. complex 8c 9c 10c 11c 12c 13c 14c 15c 16c Ea 56.5 (0.54, 0.10) 58.3 (0.45, 0.11) 75.7 (0.48, 0.17) 58.5 (0.56, 0.12) 58.9 (0.44, 0.10) 78.6 (0.47, 0.17) 61.0 (0.60, 0.11) 63.6 (0.45, 0.12) 87.9 (0.48, 0.17) Ea 4.7 (0.03) 4.6 (0.01) 6.0 (0.01) 5.9 (0.01) 5.5 (0.01) 7.5 (0.01) 6.5 (0.02) 6.0 (0.02) 8.4 (0.01) E steric 14.8 13.1 15.8 13.8 12.1 19.1 10.6 13.7 25.0 E tot 46.4 49.8 66.0 50.6 52.4 67.0 56.8 55.9 71.3 E prep 11.7 9.1 16.9 8.4 9.3 12.9 6.0 8.4 12.5 BDE 34.7 40.7 49.1 42.2 43.1 54.1 50.8 47.5 59.8

Table 7. Bond dissociation energies and energy decomposition analysis for LnMRing bonds (Ring = Cht (cycloheptatrienyl), 810; Bz (benzene), 1113; Cp (cyclopentadienyl), 1416) in Group 79 complexes. Energies are in kcal/mol. complex 8a 9a 10a 11a 12a 13a 14a 15a 16a Ea 190.7 193.2 226.9 109.0 98.4 132.4 88.8 72.7 99.2 Ea 120.8 129.8 145.5 64.6 65.6 79.4 66.0 62.2 76.7 E steric 77.2 93.7 136.1 105.6 105.2 143.6 42.5 48.0 16.9 E tot 234.3 229.4 236.3 68.0 58.7 68.3 197.2 182.9 192.8 E prep 19.0 17.6 20.1 14.6 11.9 16.4 8.6 5.6 8.5 BDE 215.4 211.8 216.2 53.4 46.8 51.8 188.7 177.3 184.4

144

Table 8. Occupied SOMO energies (eV) for the [(Ring)(NHC)M] fragments and calculated Hirshfeld charges of M and P. M M P Ed () Ed () 8c 9c 10c 11c 12c 13c 14c 15c 16c Mn Tc Re Fe Ru Os Co Rh Ir 2.10 2.17 2.35 2.23 2.38 2.52 2.54 2.65 2.77 2.37 2.40 2.27 2.47 2.58 2.41 2.71 2.91 2.71 +0.00 +0.03 +0.12 0.06 +0.12 +0.02 0.08 +0.09 0.03 0.17 0.19 0.22 0.15 0.20 0.18 0.12 0.16 0.14

145

146

Samenvatting
De ontwikkeling van eindstandige fosfinideencomplexen: de zoektocht naar toepassing gaat door
De toepassing van nucleofiele, eindstandige metaalgecomplexeerde fosfinidenen [LnM=PR] als reactieve complexen is onderontwikkeld in vergelijking tot hun electrofiele tegenhangers en de welbekende carbeen en nitreen analogen [LnM=ER] (E = C,N). Dit proefschrift beschrijft efficinte syntheseroutes om nieuwe nucleofiele fosfinideencomplexen te genereren met een verhoogde reactiviteit en bestudeert de elektronische

eigenschappen hiervan met computerchemie. De empirische diagonale relatie tussen fosfor en koolstof in het Periodiek Systeem maakt de studie naar fosfinidenen aannemelijk. In 1975 stelde Schmidt voor dat vrije fosfinidenen (R-P:) betrokken waren De bij de vorming van cyclische van vrije

fosforhoudende fosfinidenen afhankelijk

verbindingen. is

chemie schaars de

en

toepassing vrije

(R-P:) van

betrekkelijk R, in

doordat

fosfinidenen, een triplet

substituent

meeste

gevallen

grondtoestand prefereren, dat leidt tot niet-(stereo)selectieve reactiviteit. Complexatie aan een metaal levert de gewenste singlet grondtoestand op voor het fosfinideen. De eerste voorbeelden van stabiele

fosfinideencomplexen werden gerapporteerd door Lappert en collegas in 1987 en met rntgendiffractie zijn de karakteristieke structuureigenschappen opgehelderd van Cp2M=PMes* (M = Mo, W). Veruit de meeste reactiviteitstudies hebben zich beperkt tot het stabiele zirconoceencomplex Cp2(PMe3)Zr=PMes*. Maar recent zijn ook andere vroege en late overgangsmetaalcomplexen gerapporteerd met beperkte, maar toch interessante reactiviteit. De electrofiele fosfinideencomplexen (OC)5M=PR (M = Cr, Mo, W) zijn kortlevend en hun bestaan berust op de reacties met geschikte substraten. Toch zijn zij veruit het meest bestudeerd en hebben 147

een

pallet

van

fosforhoudende

verbindingen

opgeleverd.

Recente

theoretische berekeningen toonden aan dat de mate van filiciteit van een fosfinideencomplex kan worden toegeschreven aan de liganden die gebonden zijn aan het overgangsmetaal dat vergelijkbaar is met carbeencomplexen. Liganden met sterke -donor eigenschap veroorzaken ladingsoverdracht op het fosforatoom van het fosfinideen met als gevolg een gepolariseerd M=P binding en nucleofiel reactiviteit. In homogene katalyse worden de tertiaire organofosfines (PR3) en N-heterocyclisch carbenen (NHC) veelzijdig gebruikt als ligand, waarbij de substituent R niet alleen de -donor en -acceptor capaciteit ervan bepaalt maar ook voor de nodige sterische afscherming zorgt. Lang werd aangenomen dat NHC een PR3 mimetic zou zijn. Echter, door zijn goede -donerende capaciteit vormt NHC sterke MC(NHC) binding en tenonrechte wordt vaak het backbonding aandeel verwaarloosd. Terwijl de eerste NHC complexen al in 1968 werden ontdekt (onafhankelijk van elkaar door Wanzlick en Schnherr, en fele) duurde het tot in 1991 dat de eerste vrije NHC werd gerapporteerd, dat is gestabiliseerd door sterisch gehinderde adamantyl groepen op N-atoom. Maar, mede dankzij de gedeelde Nobelprijs voor de chemie in 2005 van Grubbs, Schrock en Chauvin voor hun bijdrage in de metathese reacties, nam het gebruik van NHC als ligand explosief toe. De tweede generatie carbeencomplexen met NHC als ligand, vernoemd naar Grubbs, vertonen superieure activiteit in katalytische metathese reacties dan de corresponderende eerste generatie complexen met PR3 als ligand. Is een vergelijkbare activiteit ook te verwachten voor de analoge

fosfinideencomplexen met NHC als ligand? In Hoofdstuk 1 wordt een overzicht gegeven van synthese routes naar alle bekende nucleofiele fosfinideencomplexen. Met name de laatste

ontwikkelingen in hun toepassing als reactieve deeltjes wordt besproken. In Hoofdstuk 2 wordt de tweede generatie fosfinideencomplex [(6C6H6)(NHC)Ru=PMes*] 1 gerapporteerd en de reactiviteit wordt vergeleken met de eerste generatie fosfinideencomplex [(6-C6H6)(PPh3)Ru=PMes*] 2. Een verhoogde reactiviteit is bepaald voor 1, welke in een efficinte 148

npotsreactie

bereid

werd

door

middel

van

een

selectieve

dehydrohalogeneringsreactie van het primaire fosfine complex 3 met 3 equivalenten NHC. Karakteristieke verschillen in de
31P

NMR chemische

verschuivingen tussen 1 en 2 werden toegeschreven aan de sterk -donor eigenschappen van het NHC ligand.
Mes* Ru PH2 3 Mes* Ru P
-2 NHCHCl

3 NHC

N N NHC

Cl Cl

NHC 1

Inderdaad, complex 1 reageert 40x sneller met CH2I2 dan zijn voorganger 2. Het gevormde fosfaalkeen Mes*P=CH2 is een interessante bouwsteen voor de vervaardiging van nieuwe fosforhoudende liganden en polymeren. Recycling van het metaalcomplex 4 tot fosfinideen 1 is mogelijk d.m.v. een reactie met fosfine H2PMes* en de sterke base 1,8-diazabicyclo[5.4.0]undec7-een (DBU). Verhoogde nucleofiliciteit van complex 1 is het gevolg van sterische invloeden, waardoor het NHC ligand gedwongen uit-het-vlak roteert. Hierdoor vindt er effectief meer ladingsoverdracht plaats wat resulteert in meer nucleofiel gedrag van fosfinideencomplex 1.
Mes* Ru P L 1 L = NHC 2 L = PPh3 CH2I2 toluene L 4 L = NHC 5 L = PPh3 Mes* RuI2 H 2C P

H2PMes*, DBU toluene

In Hoofdstuk 3 is het gebruik van NHC liganden uitgebreid voor de synthese van stabiele NHC-gefunctionaliseerde fosfinideencomplexen van de metalen osmium en rhodium. Van rhodium fosfinideencomplex [(5Cp*)(NHC)Rh=PMes*] 7 werd een kristalstructuur bepaald die nauwelijks verschilt van het NHC-gefunctionaliseerde iridium fosfinideencomplex [(5Cp*)(NHC)Ir=PMes*] 8. Met behulp van theoretische berekeningen aan modelstructuren van overgangsmetalen van groep 7-9 met verschillende 149

liganden (cycloheptatriene+, benzeen, cyclopentadienyl) werd hun invloed en de rotatie van het NHC ligand op het electronische karakter van de fosfinideencomplexen onderzocht. Een gunstige metaal-NHC -interactie (~ 20%) ontstaat voor complexen met het NHC ligand in-het-vlak, als is getoond voor het berekende complex 7 zonder sterische hindering, waarvan de metaal-NHC -interactie niet mag worden verwaarloosd.
L M Cl Cl L Mes* PH2 3 NHC - 2 NHC.HCl N M N P 6, LM = (pCy)Os 7, LM = (Cp*)Rh 8, LM = (Cp*)Ir Mes*

Echter, door sterische afstoting roteert het NHC ligand naar een uit-hetvlak symmetrie, zoals bepaalt is voor de complexen 1,7 en 8, waarbij de interactie de bijbehorende orbitaal energien domineert. Verder volgt uit de theoretische berekeningen dat fosfinideencomplexen met

overgangsmetaal rhenium, ruthenium en rhodium naar verwachting het meest reactief zijn van alle bestudeerde groep 7-9 overgangsmetaalcomplexen doordat de M=P binding het meest gepolariseerd is.

Rh C N N

DFT berekende structuur van 7 met NHC in-het-vlak

Rntgen kristalstructuur van 7 met NHC uit-het-vlak

In

Hoofdstuk

wordt

het

CH2Cl2-gestabiliseerde Z-9c (R = Dmp)

16-electron experimenteel

fosfinideencomplex

[(5-Cp*)Ir=PR]
31P

aangetoond en ondersteund met

NMR berekeningen. Verder wordt de

reactie met isocyanides (XyNC) gepresenteerd die via, voor de nucleofiele fosfinideencomplexen onbekende, [1+2]-cycloadditie leidt tot nieuwe iridafosfiraancomplexen 11. Eerdere pogingen naar de synthese van 150

16-electron fosfinideencomplexen [(5-Cp*)Ir=PR] 9 (R = Mes, Mes*) zonder stabiliserende liganden resulteerden, via de voorgestelde dimeren [9]2 structuren, in C-H insertieproducten. Terwijl de nitreen analoog [(5Cp*)IrNtBu] van Bergman en collegas isoleerbaar is, maar instantaan reageert met isocyanides tot iridaazirideen complexen, hebben wij experimenteel en theoretisch aangetoond dat afvangen van fosfinideen 9 met isocyanide eerst stabiele, isoleerbare 18-electron complexen 10 geeft. Vervolgens ringsluit een tweede isocyanide met het fosfinideen en geeft iridafosfiraan 11.
Cp* Ir Cl Cl R PH2 2 DBU -2 DBUHCl Cp* Ir P E/ Z-9 Cp* R Ir P C C N 11 Xy R a Mes* b Mes c Dmp R Xy-NC Cp* Ir P C N 10 Xy R

Xy-NC N Xy

Het theoretisch berekende reactiemechanisme van nitreen analoog [(5Cp)IrNH] 12 met isocyanide (HNC) geeft eerst een ringsluiting, en pas dan cordineert een tweede isocyanide als ligand om 13 te vormen.
Cp Cp Ir N H 12' H-NC Ir N C N H H H-NC Cp C H Ir N C N N 13' H H

In Hoofdstuk 5 worden de eerste difosfor-analoga van de veelzijdige Dtz intermediairen, 3-difosfavinylcarbene complexen 16, gepresenteerd. Als uitgangspunt wordt ook hier de in situ gegenereerde reactieve 16electronen fosfinidenen [LnM=PR] gebruikt. DBU-genduceerde reactie van Ir en Ru primaire fosfine complexen 14 in aanwezigheid van fosfaalkyn Mes*CP geeft eerst complex 15 en daarna pas het 16-electronen

151

fosfinideen dat wordt afgevangen door het fosfaalkyn om uitsluitend 3difosfavinylcarbeencomplexen 16 te vormen. De kristalstructuur van

complex 16 toont de unieke metallacycle waarvan het koolstofatoom carbenoid is. In aanwezigheid van PPh3 ondergaat 16 een liganduitwisseling waarbij het fosfaalkyn Mes*CP wordt vervangen en het stabiele 18electronen fosfinideencomplex [(6-pCy)(PPh3)Ru=PMes*] ontstaat.
Mes* DBU RC P, DBU LM PH Cl - DBUHCl - DBUHCl Cl Cl 15 LM = 14a, 5-Cp*Ir R = Mes*, t Bu 14b, 6-pCyRu; Mes* M PH2 t Bu t Bu LM P P R 16 R = Mes* t Bu LM t Bu for R = t Bu R P t Bu 17 R = t Bu t Bu P L

Wanneer het sterisch minder gehinderde fosfaalkyn tBuCP wordt gebruikt, is 16 niet stabiel en vindt er een omlegging plaats tot complex 17 dat overeenkomt met de Dtz benzannulatie reactie. De stapsgewijze dehydrohalogeneringsreactie van complex 14 werd m.b.v variabele temperatuur NMR spectroscopy gevolgd. DFT berekeningen aan

modelstructuren toonden aan dat omlegging van 16 via een reeks reactiestappen verloopt, met P-P spliting,

2-gecoordineerd fosfaalkyn

rotatie, P-C bond vorming en electrofiele aanval op het aromaat als belangrijkste.

152

Curriculum Vitae
Halil Akta was born in the Karyaka district of the province zmir in Turkey on the 20th March 1976. In 1987 he moved together with his brother, sisters and mother to the Netherlands due to the family reunification. At the age of 14 he decided to study chemistry and in 2001 he obtained his B.Sc. in Organic Chemistry at the Hogeschool van Utrecht. His B.Sc. thesis project concerned the study towards Schrock-type phosphinidenes at the Vrije Universiteit, Amsterdam. He received his M.Sc. in Organic Chemistry at the same institute in 2003. Under the supervision of prof.dr. Romano Orru, Halils M.Sc. project thesis consisted of the synthetic study towards 3-deoxyribolactones using a hydrolysis-induced lactonization cascade reaction as synthons to utilize "sugar" part of antibiotic Mureidomycin A. Then, in September 2003 Halil initiated his Ph.D. project at the Vrije Universiteit, Amsterdam, under the supervision of prof.dr. Koop Lammertsma towards the nucleophilic phosphinidene complexes. His current work as R&D Scientist at Tate and Lyle comprises renewable ingredients and started in March 2008. His work focuses on developing food and industrial starches manufactured by means of chemical, thermal, and physical modification processes. He advices up- and down-scaling of product manufacturing, provides chemical

understanding of products and the corresponding processes, develops analytical methodologies, and assists in technical support. Halil is an amateur guitar, saz, and lute player.

153

154

List of publications
Nucleophilic Phosphinidene Complexes Access and Applicability, Halil Aktas, J. Chris Slootweg, Koop Lammertsma, Angew. Chem. 2009, accepted. Iridium Phosphinidene Complexes: A Comparison with Iridium Imido Complexes in Their Reaction with Isocyanides, Halil Aktas, Jos Mulder, Frans J. J. de Kanter, J. Chris Slootweg, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, Koop Lammertsma, J. Am. Chem. Soc. 2009, 131, 1353113537. N-heterocyclic Carbene Functionalized Group 7-9 Transition Metal Phosphinidene Complexes, H. Aktas, J. C. Slootweg, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Organometallics 2009, 28, 51665172. 3-Diphosphavinylcarbene: A P2 Analogue of the Dtz Intermediate, H. Aktas, J. C. Slootweg, A. W. Ehlers, M. Lutz, A. L. Spek, K. Lammertsma, Angew. Chem. 2009, 121, 31543157; Angew. Chem. Int. Ed. 2009, 48, 31083111. N-heterocyclic Carbene Functionalized Ruthenium Phosphinidenes: What a Difference a Twist Makes, Halil Aktas, Marius Schakel, J. Chris Slootweg, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, Koop Lammertsma, J. Am. Chem. Soc. 2009, 131, 66666667. Synthesis of 3-Deoxyribolactones using a Hydrolysis-Induced Lactonization Cascade Reaction of Epoxy Cyanohydrins, Danielle J. Vugts, Halil Aktas, Kanar Al-Mafraji, Renske, Frans J. J. de Kander, Eelco Ruijter, Marinus B. Groen, Romano V. A. Orru, Eur J. Org. Chem. 2008, 13361339. Terminal Phosphinidene Complexes Cp[R](L)M=PAr of the Group 9 Transition Metals Cobalt, Rhodium, and Iridium. Synthesis, Structures, and Properties, Arjan T. Termaten, Halil Aktas, Marius Schakel, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, Koop Lammertsma, Organometallics 2003, 22, 18271834.
155

156

Dankwoord
Eindelijk is het zo ver, mijn proefschrift is af! Het is een mooi moment om iedereen te bedanken die op welke manier dan ook een bijdrage heeft geleverd aan de totstandkoming van mijn boekje. Een aantal mensen wil ik in het bijzonder bedanken. Mijn promotor prof.dr. Koop Lammertsma, bedankt dat u mij heeft gevraagd en de kans heeft gegeven om aan dit promotieonderzoek te beginnen. Dat mijn hart in het onderzoek lag, was snel duidelijk, maar de verleiding om naar de industrie te gaan was ook groot. Gelukkig hebben we aan het begin van deze rit de nodige discussies gehad en uiteindelijk kwam alles op zijn pootjes terecht. Wat betreft het onderzoek kreeg ik vanaf dag 1 alle vrijheid. Ondanks mijn exotische ideen liep de chemie echter pas in de tweede helft van mijn Ph.D. Wel bleven we altijd lachen, ook op momenten wanneer de chemie niet liep. Ondanks uw drukke agenda, wist u toch tijd vrij te maken. Marius Schakel, bedankt voor de dagelijkse begeleiding in de experimentele werk en je kritische, maar heldere blik op de materie. Wanneer experimenten mislukten, wist jij wel een manier om het toch nog te laten lukken. Andreas Ehlers, bedankt voor al je hulp op het gebied van computerchemie. Onze gesprekken in de OV en tijdens de korte wandelstukjes naar de VU waren erg stimulerend. Beste Chris Slootweg, mede dankzij jouw sterke betrokkenheid vanaf jouw dag 1, je goede ideen, je kritische blik op het onderzoek en de correcties aan mijn manuscripten heeft mij enorm geholpen in het zetten van grote stappen. Bedankt voor al je hulp en de dagelijkse begeleiding. I would like to extend my gratitude to the reading committee members prof.dr. F.M. Bickelhaupt, dr. B. de Bruin, prof.dr. C.J. Elsevier, prof.dr. E. HeyHawkins, and dr. C. Mller for being the referees of this thesis and being part of the opposition. Beste Arjan Termaten, mijn dank gaat ook naar jouw enthousiaste dagelijkse begeleiding die ik tijdens mijn stage periode kreeg. Mede dankzij jou groeide mijn interesse naar fosfinidenen en gelukkig mocht ik het 157

onderzoek verder oppakken. Tevens wil ik Sander van Assema bedanken voor de gezelschap op het laboratorium en de buurman Bas de J. voor de nodige glaswerk. Het was altijd zoeken naar die laatste schone NMR buisje! Mijn studenten Abdul Karim (laat even weten waar je bent!), en Jos. Jullie beide wil ik bedanken voor jullie bijdragen, beschreven in Hoofdstuk 4, naar het oplosmiddel gestabiliseerde 16-elektron deeltje. Frans de Kanter, bedankt voor je hulp met NMR metingen onder extreme condities en de 2D verzoekjes. Marek Smoluch, thank you very much for the HRMS

measurements of the compounds, in a relatively short period of time. J.W. Han Peeters, ook jij bedankt voor de FAB metingen wanneer EI een te krachtig methode bleek. Martin Lutz, bedankt voor de kristalstructuur analyses. Vaak bleven de kristallen op hun plek zitten. Ik hield altijd mijn hart vast en gelukkig klopte de verwachtte structuur als een bus. Rob Schmitz, bedankt voor al je hulp bij experimentele opstellingen en de vacum pompen. Andr en Erik, jullie beide wil ik bedanken voor jullie hulp met Spartan, Gaussian en ADF tijdens de promotietijd. Andr, zonder jou zouden we op lauwe Dommelsch moeten leven en Erik, jij had altijd wel een neus voor een goed wijn. O ja, nog bedankt voor je hulp met de veldcodes! Also, Id like to thank all the international Ph.D. students and Post-docs for the great company. Nuria Ortega, Federico, Federica, Masahiro, Sebastian Burck, Hiroshi Naka, and Robert Wolf: thanks. Verder wil ik emeritus prof.dr. F. Bickelhaupt, prof.dr. Romano Orru, Eelco, Fedor, Mark B., Bas G., Niels T., Niels E., Helen, Robin, Lisette, Maurice, Rachel, Amos, Shen, Maarten, Wannes, Anass, alle AIOs, studenten, en OACers die ik bij naam ben vergeten bedanken voor de vele onvergetelijke momenten, de conferentie(s) (16th ICPC Birmingham, NCCC en 4th PhD Workshop on Phosphorus Chemistry), de borrels, de jaarlijks terugkerende recepties en de gezellige diners. Judith bedankt voor de onvergetelijke, heerlijke Indische gerechten, maar ook voor even bijpraten. Miep, jij ook bedankt voor het regelen van allerlei papier zaken, maar ook voor de goede praatjes. Mijn collegas en vrienden van Tate and Lyle wil ik hierbij ook bedanken voor hun interesse in mijn promotieonderzoek. Mark, Nicoline, dr. Jere 158

Koskinen, Serge, Emiel, Sadi, (ha!) Tom (tom) en Coen. Furthermore, Id like to thank the people who supported me outside the VUA and T&L and who shared the great fun during my Ph.D. period. Tommy & Celine, Ana, Micaela, Nalan and Harry. Thank you for the wonderful moments, drinks, BBQs, trips and parties. Ook de jongens band, dostlarm waarvan sommige tot de VUA meubilair behoren wil ik bedanken voor de vele uitstapjes, heftige discussies, de etentjes, en de feestjes. Atilla, Ahmet, Blent, Kaan, Timur, Volkan, Uur, Ekrem en hola dos Ouz. Mijn fysiotherapeut arkada R. Meents moet ik ook bedanken. Zonder zijn therapie zou ik de lange schrijfwerkdagen niet door kunnen komen. My friends Engin, aban, Murat, and Sevda, Id like to thank you being great company during the holidays. Thank you for the true friendship, which started as classmates back in zmir. Dostum Okan Akn wil ik speciaal bedanken. Niet alleen maar voor zijn bijdrage in de omslag en de binnenwerk, maar ook voor zijn oprechte vriendschap en broederschap. Succes Mustafa met alles wat je onderneemt! Mijn schatjes Glay, Meneke-Naz, pek, en mijn neefje grote mannetje Eray. Ik sta altijd voor jullie klaar. Ook voor jou Emel! Erol, pas goed op haar! Mijn broer Medayim, en mijn nichtje Emel wil ik bedanken omdat ze mijn paranimf wilden zijn. Bedankt voor al je steun broer. Je was er altijd wanneer ik je nodig had en ik heb veel van je geleerd. Mijn zussen Hacer, Ycel, Hazime, en schoonzus Ebru wil ik bedanken voor alle geduld tijdens het onderzoek en het schrijfperiode. Ook jullie steunden mij in alles wat ik deed. Canm haclarm anneciim ve babacm. zerimde deyemiyeceim kadar ok hakknz var. Nacizhane bu kitab aileme ve sevdiklerime hediye ediyorum. Rabbime binlerce kr ve hamd senalar olsun. Nu het boekje eindelijk echt af is, heb ik meer tijd voor mijn dierbaren.

Zaandam, oktober 2009 Halil AKTA

159

You might also like