Sunteți pe pagina 1din 8

FULL PAPER

DOI: 10.1002/ejic.201100131

Aminotroponiminate and Aminotroponate Complexes of Group 15 (P, As)


Elements: Synthesis, X-ray Diffraction Analysis and Reactivity

Lucian-Cristian Pop,[a,b] Annie Castel,*[a] Luminita Silaghi-Dumitrescu,*[b] and


Nathalie Saffon[c]

Keywords: Phosphorus / Arsenic / N,N ligands / N,O ligands / Cycloaddition / Tungsten

New phosphenium and arsenium cations stabilized by biden- cycles. The reactivities of these cationic species were investi-
tate monoanionic N-isopropyl-2-(isopropylamino)troponim- gated. The phosphenium cations were found to be more reac-
inate or 2-(isopropylamino)troponate units have been synthe- tive than their arsenium analogues, especially towards an o-
sized. These complexes were characterized by 31P, 1H and quinone, with the formation of a phosphate ester derivative.
13
C NMR spectroscopy and the molecular structures were de- Complexation reactions with [(THF)W(CO)5] afforded the
termined by X-ray crystallography. These data indicate the first aminotroponiminato hydroxyphosphenium cation stabi-
formation of N,N⬘- and N,O-chelate derivatives with three- lized by coordination to a transition metal.
coordinate Group 15 atoms included in planar heterobi-

Introduction like other low-coordinate Group 15 elements, are conspicu-


ous by virtue of their expected ambiphilic (Lewis acid and
Aminotroponiminates ([ATI]–) are a well-known class of base) properties. They are six-electron dicoordinate
ligands that have found extensive use in coordination chem- Group 15 species and are isovalent analogues of carbenes.
istry leading to a large number of transition-metal and The stability of these species depends on charge delocaliza-
main-group complexes.[1] On the one hand, this ligand dis- tion through heteroatom bonding and/or by incorporation
plays heteroatom bonding and a highly conjugated 10π elec- of the pnictogen atom into a conjugated π system. Among
tron system that should allow charge delocalization when the numerous types of pnictogenium cations that have been
incorporating the pnictogenium atom into the π electron prepared and studied,[3] the most important class is that
backbone. On the other hand, and in marked contrast to analogous to N-heterocyclic carbenes that display two nor-
the β-diketiminate scaffold, [ATI]– is expected to form mal covalent E15–N bonds (E15 = P,[4] As,[5] Sb[4c,5b,6]).
stable five-membered chelate rings and its anticipated fused More recently, some species having amino/imino ligands
ring structure should disfavour side-reactions. In contrast, with a covalent E15–N bond and a coordinative E15씮N
aminotroponates ([AT]–), which may formally be considered bond have been reported. In this context, Cowley and co-
as a combination of amido and carbonyl donors, have been workers synthesized the first example of a phosphenium
much less investigated.[2] However, the coordination chemis- cation stabilized by the N,N⬘-chelation of a β-diketiminato
try of both these ligand types with Group 15 divalent ele- ligand, which can be regarded as an intramolecular donor
ments remains totally unexplored. Pnictogenium cations, adduct.[7] Surprisingly, only a few Group 15 element com-
plexes of this type have been described,[8] whereas a large
number of intermolecular donor adducts of both phos-
phenium[3a,9] and arsenium[10] cations are known. As part
[a] Laboratoire Hétérochimie Fondamentale et Appliquée, UMR/ of our ongoing investigation of main-group elements in un-
CNRS 5069, Université Paul Sabatier, usually low oxidation states, we have recently shown that
118, Route de Narbonne, 31062 Toulouse Cedex 9, France
Fax: +33-5-61558204 chlorophosphenium cations can be stabilized by aminotro-
E-mail: castel@chimie.ups-tlse.fr poniminate and aminotroponate ligands.[11] In this paper,
[b] Babes-Bolyai University, we report an extension of this work to the N,N- and N,O-
1, Kogalniceanu Street, Cluj-Napoca, Romania
Fax: +40-264-59-19-06 chelated stabilization of the arsenic element. Some reac-
E-mail: lusi@chem.ubbcluj.ro tions, namely oxidative cycloaddition with an o-quinone
[c] Structure Fédérative Toulousaine en Chimie Moléculaire
(FR2599), Université Paul Sabatier, and complexation with a transition-metal complex, will be
118, Route de Narbonne, 31062 Toulouse Cedex 9, France also described.

Eur. J. Inorg. Chem. 2011, 3357–3364 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 3357
L.-C. Pop, A. Castel, L. Silaghi-Dumitrescu, N. Saffon
FULL PAPER
Results and Discussion at –30 °C. The molecule crystallizes in the monoclinic
P21/c space group with one H2O molecule per molecule of 2
Synthesis, NMR Spectroscopy and Structural Studies of (Figure 1). The aminotroponiminato ligand chelates to the
Pnictogenium Cations 1–4 arsenic centre with near perfect planarity of the two cycles.
The most noteworthy structural feature concerns the As–
In previous work,[11] we have shown the impact of the Cl bond lengths, which differ significantly [2.3759(4) and
synthetic process on the nature of the product obtained 2.7974(6) Å]. The first value is in the range of As–Cl cova-
when phosphorus bears an aminotroponiminato substitu- lent bond lengths observed in the related compound
ent. Indeed, the nucleophilic substitution reaction of PCl3 ClAs(tBuNCH2CH2NtBu) [2.375(2) Å][4a] and ClAs(iPrN)2-
and the monolithiated ligand gave a mixture of dissociated C10H6 [2.2820(8) Å].[4c] In contrast, the second is much
and non-dissociated forms 1a and 1b whereas the base-in- elongated and is in the range of the bridging As–Cl bonds
duced dehydrochlorination coupling reaction selectively observed in chlorodiarsenate Ph4P[As2Cl8]·CH3CN (2.639
yielded the phosphenium cation 1a (Scheme 1). and 3.042 Å).[12] However, the corresponding chloride atom
is hydrogen-bonded to the H2O molecule with a distance of
3.209 Å for the Cl···H(O) interaction. The presence of such
an interaction could have an effect on the As–Cl bond
length. Accordingly, it was of interest to examine the struc-
ture of an arsenium salt that does not present hydrogen
bonding.

Scheme 1.

The analogous arsenic system is more complex because


only one form (2 or 3) was obtained regardless of the exper-
imental conditions used and the nature of the ligand Figure 1. Molecular structure of 2 (50 % probability level for the
(Scheme 2). These compounds were isolated as yellow solids thermal ellipsoids) with hydrogen atoms and solvent molecules
in 50–60 % yields. Note that the N,O-chelated product 3 is omitted. Selected bond lengths [Å] and angles [°]: As1–N1
1.8445(12), As1–N2 1.8502(12), As1–Cl1 2.3759(4), As1–Cl2
more unstable than the N,N⬘-chelated analogue 2, particu- 2.7974(6), N1–C1 1.3611(17), N2–C2 1.3585(18), C1–C7 1.398(2),
larly in solution, leading to the formation of aminotropone. C1–C2 1.448(2), C2–C3 1.400(2), C3–C4 1.385(2), C4–C5 1.381(2),
C5–C6 1.378(2), C6–C7 1.386(2); N1–As1–N2 84.28(5), N1–As1–
Cl1, 95.37(4), N2–As1–Cl1 96.46(4), N1–As1–Cl2, 88.26(4), N2–
As1–Cl2 91.16(4), Cl1–As1–Cl2 171.85(1), C1–N1–As1 115.53(9),
C2–N2–As1 115.58(10).

The triflate salt 4 was easily prepared from 2 by chloride


abstraction using trimethylsilyl triflate (Me3SiOTf,
Scheme 2. OTf=OSO2CF3; Scheme 2). As expected, the 1H NMR
spectrum shows a more pronounced deshielding of the ring
The 1H NMR spectra of compounds 2 and 3 exhibit res- protons (Δδ ≈ 1.5 ppm) compared with in the starting li-
onances for the ring protons that are significantly shifted gand. Its molecular structure (including selected bond
downfield (Δδ ≈ 1.0 ppm) relative to those of the starting lengths and angles) is shown in Figure 2. Compound 4
ligands, presumably because of the increased conjugation of shows the same general features as phosphenium cation 1[11]
π electrons in the newly formed metallacycles. A similar ef- consisting of separated ion-pairs with triflate as the anion.
fect, but more pronounced, has already been observed for The nearest distance between the ions is F–···H(cation)
the phosphenium cation 1a. The spectroscopic data indicate (2.73 Å). The shortest distance between the anion and the
the possible existence of dissociated forms. To gain more arsenic atom is 6.91 Å. The cationic part is formed by the
insight into their structures, we performed X-ray diffraction AsCl unit, which is stabilized by the chelating ligand sys-
analysis. Single crystals of compound 2 were isolated by tem. The three-coordinate arsenic atom adopts a pyramidal
very slow crystallization from a dichloromethane solution geometry (sum of the angles is 282.8°) with the bonded

3358 www.eurjic.org © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Inorg. Chem. 2011, 3357–3364
Aminotroponiminate and Aminotroponate Complexes

chloride atom in an approximately orthogonal position with


respect to the heterobicyclic plane. The As–N bond lengths
[As–N1 = 1.837(4), As–N2 = 1.840(4) Å] are similar to
those reported for a five-membered aromatic arsenium cat-
ion [1.8377(14) and 1.8271(13) Å].[5b] Likewise, the C–C
and C–N bond lengths are only marginally different, which
suggests an almost similar degree of delocalization of the
electrons in the systems.

Figure 3. Molecular structure of the dimeric unit of compound 3


(50 % probability level for the thermal ellipsoids) with hydrogen
atoms omitted. The two independent molecules of the asymetric
unit are shown. Selected bond lengths [Å] and [°]: As1–O1
1.8194(16), As1–N1 1.9250(17), As1–Cl1 2.3202(6), As1–Cl3
2.7721(7), As1–Cl3A 2.8740(7), N1–C1 1.333(3), O1–C2 1.348(3),
C1–C7 1.420(3), C1–C2 1.446(3), C2–C3 1.365(3), C3–C4 1.404(3),
C4–C5 1.371(3), C5–C6 1.393(3), C6–C7 1.373(3); O1–As1–N1
83.56(7), O1–As1–Cl1 93.76(5), N1–As1–Cl1 93.24(6), C1–N1–As1
114.19(14), C2–O1–As1 115.97(13), O1–As1–Cl3 84.91(5), N1–
Figure 2. Molecular structure of 4 (50 % probability level for the As1–Cl3 87.28(6), O1–As1–Cl3A 81.81(5), N1–As1–Cl3A
thermal ellipsoids). All H atoms have been omitted for clarity. Se- 165.36(6), Cl1–As1–Cl3 88.41(2), Cl1–As1–Cl3A 178.51(3), Cl3–
lected bond lengths [Å] and angles [°]: As1–N1 1.837(4), As1–N2 As1–Cl3A 90.73(2); As2–O2 1.8150(16), As2–N2 1.9208(19), As2–
1.840(4), As1–Cl1 2.2808(15), N1–C1 1.356(5), N2–C2 1.360(5), Cl2 2.3088(8), As2–Cl4 2.7222(8), As2–Cl4A 2.9884(5), N2–C11
C1–C7 1.396(6), C1–C2 1.447(6), C2–C3 1.396(6), C3–C4 1.385(6), 1.337(3), O2–C12 1.344(3), C11–C17 1.418(3), C11–C12 1.437(3),
C4–C5 1.393(7), C5–C6 1.371(7), C6–C7 1.390(7); N1–As1–N2 C12–C13 1.382(3), C13–C14 1.398(3), C14–C15 1.364(4), C15–C16
84.77(16), N1–As1–Cl1, 99.44(12), N2–As1–Cl1 98.63(12), C1– 1.401(4), C16–C17 1.367(3); O2–As2–N2 84.00(8), O2–As2–Cl2
N1–As1 115.5(3), C2–N2–As1 114.9(3). 92.62(6), N2–As2–Cl2 93.28(6), C11–N2–As2 113.54(15), C12–O2–
As2 115.54(14), O2–As2–Cl4 86.05(6), N2–As2–Cl4 87.49(6), O2–
As2–Cl4A 81.94(6), N2–As2–Cl4A 164.91(6), Cl2–As2–Cl4
178.39(3), Cl2–As2–Cl4A 92.83(3), Cl4–As2–Cl4A 86.08(2).
Interestingly, compound 3 adopts a very different struc-
ture. In the solid state, the neutral dichloroarsine unit pres-
Reactivity of Pnictogenium Cations
ents a dimeric arrangement involving a four-membered
As2Cl2 ring (Figure 3). The asymmetric unit contains two It is well known that phosphenium cations act like dieno-
chemically similar but crystallographically different dimers philes with 1,3-dienes.[3a] However, neither phosphenium
of 3. The main difference lies in the As–Cl distances. The cation 1a nor the arsenium derivative 2 react with dimethyl-
As–Cl distances [2.7221(7) and 2.8740(7) Å, 2.7222(8) and butadiene, whatever the experimental conditions used (in
2.9884(5) Å] of the As2Cl2 parallelogram are significantly dichloromethane solution[16] or in a sealed tube at
elongated compared with those of the exocyclic As–Cl 100 °C[17]). In contrast, an immediate reaction was observed
bonds [2.3202(6) and 2.3088(8) Å], which are in the range between phosphenium cation 1a and 3,5-di-tert-butyl-1,2-
of covalent As–Cl bonds.[4a,4c] Similar halogen-bridged ar- benzoquinone at room temperature, however, the arsenium
rangements have been reported for chloroarsenic com- 2 failed to react.
pounds such as chlorobenzothiazarsole,[13] chloroarsines,[14] The dropwise addition of the o-quinone to a solution of
and chlorodiarsenates.[12] The heterobicyclic unit is planar compound 1a in CH2Cl2 caused an immediate discoloration
and perpendicular to the As2Cl2 system. The As–O bond of the quinone solution. The reaction was monitored by 31P
lengths [1.8194(16) and 1.8150(16) Å] are close to that of a NMR, which showed the formation of a single product (δ
single bond (1.78 Å)[15] and the As–N distances [1.9250(16) = 7.83 ppm), identified as 5, that was easily isolated after
and 1.9208(19) Å] are elongated compared with those in 4, evaporation of the solvent (Scheme 3). Compound 5 is very
which confirms a strong interaction between the arsenic sensitive to moisture in solution and rapidly gives the phos-
atom and the chelating system even in the case of the non- phate ester 6, which is the tautomeric form of hydroxyphos-
dissociated form. phorane ⬎P(OH)O–.[18] Slow crystallization in CH2Cl2 at

Eur. J. Inorg. Chem. 2011, 3357–3364 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.eurjic.org 3359
L.-C. Pop, A. Castel, L. Silaghi-Dumitrescu, N. Saffon
FULL PAPER
–30 °C gave suitable crystals for X-ray analysis (Figure 4).
No interaction was detected between the phosphorus atom
and the chloride anion (P+···Cl–, 3.920 Å) and the two fused
cycles (five- and seven-membered rings) are almost in the
same plane. The phosphorus atom is tetracoordinate. The
P–O bond length [1.568(3) Å] lies between a typical P–O
single bond (1.71 Å) and a P=O double bond (1.40 Å).[19]
The P=O bond length [1.456(3) Å] is very close to the value
of a standard P=O double bond. The dioxygenated phenyl
system lies in a plane perpendicular to the cationic part to
reduce steric constraints.

Figure 4. Molecular structure of 6 (50 % probability level for the


thermal ellipsoids). All H atoms (except H3) and solvent molecules
have been omitted for clarity. Selected bond lengths [Å] and angles
[°]: P1–O1 1.456(3), P1–O2 1.568(3), P1–N1 1.667(4), P1–N2
1.669(4), N1–C1 1.384(5), C1–C7 1.380(6), C1–C2 1.440(6), C2–
C3 1.407(6), C3–C4 1.375(6), C4–C5 1.382(6), C5–C6 1.381(6), C6–
C7 1.384(6); O1–P1–O2 115.05(18), O1–P1–N1 120.62(18), O2–P1–
N1 101.95(18), O1–P1–N2 115.43(19), O2–P1–N2 108.67(18), N1–
P1–N2 92.20(19), C1–N1–P1 113.2(3), C2–N2–P1 113.1(3).

[{(C5Me5)(tBuNH)P}W(CO)3(MeCN)2]+ AlCl4– (404 Hz),[21]


but smaller than that reported for the W=P double bond in
Scheme 3. [Cp(CO)2W=P{OC(CH3)2C(CH3)2O}] (847 Hz),[22] results
that suggest a small amount of π character in the W–P
DFT calculations were performed on phosphenium[11] bond.
and arsenium cations.[20] They revealed a high-lying phos- The 13C NMR spectrum shows two sharp resonances at
phorus lone-pair orbital (HOMO–1) which should allow an low field of relative intensity 4:1 belonging to the equatorial
easy coordination of phosphenium cations to metal car- and axial carbonyl groups that are split by C–P coupling: δ
bonyl fragments. In contrast, the arsenic lone-pair resides = 195.20 (2JC,P = 9.3 Hz) and 196.18 ppm (2JC,P = 8.6 Hz).
in the HOMO–4 and seems to be less accessible for metal The trans carbonyl resonance is slightly deshielded relative
complexation. This proved to be the case during its reaction to the cis carbonyl, as previously observed in the literature
with [W(CO)5THF] as no arsenium complex could be char- for complexes [Y3PW(CO)5] (Y = Me, Bu, Ph, OPh).[23] The
acterized or isolated. In contrast, stoichiometric amounts IR spectrum exhibits a pattern characteristic of the
of phosphenium cation 1a and complex [W(CO)5THF] in [LW(CO)5][23a] complex with three CO stretching bands at
THF at room temperature easily led to the transient chloro 1833, 1904, and 2017 cm–1.
complex identified by 31P NMR spectroscopy (δ = Slow crystallization in chloroform solution at room tem-
150.03 ppm). This latter rapidly hydrolysed to give the hy- perature gave brown crystals identified as complex 7 by X-
droxy complex 7 (Scheme 4). The 31P NMR spectrum ray analysis (Figure 5). The phosphorus atom is tetracoor-
shows a signal at δ = 128.65 ppm accompanied by satellites dinate and the N-heterobicyclic unit is nearly planar and
of tungsten with a coupling constant JWP = 386 Hz. This almost perpendicular to the W(CO)5 system. The long dis-
value is close to that observed for the similar complex tance between the phosphorus atom and the chloride anion

Scheme 4.

3360 www.eurjic.org © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Inorg. Chem. 2011, 3357–3364
Aminotroponiminate and Aminotroponate Complexes

[P+···Cl–, 3.833(2) Å] suggests that there is no interaction Conclusions


between them. The OH group of the cation is hydrogen-
bonded to the chloride anion with a distance of 2.903 Å for An easy and direct route to the first phosphenium and
the O1···Cl1 interaction. The P–N bond lengths [1.717(3) arsenium cations stabilized by N,N⬘- or N,O-chelation of a
and 1.724(3) Å] are slightly longer than those of 1a tropolone scaffold has been described. The presence of a π-
[1.697(1) and 1.708(1) Å]. The P–O bond length unsaturated backbone in the aminotroponiminate and ami-
[1.573(3) Å] is in the range of P–OH distances (1.503– notroponate ligands not only prevents the competitive γ-
1.611 Å).[24] The tungsten atom is octahedrally coordinated carbon substitution previously observed for the diketimin-
with an almost linear Cax–W–P angle of 173.33(15)°. The ate ligand but also allows complete positive charge delocal-
bond length of P–W [2.4584(9) Å] lies between the P–W ization onto the seven-membered cycles. This result was
lengths in [W(PCl3)(CO)5] [2.378(2) Å][25] and confirmed by spectroscopic studies and X-ray structural
[W(PMe3)(CO)5] [2.516(2) Å][25] and is nearly identical to data. The cycloaddition of a phosphenium cation to an o-
the P–W bond length [2.460(2) Å] of the hydroxyphosphide quinone with the formation of a phosphate ester derivative
complex [WClCp{PH(OH)R}(CO)2] (R = 2,4,6- confirms its dienophilic nature. Another interesting feature
C6H2tBu3).[26] is its ability to act as a ligand in the coordination of a tran-
sition-metal complex. A novel cationic tungsten complex,
which is isolobal with transition-metal carbene complexes,
was synthesized and structurally characterized.

Experimental Section
General: All manipulations with air-sensitive materials were per-
formed in a dry and oxygen-free atmosphere of Argon by using
standard Schlenk-line and glove-box techniques. Solvents were
purified with the MBRAUN SBS-800 purification system except
for THF, which was distilled over Na/benzophenone. NMR spectra
were recorded with the following spectrometers: 1H: Bruker Avance
II 300 (300.13 MHz); 13C: Bruker Avance II 300 (75.47 MHz) and
Avance II 500 (100.61 MHz); 31P: Bruker Avance II 300
Figure 5. Molecular structure of 7 (50 % probability level for the (121.49 MHz) with H3PO4 as reference; 19F: Bruker Avance 300
thermal ellipsoids). All H atoms (except H1) have been omitted for (282 MHz) with CFCl3 as reference. Mass spectra were measured
clarity. Selected bond lengths [Å] and angles [°]: P1–O1 1.573(3), with a Hewlett-Packard 5989A spectrometer in the electron impact
P1–N1 1.717(3), P1–N2 1.724(3), P1–W1 2.4584(9), N1–C1
mode (70 eV) and only characteristic fragments are reported. Melt-
1.372(4), N2–C2 1.368(5), W1–C16 2.025(5), W1–C18 2.049(5),
O2–C14 1.136(6), O(4)–C16 1.129(6), C1–C7 1.396(5), C1–C2 ing points were measured with a Leitz microscope or an Electro-
1.444(5), C2–C3 1.394(5), C3–C4 1.376(6), C4–C5 1.381(6), C5– thermal apparatus (capillary). IR spectra were measured with a
C6 1.384(6), C6–C7 1.380(5); O1–P1–N1 105.73(16), N1–P1–N2 Varian 640-IR FTIR spectrometer. Elemental analyses were per-
88.44(15), C1–N1–P1 114.0(2), C2–N2–P1 113.9(2), O1–P1–W1 formed at the Centre de Microanalyse de l’Ecole Nationale Supé-
107.42(11), N1–P1–W1 123.99(11), N2–P1–W1 122.87(12), C16– rieure des Ingénieurs en Arts Chimiques et Technologiques. N-Iso-
W1–C14 89.05(19), C16–W1–P1 173.33(15). propyl-2-(isopropylamino)troponimine, 2-(isopropylamino)trop-
one[30] and the phosphenium cation 1a[11] were prepared according
to literature procedures. The atom labelling used in the NMR as-
signments of 2 is given below:
Compound 7 is the first hydroxyphosphenium cation sta-
bilized both by an intramolecular donor–acceptor bond
and coordination to a transition metal. It is well known
that phosphinous acid (R2POH) is in equilibrium with its
phosphane oxide tautomeric form (R2HP=O) and that this
equilibrium is generally shifted to the phosphane oxide in
the case of the neutral compound.[27] However, phos-
phinous acid can be stabilized by CF3 substituents[28] or by Synthesis of Dichloroarsine 2
complexation with a transition metal.[29] In a series of cat-
ionic species, Cowley and co-workers reported the prepara- Method 1. By a Dehydrohalogenation Coupling Reaction: AsCl3
tion and characterization of a β-diketiminato hydroxyphos- (0.36 g, 1.96 mmol) in toluene (5 mL) was added dropwise to a
solution of (iPr2-ATI)H (0.40 g, 1.96 mmol) and Et3N (0.20 g,
phenium.[24b] The preference for the hydroxy tautomeric
1.96 mmol) in toluene (15 mL) over a period of 5 min. The mixture
form is probably due to the electronic and steric effects of was stirred for 3 h. After filtration, the volatiles were removed un-
the N,N⬘-chelating unit. In our case, the presence of the der reduced pressure and the residue was dissolved in CH2Cl2. The
aminotroponiminato moiety is not sufficient and additional saturated solution was placed in a freezer at –30 °C for 2 d to yield
coordination to tungsten seems to be required for stabiliz- yellow crystals of 2 (0.34 g, 50 %). M.p. 138–142 °C. 1H NMR
ing this form. (300 MHz, CDCl3): δ = 1.74 (d, 3JH,H = 6.6 Hz, 12 H, CHCH3),

Eur. J. Inorg. Chem. 2011, 3357–3364 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.eurjic.org 3361
L.-C. Pop, A. Castel, L. Silaghi-Dumitrescu, N. Saffon
FULL PAPER
4.40 (sept., 3JH,H = 6.6 Hz, 2 H, CHCH3), 7.24 (t, 3JH,H = 9.6 Hz, Reaction of 1a with 3,5-Di-tert-butyl-1,2-benzoquinone: A solution
1 H, 5-H), 7.30 (d, 3JH,H = 11.4 Hz, 2 H, 3-H, 7-H), 7.69 (pseudo- of 3,5-di-tert-butyl-1,2-benzoquinone (0.13 g, 0.6 mmol) in CH2Cl2
t, 3JH,H = 9.6 Hz, 2 H, 4-H, 6-H) ppm. 13C{1H} NMR (75 MHz, (2 mL) was added to a solution of 1a (0.20 g, 0.7 mmol) in CH2Cl2
CDCl3): δ = 21.62 (CHCH3), 51.00 (CHCH3), 121.99 (C-3, C-7), (4 mL). The mixture was stirred for 1 h at room temperature and
130.00 (C-5), 139.97 (C-4, C-6), 158.78 (C-2, C-8) ppm. MS (EI): then heated at 40 °C for 1 h. The mixture was kept at –30 °C for
m/z (%) = 313 (100) [M – Cl]+, 235 (95) [M – 2Cl – iPr]+. 12 h and then the resulting yellow powder was filtered and dried
under vacuum to give compound 5 (0.24 g, 77 %). M.p. 176 °C.
Method 2. By Nucleophilic Substitution Reaction Using an Amino- 31
P{1H} NMR (121 MHz, CDCl3): δ = 7.83 ppm. 1H NMR
lithiated Derivative: A 1.6 m solution of nBuLi in hexanes (300 MHz, CDCl3): δ = 1.22 [s, 9 H, C(CH3)3], 1.31 [s, 9 H,
(1.60 mmol) was slowly added at 0 °C to a solution of (iPr2- C(CH3)3], 1.51 (d, 3JH,H = 6.5 Hz, 6 H, CHCH3), 1.54 (d, 3JH,H =
ATI)H (0.29 g, 1.42 mmol) in diethyl ether (10 mL). The yellow 6.7 Hz, 6 H, CHCH3), 4.63 (sept., 3JH,H = 6.7 Hz, 2 H, CHCH3),
solution was warmed to room temperature, stirred for 30 min and 7.04 (d, 4JH,H = 2.2 Hz, 1 H, C6H2), 7.47 (d, 4JH,H = 2.2 Hz, 1 H,
slowly added to a cooled (–75 °C) solution of AsCl3 (0.26 g, C6H2), 7.78 (t, 3JH,H = 9.4 Hz, 1 H, C7H5), 8.16–8.30 (m, 4 H,
1.42 mmol) in diethyl ether (5 mL). The reaction mixture was C7H5) ppm. 13C{1H} NMR (75 MHz, CDCl3): δ = 19.82 (d, 3JC,P
warmed to room temperature and stirred for 3 h. After filtration, = 1.5 Hz, CHCH3), 21.21 (CHCH3), 29.50 and 31.55 (C-CH3),
the remaining solid was washed with diethyl ether and dried under 34.49 and 35.41 (C-CH3), 50.19 (d, 2JC,P = 3.3 Hz, CHCH3), 113.63
vacuum, leading to a yellow powder identified as 2 (0.28 g, 57 %). and 120.17 (C6H2, CH), 124.89 (d, 3JC,P = 10.6 Hz, C-3, C-7),
135.62 (C-5), 137.95 and 141.05 (Cquat of C6H2), 139.61 (d, 2JC,P =
Dichloroarsine 3
8.0 Hz) and 143.64 (d, 2JC,P = 8.7 Hz, Cquat of C6H2), 144.31 (C-
Method 1. By a Dehydrohalogenation Coupling Reaction: By using 4, C-6), 151.24 (d, 2JC,P = 18.0 Hz, C-2, C-8) ppm. MS (CI/NH3):
the same procedure as described for 2, from (iPr-AT)H (0.40 g, m/z (%) = 490 (14) [M – Cl + H]+. C27H39Cl2N2O2P (524.21): calcd.
2.45 mmol), Et3N (0.25 g, 2.45 mmol) and AsCl3 (0.44 g, C 61.71, H 7.48, N 5.33; found C 60.94, H 7.90, N 5.88.
2.45 mmol) in toluene (15 mL) a yellow powder identified as 3 was Presumably due to the adventitious presence of water in the reac-
obtained. Crystallization from dichloromethane at –30 °C gave 3 tion mixture, a very slow recrystallization at –30 °C from CH2Cl2
as yellow crystals (0.42 g, 56 %). M.p. 106 °C (dec.). 1H NMR gave product 6 (0.06 g, 18 %). M.p. 140–142 °C. 31P{1H} NMR
(300 MHz, CDCl3): δ = 1.74 (d, 3JH,H = 6.4 Hz, 6 H, CHCH3), (121 MHz, CDCl3): δ = –3.33 ppm. 1H NMR (300 MHz, CDCl3):
4.36 (sept., 3JH,H = 6.7 Hz, 1 H, CHCH3), 7.40–7.50 (m, 2 H, δ = 1.13 [s, 9 H, C(CH3)3], 1.28 [s, 9 H, C(CH3)3], 1.48 (d, 3JH,H =
C7H5), 7.65–7.75 (m, 2 H, C7H5), 7.85–7.95 (m, 1 H, C7H5) ppm. 6.4 Hz, 12 H, CHCH3), 3.89–4.00 (m, 2 H, CHCH3), 6.81–7.05 (m,
13
C{1H} NMR (75 MHz, CDCl3): δ = 21.54 (CHCH3), 51.20 5 H, C6H2 and C7H5), 7.37–7.44 (m, 2 H, C7H5) ppm. 13C{1H}
(CHCH3), 124.33 (C-7), 128.25 (C-3), 133.40 (C-5), 140.69 and NMR (75 MHz, CDCl3): δ = 20.35 (CHCH3), 28.55 and 30.51 (C-
142.56 (C-4, C-6), 150.00 (C-2), 158.13 (C-8) ppm. MS (EI): m/z CH3), 33.25 and 34.11 (C-CH3), 46.82 (CHCH3), 115.19 and 124.27
(%) = 272 (100) [M – Cl]+. (CH of C6H2), 125.04 (C-5), 127.20 and 128.00 (C-3, C-7), 135.70
and 140.28 (Cquat of C6H2), 136.81 and 142.77 (Cquat of C6H2),
Method 2. By Nucleophilic Substitution Reaction Using an Amino-
138.74 (C-4, C-6), 149.58 (C-2, C-8) ppm. MS (EI): m/z (%) = 506
lithiated Derivative: A 1.6 m solution of nBuLi in hexanes
(3) [M – 1]+.
(0.85 mmol) was slowly added at 0 °C to a solution of 2-(isopro-
pylamino)tropone (0.20 g, 1.23 mmol) in diethyl ether (15 mL). The Reaction of 1a with [W(CO)5THF]: A solution of [W(CO)5THF]
yellow solution was warmed to room temperature, stirred for (0.44 g, 1.10 mmol), freshly prepared by the irradiation of
30 min, and slowly added to a diethyl ether (5 mL) solution of [W(CO)6] in THF (20 mL), was added to a stirred suspension of 1
AsCl3 (0.22 g, 1.23 mmol) at –75 °C. The reaction mixture was (0.31 g, 1.00 mmol) in THF (6 mL). The reaction mixture was
warmed to room temperature and stirred for 3 h. After filtration, stirred for 6 h at room temperature and then the solvent was re-
the remaining solid was washed with diethyl ether (3 ⫻ 10 mL). The moved from the solution leaving an orange-brown powder. All
volatiles were removed under reduced pressure to yield a yellow attempts to isolate the chloro complex from the crude reaction mix-
powder. 1H NMR analysis showed the formation of 3 (50 %) and ture caused its progressive transformation into 7. A very slow
the presence of the starting ligand, 2-(isopropylamino)tropone recrystallization at room temperature from CDCl3 gave crystals of
(50 %). 7 (0.10 g, 16 %). 31P{1H} NMR (121 MHz, CDCl3): δ = 128.65
(JWP = 386 Hz) ppm. 1H NMR (300 MHz, CDCl3, 25 °C): δ = 1.74
Arsenium Cation 4: Me3SiOTf (0.13 g, 0.59 mmol) was added drop- (d, 3JH,H = 7.0 Hz, 6 H, CHCH3), 1.84 (d, 3JH,H = 7.1 Hz, 6 H,
wise to a solution of 2 (0.20 g, 0.57 mmol) in CH2Cl2 (5 mL) over CHCH3), 4.72–4.90 (m, 2 H, CHCH3), 7.50–7.60 (m, 1 H), 7.68–
a period of 5 min. The orange solution was stirred for 30 min. After 7.95 (m, 4 H) ppm. 13C{1H} NMR (75 MHz, CDCl3): δ = 20.63
evaporation of the solvent under reduced pressure, the precipitate (d, 3JC,P = 3.7 Hz, CHCH3), 49.05 (d, 2JC,P = 13.2 Hz, CHCH3),
was dissolved in CH2Cl2 and cooled for 24 h at –30 °C. Compound 122.43 (d, 3JC,P = 4.7 Hz, C-3, C-7), 133.22 (C-5), 143.37 (C-4, C-
4 was then isolated as yellow crystals (0.15 g, 59 %). M.p. 160– 6), 154.06 (C-2, C-8), 195.20 (d, 2JC,P = 9.3 Hz, COeq), 196.18 (d,
163 °C. 19F{1H} NMR (282 MHz, CDCl3): δ = –78.44 ppm. 1H 2
JC,P = 8.6 Hz, COax) ppm. IR: ν̃(CO) = 1833, 1904, 2017 cm–1.
NMR (300 MHz, CDCl3): δ = 1.58 (d, 3JH,H = 6.4 Hz, 6 H,
X-ray Structural Determination: Structural data were collected with
CHCH3), 1.81 (d, 3JH,H = 6.4 Hz, 6 H, CHCH3), 4.55 (sept., 3JH,H
a Bruker-AXS SMART APEX II diffractometer with Mo-Kα radia-
= 6.5 Hz, 2 H, CHCH3), 7.56 (t, 3JH,H = 9.5 Hz, 1 H, 5-H), 7.63
tion (λ = 0.71073 Å) at low temperatures (173 K for 6, 193 K for
(d, 3JH,H = 11.1 Hz, 2 H, 3-H, 7-H), 7.95 (pseudo-t, 3JH,H = 9.6 Hz,
2, 3, 4 and 7) using an oil-coated shock-cooled crystal. The struc-
2 H, 4-H, 6-H) ppm. 13C{1H} NMR (75 MHz, CDCl3): δ = 20.81
tures were solved by direct methods[31] and all non-hydrogen atoms
(CHCH3), 51.16 (CHCH3), 118.40 (q, JCF = 317.4 Hz, CF3), 124.46
were refined anisotropically by using the least-squares method on
(C-3, C-7), 134.08 (C-5), 141.52 (C-4, C-6), 157.95 (C-2, C-8) ppm.
F2.[32]
MS (EI): m/z (%) = 427 (22) [M – Cl]+, 313 (100) [M – OSO2-
CF3]+. C14H19AsClF3N2O3S (462.74): calcd. C 36.33, H 4.14, N 2: C13H21AsCl2N2O, M = 367.14, monoclinic P21/c, a =
6.06; found C 35.92, H 3.91, N 5.90. 11.6250(13), b = 14.7685(18), c = 9.4324(10) Å, β = 98.6776(6)°, V

3362 www.eurjic.org © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Inorg. Chem. 2011, 3357–3364
Aminotroponiminate and Aminotroponate Complexes

= 1600.9(3) Å3, Z = 4, 20953 reflections (4796 independent, Rint [6] D. Gudat, T. Gans-Eichler, M. Nieger, Chem. Commun. 2004,
= 0.0309) were collected. Largest diff. peak and hole: 0.339 and 2434–2435.
–0.277 e Å–3, R1 [for I ⬎ 2σ(I)] = 0.0252 and wR2 (all data) = 0.0584. [7] D. Vidovic, Z. Lu, G. Reeske, J. A. Moore, A. H. Cowley,
Chem. Commun. 2006, 3501–3503.
3: C10H12AsCl2NO, M = 308.03, triclinic P1̄, a = 9.6348(2), b = [8] a) T. Kaukorat, I. Neda, R. Schmutzler, Coord. Chem. Rev.
11.0251(2), c = 13.3230(2) Å, α = 98.4130(10), β = 100.5440(10), γ 1994, 137, 53–107; b) P. Kilian, A. M. Z. Slawin, Dalton Trans.
= 113.0890(10)°, V = 1242.12(4) Å3, Z = 4, 19949 reflections (6121 2007, 3289–3296; c) S. P. Green, C. Jones, G. Jin, A. Stasch,
independent, Rint = 0.0418) were collected. Largest diff. peak and Inorg. Chem. 2007, 46, 8–10; d) B. Lyhs, S. Schulz, U. Westphal,
hole: 0.474 and –0.502 e Å–3, R1 [for I ⬎ 2σ(I)] = 0.0314 and wR2 D. Bläser, R. Boese, M. Bolte, Eur. J. Inorg. Chem. 2009, 2247–
(all data) = 0.0692. 2253.
[9] a) C. W. Schultz, R. W. Parry, Inorg. Chem. 1976, 15, 3046–
4: C14H19AsClF3N2O3S, M = 462.74, monoclinic P21/n, a = 3050; b) M. G. Thomas, C. W. Schultz, R. W. Parry, Inorg.
11.5208(6), b = 8.7814(4), c = 19.0636(9) Å, β = 95.613(3)°, V = Chem. 1977, 16, 994–1001; c) N. Burford, T. S. Cameron, D. J.
1919.39(16) Å3, Z = 4, 5938 reflections [4763 independent, R(int) LeBlanc, P. Losier, S. Sereda, G. Wu, Organometallics 1997, 16,
= 0.0590] were collected. Largest diff. peak and hole: 0.604 and 4712–4717; d) N. Burford, T. S. Cameron, P. J. Ragona, J. Am.
–0.560 e Å–3, R1 [for I ⬎ 2σ(I)] = 0.0531 and wR2 (all data) = 0.1488. Chem. Soc. 2001, 123, 7947–7948; e) A. Karaçar, M. Freytag,
H. Thönnessen, P. G. Jones, R. Bartsch, R. Schmultzler, J. Or-
6: C29H44Cl5N2O3P, M = 676.88, triclinic P1̄, a = 10.0784(15), b = ganomet. Chem. 2002, 643–644, 68–80; f) N. Burford, P. J. Ra-
10.7719(17), c = 17.733(3) Å, α = 73.226(3), β = 78.009(3), γ = gona, R. McDonald, M. J. Ferguson, J. Am. Chem. Soc. 2003,
70.202(3)°, V = 1721.3(5) Å3, Z = 2, 10459 reflections [4832 inde- 125, 14404–14410.
pendent, R(int) = 0.1278] were collected. Largest diff. peak and [10] a) K. A. Porter, A. C. Willis, J. Zank, S. B. Wild, Inorg. Chem.
2002, 41, 6380–6386; b) N. L. Kilah, S. Petrie, R. Stranger, J. W.
hole: 0.233 and –0.210 e Å–3, R1 [for I ⬎ 2σ(I)] = 0.0461 and wR2
Wielandt, A. C. Willis, S. B. Wild, Organometallics 2007, 26,
(all data) = 0.0904. 6106–6113; c) N. L. Kilah, M. L. Weir, S. B. Wild, Dalton
7: C18H20ClN2O6PW, M = 610.63, orthorhombic Pbca, a = Trans. 2008, 2480–2486.
9.5289(2), b = 13.1917(2), c = 35.3309(6) Å, V = 4441.18(14) Å3, Z [11] L.-C. Pop, N. Katir, A. Castel, L. Silaghi-Dumitrescu, F.
= 8, 75313 reflections [5470 independent, R(int) = 0.0557] were Delpech, I. Silaghi-Dumitrescu, H. Gornitzka, D. MacLeod-
Carey, N. Saffon, J. Organomet. Chem. 2009, 694, 1562–1566.
collected. Largest diff. peak and hole: 0.969 and –1.617 e Å–3, R1
[12] W. Czado, S. Rabe, U. Mueller, Z. Naturforsch. B: Chem. Sci.
[for I ⬎ 2σ(I)] = 0.0325and wR2 (all data) = 0.0648.
1999, 54, 288–290.
CCDC-811691 (for 2), -811692 (for 3), -811693 (for 4), -811694 (for [13] N. Burford, T. M. Parks, B. W. Royan, J. F. Richardson, P. S.
6) and -81695 (for 7) contain the supplementary crystallographic White, Can. J. Chem. 1992, 70, 703–709.
data for this paper. These data can be obtained free of charge [14] a) M. Veith, B. Bertsch, Z. Anorg. Allg. Chem. 1988, 7, 557;
from The Cambridge Crystallographic Data Centre at b) D. J. Willliams, M. G. Newton, K. J. Wynne, Cryst. Struct.
Commun. 1977, 6, 167–170.
www.ccdc.cam.ac.uk/data_request/cif.
[15] a) J. Kopf, K. Von Deuten, G. Klar, Inorg. Chim. Acta 1980,
38, 67–69; b) A. L. Rheingold, A.-J. DiMalo, Organometallics
1986, 5, 393–394.
Acknowledgments [16] a) C. A. Caputo, J. T. Price, M. C. Jennings, R. McDonald,
N. D. Jones, Dalton Trans. 2008, 3461–3469; b) C. K. SooHoo,
We thank the ECONET Program (no. 18824SD). L.-C. P. is grate- S. G. Baxter, J. Am. Chem. Soc. 1983, 105, 7443–7444.
ful to The French Ministry of Foreign and European Affairs for [17] A. Bond, M. Green, S. C. Pearson, J. Chem. Soc. B 1968, 929–
an Eiffel Excellence Scholarship (no. 625658J) and the Romanian 931.
Ministry of Education Research, Youth and Sports (grant no. PN- [18] F. Ramirez, M. Nowakowski, J. F. Marecek, J. Am. Chem. Soc.
II-ID-564). 1977, 99, 4515–4517.
[19] L. Pauling, Nature of the Chemical Bond, Cornell University
Press, Ithaca, NY, 1960.
[1] a) P. W. Roesky, Chem. Soc. Rev. 2000, 29, 335–345; b) H. V. R. [20] L.-C. Pop, D. MacLeod Carey, A. Munoz-Castro, L. Silaghi-
Dias, Z. Wang, W. Jin, Coord. Chem. Rev. 1998, 176, 67–86. Dumitrescu, A. Castel, R. Arratia-Perez, Polyhedron 2011, 30,
[2] a) D. Pappalardo, M. Mazzeo, P. Montefusco, C. Tedesco, C. 841–845.
Pellecchia, Eur. J. Inorg. Chem. 2004, 1292–1298; b) S. Dehnen, [21] D. Gudat, M. Nieger, E. Niecke, J. Chem. Soc., Dalton Trans.
M. R. Bürgstein, P. W. Roesky, J. Chem. Soc., Dalton Trans.
1989, 693–700.
1998, 2425–2430; c) F. A. Hicks, M. Brookhart, Organometal-
[22] E. Gross, K. Jörg, K. Fiederling, A. Göttlein, W. Malish, R.
lics 2001, 20, 3217–3219; d) N. Meyer, K. Löhnwitz, A. Zulys,
P. W. Roesky, M. Dochnahl, S. Blechert, Organometallics 2006, Boese, Angew. Chem. Int. Ed. Engl. 1984, 23, 738–739.
25, 3730–3734; e) S. Datta, P. W. Roesky, S. Blechert, Organo- [23] a) M. S. Davies, R. K. Pierens, M. J. Aroney, J. Organomet.
metallics 2007, 26, 4392–4394; f) N. Meyer, R. Rüttinger, P. W. Chem. 1993, 458, 141–146; b) G. M. Bodner, Inorg. Chem.
Roesky, Eur. J. Inorg. Chem. 2008, 1830–1833. 1975, 14, 2694–2699; c) W. Buchner, W. A. Schenk, Inorg.
[3] a) M. Sanchez, M. R. Mazières, L. Lamandé, R. Wolf, Phos- Chem. 1984, 23, 132–137; d) G. G. Mather, A. Pidcock, J.
phorous Chemistry, Thieme, Stuttgart, 1990, p. 129; b) D. Gu- Chem. Soc. A 1970, 1226–1229.
dat, Coord. Chem. Rev. 1997, 163, 71–106. [24] a) M. J. Pilkington, A. M. Z. Slawin, D. J. Williams, J. D. Wool-
[4] a) C. J. Carmalt, V. Lomeli, B. G. McBurnett, A. H. Cowley, ins, Main Group Chem. 1995, 1, 145–151; b) Z. Lu, M. Find-
Chem. Commun. 1997, 2095–2096; b) G. Reeske, C. R. Hoberg, later, A. H. Cowley, Chem. Commun. 2008, 184–186.
N. J. Hill, A. H. Cowley, J. Am. Chem. Soc. 2006, 128, 2800– [25] M. S. Davies, M. J. Aroney, I. E. Buys, T. W. Hambley, J. L.
2801; c) H. A. Spinney, I. Korobkov, G. A. DiLabio, G. P. A. Calvert, Inorg. Chem. 1995, 34, 330–336.
Yap, D. S. Richeson, Organometallics 2007, 26, 4972–4982. [26] M. Alonso, M. A. Alvarez, M. E. Garcia, D. Garcia-Vivo,
[5] a) N. Burford, T. M. Parks, B. W. Royan, B. Borecka, T. S. M. A. Ruiz, Inorg. Chem. 2010, 49, 8962–8976.
Cameron, J. F. Richardson, E. J. Gabe, R. Hynes, J. Am. Chem. [27] D. E. C. Corbridge in Phosphorus: An Outline of its Chemistry,
Soc. 1992, 114, 8147–8153; b) T. Gans-Eichler, D. Gudat, M. Biochemistry and Uses, 5th ed., Elsevier Science, Amsterdam,
Nieger, Heteroat. Chem. 2005, 16, 327–338. 1995, p. 336.

Eur. J. Inorg. Chem. 2011, 3357–3364 © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.eurjic.org 3363
L.-C. Pop, A. Castel, L. Silaghi-Dumitrescu, N. Saffon
FULL PAPER
[28] a) B. Hoge, S. Neufeind, S. Hettel, W. Wiebe, C. Thösen, J. [30] H. V. R. Dias, W. Jin, R. E. Ratcliff, Inorg. Chem. 1995, 34,
Organomet. Chem. 2005, 690, 2382–2387; b) B. Hoge, P. Garcia, 6100–6105.
H. Willner, H. Oberhammer, Chem. Eur. J. 2006, 12, 3567– [31] G. M. Sheldrick, Acta Crystallogr., Sect. A 1990, 46, 467–473.
3574. [32] G. M. Sheldrick, SHELXL-97, Program for Crystal Structure
[29] a) J. Chatt, B. T. Heaton, J. Chem. Soc. A 1968, 2745–2757; b) Refinement, University of Göttingen, Göttingen, 1997.
C. S. Kraihanzel, C. M. Bartish, J. Am. Chem. Soc. 1972, 94, Received: February 7, 2011
3572–3575; c) E. Lindner, B. Schilling, Chem. Ber. 1977, 110, Published Online: June 28, 2011
3266–3271.

3364 www.eurjic.org © 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Inorg. Chem. 2011, 3357–3364

S-ar putea să vă placă și