Vous êtes sur la page 1sur 21

Accepted Manuscript

The synthesis of PNP-supported low-spin nitro manganese(I) carbonyl com-


plexes

Aaron M. Tondreau, James M. Boncella

PII: S0277-5387(16)30079-1
DOI: http://dx.doi.org/10.1016/j.poly.2016.04.007
Reference: POLY 11933

To appear in: Polyhedron

Received Date: 26 January 2016


Accepted Date: 6 April 2016

Please cite this article as: A.M. Tondreau, J.M. Boncella, The synthesis of PNP-supported low-spin nitro
manganese(I) carbonyl complexes, Polyhedron (2016), doi: http://dx.doi.org/10.1016/j.poly.2016.04.007

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
The synthesis of PNP-supported low-spin nitro manganese(I) carbonyl complexes.

Aaron M. Tondreau, a James M. Boncellaa


a.
LANL

boncella@lanl.gov

Abstract.
The coordination chemistry of Mn(CO)5Br was investigated with a series of
iPr
PNP-pincer ligands. The ligands PONOP (iPrPONOP = 2,6-
iPr H iPr H
bis(diisopropylphosphinito)pyridine) and PN P ( PN P = HN{CH2CH2(PiPr2)}2)
gave the desired organometallic manganese complexes (iPrPONOP)Mn(CO)2Br and
(iPrPNHP)Mn(CO)2Br, respectively, upon chelation to Mn(CO)5Br. The reactivity of
iPr
PNNNP (iPrPNNNP = N,N′-bis(diisopropylphosphino)-2,6-diaminopyridine) with
Mn(CO)5Br yielded a pair of products, [(iPrPNNNP)Mn(CO)3][Br] and
(iPrPNNNCO)Mn(CO)3. The formation of the asymmetric chelate arises from a formal
loss of iPr2PBr and C-N bond formation from a carbonyl ligand and NH, yielding a
Mn(I) amide core. The nitration reactions of (iPrPONOP)Mn(CO)2Br and
(iPrPNHP)Mn(CO)2Br were carried out using silver nitrite, yielding the nitro compounds
(iPrPONOP)Mn(CO)2(NO2) and (iPrPNHP)Mn(CO)2(NO2), respectively. The analogous
iron complex (iPrPONOP)Fe(CO)Cl2 was nitrated under the same conditions to yield the
salt pair [(iPrPONOP)Fe(CO)2][FeCl3NO]. This reactivity underlines the difference
between iso-valent iron and manganese centers. The manganese complexes
(iPrPONOP)Mn(CO)2(NO2) and (iPrPNHP)Mn(CO)2(NO2) were ineffective as oxygen
atom transfer reagents for a variety of substrates.

1. Introduction.
Manganese is a bioactive metal present in a variety of metalloenzymes. 1 Often
the manganese center catalyzes oxidative transformations, and specifically we are
interested in the reactivity of oxalate decarboxylase, wherein oxalate is converted to
carbon dioxide and formate. 2 As a transition metal catalyst in the laboratory,
manganese is also well known for its oxidative transformations. 3 Metal nitrites have
found use as oxidants in both organic synthesis as well as transition metal coordination
chemistry.4 We have been investigating the synthesis of a PNP supported manganese
nitro compound for the express purpose of oxygen atom transfer from the nitro group to
relevant substrates.

Figure 1. The three ligands explored in this study and their shorthand names are shown above.
Successful catalysis and reaction chemistry has been observed with a variety of
tridentate PNP ligand supported iron complexes. We have chosen three related PNP
ligands (Figure 1) with which to investigate the coordination of Mn(CO)5Br. The
oxygen-linked iPrPONOP ligand has been used as a support for iron complexes by
different groups and catalytically competent coordination complexes for the transfer
hydrogenation of aldehydes have been developed. 5 Kirchner has investigated the
coordination and reactivity of iPrPNNNP supported iron complexes and has found
certain complexes catalytically competent for ketone hydrogenation.6 Recently, iPrPNHP
complexes of iron have been successfully used for the dehydrogenation of formic acid
and methanol, 7 the hydrogenation of esters to alcohols, 8 the hydrogenation and
dehydrogenation of N-heterocycles, 9 alcohols and ketones, 10 as well as the
hydrogenation of aromatic and aliphatic nitriles. 11 Ozerov has investigated diarylamido-
based PNP ligands on manganese carbonyl complexes along with subsequent ligand
centered reactivity.12 Certain pentadentate bis(imino)pyridine derived PNP ligands have
also been used to synthesize manganese complexes that show high catalytic activity for
the hydrosilylation of aldehydes and ketones. 13 Although the PNP ligand class has
found wide-ranging use with iron, tridentate PNP-supported manganese complexes
remain rare in the literature.
During our initial efforts to effect the conversion of oxalic acid into its
component parts, carbon dioxide and dihydrogen, one course of experiments led to the
investigation of PNP supported manganese carbonyl compounds. Within this line of
investigation we found a rational route to an uncommon class of nitro-manganese
complexes. The paucity of such complexes coupled with the lure of ligand-mediated
reactivity, specifically oxygen atom transfer reactions from a coordinated nitro group,
prompted us to investigate these nitro-manganese coordination complexes. Reported
herein are the reactions of ligation, nitration, and attempts at deoxygenation of a series
of PNP supported manganese carbonyl bromide complexes derived from Mn(CO)5Br.
Silver nitrite was used as the nitro-source due to the facile removal of AgBr and fast
reaction times. Multiple attempts were made to deoxygenate the nitro ligands, but they
proved resilient to the reducing agents that were employed.

2. Experimental

General Considerations.
All air- and moisture-sensitive manipulations were carried out using standard
Schlenk techniques or in an MBraun dry box containing a purified nitrogen atmosphere.
THF, diethyl ether, toluene, and N-hexane were dried on molecular sieves and shaved
sodium before use. N,N,-dimethyl formamide (DMF) and DMSO were purchased as
anhydrous solvents from Fisher Scientific and stored over 4 Å molecular sieves. THF-
d8, DMSO-d6, and C6D6 were purchased from Cambridge Isotope Laboratories and
dried over 4 Å molecular sieves. The chemicals: Mn(CO)5Br, AgNO2, iPrPNHP, FeCl2,
PMe3, Phenylsilane, 1,2-dihydroanthracene, and N-butyllithium were ordered from
Fisher Scientific and were used as received. The compounds: iPrPONOPFe(CO)Cl2,5b
iPr
PONOP,14 iPrPNNNP,6d were synthesized according to published procedures.
1
H NMR, 13C NMR, and 31P NMR spectra were recorded on a Bruker Avance
400 MHz spectrometer operating at 400.132 MHz, 100.627 MHz, and 161.978 MHz,
respectively. All 1H and 13C NMR chemical shifts are reported relative to SiMe4 using
the 1H (residual in the deuterated solvents) and 13C chemical shifts of the solvent as a
secondary standard. Paramagnetically shifted peaks are listed with the peak width at
half height (Hz). Infrared spectra were collected on a Thermo Scientific Nicolet iS10
spectrometer equipped with a Smart Omni transmission tool for the collection.
Elemental analysis was performed by Atlantic Microlab, inc. Several EA’s have values
outside of the accepted values; therefore NMR spectra are provided (SI) that label
impurities.
Single crystals suitable for X-ray diffraction were coated with n-paratone (dried
under reduced pressure overnight at 100 °C) oil in a drybox, placed on a nylon loop and
then transferred to the goniometer head of a Bruker AXS APEX II diffractometer
equipped with a graphite-monochromatized molybdenum Kα X-ray tube ( = 0.71073
Å) and an APEX II CCD detector (complexes ) or to a Bruker D8 Quest equipped with
a graphite-monochromatized molybdenum Kα X-ray tube ( = 0.71073 Å) and a
CMOS detector (complexes ).15 A hemisphere routine was used for data collection and
determination of lattice constants. The space group was identified and the data were
processed using the Bruker SAINT+ program and corrected for absorption using
SADABS.16 The structures were solved using direct methods (SHELXS) completed by
subsequent Fourier synthesis and refined by full-matrix least-squares procedures. 17
Complexe 7 was solved using Olex2 software with solvent suppression due to the
presence of a highly disordered diethyl ether solvent molecule.18

Synthesis of (iPrPONOP)Mn(CO)2Br (1).


To a 100 mL Schlenk flask was added 25 mL of fluorobenzene, 2.00 g (7.275 mmol) of
MnBr(CO)5, and a stir bar. A syringe containing 5 mL of fluorobenzene and 2.620 g
(7.640 mmol) of iPrPONOP was used to add the ligand to the stirring manganese
solution. Upon addition of the ligand, the solution evolved gas. The reaction was stirred
at room temperature until gas generation ceased and the MnBr(CO)5 completely
dissolved, at which point the solution was brought to reflux for approximately ten
minutes. The solution was cooled to room temperature and volatiles were removed
under reduced pressure. The product was isolated by adding hexane to the residue and
mobilizing the solid, followed by isolation of the orange powder on a glass frit. This
yielded 3.66 g (94%) of (iPrPONOP)Mn(CO)2Br. Analysis for C19H31BrMnNO4P2:
Calculated: C = 42.72%, H = 5.85%, N = 2.62%. Found: C = 43.04%, H = 5.93%, N =
2.51%. 1H NMR (C6D6): 6.61 (t, 1H, J= 7.9 Hz, p-Py), 6.10 (d, 2H, J= 7.9 Hz, m-Py),
4.05 (ddt, 2H, J= 14.5, 11.9, 7.2 Hz, iPr-CH), 2.72 (ddt, 2H, J= 14.3, 12.3, 7.9 Hz, iPr-
CH), 1.52-1.36 (m, 6H, iPr-CH3), 1.36-1.15 (m, 18H, iPr-CH3). 1H NMR (THF-d8):
7.72 (t, 1H, J= 7.9 Hz, p-Py), 6.76 (d, 2H, J= 7.9 Hz, m-Py), 3.86 (ddt, 2H, J= 14.5,
11.9, 7.2 Hz, iPr-CH), 3.05 (ddt, 2H, J= 14.3, 12.3, 7.9 Hz, iPr-CH), 1.56-1.40 (m,
18H, iPr-CH3), 1.39-1.27 (m, 6H, iPr-CH3). 31P NMR (C6D6): 231.57. 31P NMR (THF-
d8): 230.38. 13C{1H} NMR (THF-d8): 228.09, 225.74, 163.59 (t, J= 5.4 Hz), 142.15,
102.43 (t, J= 2.3Hz), 29.39 (t, 5.9 Hz), 28.69 (t, 9.4 Hz), 16.91 (t, 4.0 Hz), 16.82 (t, 4.4
Hz), 16.42, 15.65 (t, J= 1.4 Hz). IR (KBr, ν CO): 1866, 1946 cm-1.

Synthesis of iPrPNHPMn(CO)2Br (2).


To a 100 mL Schlenk flask was added 25 mL of fluorobenzene, 2.25 g (8.185 mmol) of
MnBr(CO)5, and a stir bar. A syringe containing 5 mL of fluorobenzene and 2.50 g
(8.185 mmol) of bis[(2-diisopropylphosphino)ethyl]amine was used to add the ligand to
the stirring manganese solution. Upon addition of the ligand, the solution evolved gas,
presumably CO. The reaction was stirred at room temperature until gas generation
ceased, at which point the solution was brought to reflux for approximately ten minutes.
The solution was cooled to room temperature and volatiles were removed under
reduced pressure. The product was isolated by adding hexane to the residue and
mobilizing the solid, followed by isolation of the orange powder on a glass frit. This
yielded 3.49g (86%) of iPrPNHPMn(CO)2Br. A second batch of product could be
isolated by removing the volatiles of the mother liquor and re-mobilizing with hexane,
followed by isolation of the orange powder on a glass frit to yield an additional 0.330 g
for an overall 94% yield. Analysis for C18H37BrMnNO2P2: Calculated: C = 43.56%, H
= 7.51%, N = 2.82%. Found: C = 43.71%, H = 7.40%, N = 2.75%. 1H NMR (C6D6):
3.24 (ddt, 2H, J= 13.6, 9.7, 4.9 Hz), 2.96 (t, 1H, N-H, J= 12.8 Hz), 2.59 (q, 2H, J=
13.6, 12.7 Hz), 2.23 (dtt, 2H, J= 16, 8.7, 4.3 Hz), 1.97 (q, 2H, J= 12.6 Hz), 1.68 (dq,
2H, J= 14.4, 3.9 Hz), 1.55-1.47 (m, 6H, iPr-CH3), 1.37-1.27 (m, 6H, iPr-CH3), 1.27-
1.15 (m, 6H, iPr-CH3), 1.14-1.04 (m, 6H, iPr-CH3). 31P NMR (C6D6): 81.22. 13C{1H}
NMR (C6D6): 232.72, 229.62, 52.77 (t, J= 5.2 Hz), 27.23 (t, 5.9 Hz), 26.17 (t, 9.4 Hz),
24.49 (t, 9.42 Hz), 20.38, 20.20, 18.93, 18.49 (t, J= 2.3 Hz). IR (KBr, ν CO): 1824,
1916 cm-1.

Synthesis of (iPrPNNNCO)Mn(CO)3 (3) and [(iPrPNNNP)Mn(CO)3][Br] (4).


A 100 mL Schlenk flask was charged with 25 mL of fluorobenzene, 1.00 g (3.638
mmol) of MnBr(CO)5, and a stir bar. Under nitrogen flow, 1.245 g (3.640 mmol) of
iPr
PNNNP was added as a solid. The reaction mixture immediately began to evolve gas
and iPrPNNNP dissolved quickly, while MnBr(CO)5 dissolved after roughly one hour.
The reaction was stirred at room temperature overnight and the formed off-white solid
was collected on a glass frit and washed with 2x10 mL of fluorobenzene followed by
2x10 mL of N-hexane. This yielded a roughly 50:50 mixture of 3 and 4.

Isolation of 3.
Small quantities of (iPrPNNNCO)Mn(CO)3 could be isolated in pure form by stirring
the mixture of 3 and 4 in 50 mL of THF warmed to 60 °C followed by a hot filtration of
the mixture. Concentration of the THF down to 10 mL, followed by cooling to -30 °C
gave an off white powder of (iPrPNNNCO)Mn(CO)3 in a yield of 0.025 g.
Alternatively, the mixture of 3 and 4 was extracted with warm DMF (7 mL) and filtered
on a glass frit. The solid was washed with 3 mL of DMF, followed by 2 x 5 mL of
benzene then 2 x 5 mL of diethyl ether. The mother liquor was concentrated under
reduced pressure to remove benzene and ether. The yellow mother liquor (~10 mL) was
then mixed with 3 mL of benzene and 3 mL of N-hexane and cooled to -30 °C. The
white solid that formed was filtered off and washed with benzene. The remaining
mother liquor was removed to a Schlenk line in a Schlenk flask and the DMF was
removed with heating. The residue was mobilized with THF and filtered, yielding 0.463
g (64% of ½ of the starting manganese) of analytically pure (iPrPNNNCO)Mn(CO)3.
Analysis for C15H19MnN3O4P: Calculated: C = 46.05%, H = 4.90%, N = 10.75%.
Found: C = 46.56%, H = 5.92%, N = 8.09%. 1H NMR (C6D6): (DMSO-d6): 10.10 (s,
1H, CO-NH-Py), 8.48 (d, 1H, P-NH-Py, J= 4.5 Hz), 7.41 (t, 1H, p-Py, J= 8.0 Hz), 6.27
(d, 1H, m-Py, J= 7.9 Hz), 6.17 6.27 (d, 1H, m-Py, J= 7.8 Hz), 2.39 (dp, 2H, iPr-CH, J=
10.8, 7.0 Hz), 1.28-1.08 (m, 12H, iPr-CH3). 31P NMR (DMSO-d6): 144.49. 13C{1H}
NMR (DMSO-d6): 225.75 (d, J= 11.0 Hz), 223.74, 217.53 (d, J= 12.0 Hz), 161.13 (d,
J= 12.6 Hz), 157.16 (d, J= 3.7 Hz), 140.19, 98.20 (d, J= 6.5 Hz), 96.51, 31.16 (d, J=
23.5 Hz), 18.58 (d, J= 6.0 Hz), 18.42 (d, J= 2.5 Hz). IR (KBr, ν CO): br. 1977-1821
cm-1.

Isolation of 4.
The solid that formed from the DMF solution was washed successively with THF (2
x10 mL), diethyl ether (2 x 30 mL), and finally N-hexane (2 x 10 mL). The solid was
kept under reduced pressure to remove volatiles. This left an off-white powder (0.890
g, 87% from ½ of the starting manganese) that was identified as
[(iPrPNNNP)Mn(CO)3][Br]. Analysis for C20H33BrMnN3O3P2: Calculated: C =
42.87%, H = 5.94%, N = 7.50%. Found: C = 47.29%, H = 6.12%, N = 7.11%. 1H NMR
(DMSO-d6): 9.09 (s, 2H, N-H), 7.48 (t, 1H, p-Py, J= 8.0 Hz), 6.45 (d, 2H, m-Py, J= 8.0
Hz), 2.81-2.70 (m, 4H, iPr-CH), 1.41 (dd, 12H, iPr-CH3, J= 15.7, 6.4 Hz), 1.28 (dd,
12H, iPr-CH3, J= 16.6, 7.3 Hz). 31P NMR (DMSO-d6): 133.79. 13C{1H} NMR
(DMSO-d6): 221.01, 215.18, 160.75 (t, J= 7 Hz), 140.85, 99.16 (t, J= 3.1 Hz), 30.43 (t,
J= 13.2 Hz), 17.98, 17.80. IR (KBr, ν CO): br. 1996-1826 cm-1. Rinsing 4 with copious
amounts of ether and removing the volatiles under reduced pressure for extended
periods of time was not sufficient to remove DMF and THF from the compound, as
seen in the 1H and 13C NMR spectra. We attribute the high carbon and hydrogen
numbers observed in the elemental analysis to the presence of these solvents in the
solid samples.

Synthesis of (iPrPNNNCO)Mn(CO)3 (3) and [(iPrPNNNP)Mn(CO)3][Br] (4) in a


closed vessel.
A 50 mL stainless steel vessel was charged with 0.110 g (0.400 mmol) of MnBr(CO)5
and a stir bar in an inert atmoshere glovebox. This vessel was transferred to a
manometer apparatus outside of the glovebox while maintaining inert atmosphere in the
vessel. Under a flow of N2, 0.140 g (0.410 mmol) of iPrPNNNP, dissolved in 12 mL of
fluorobenzene, was syringed into the stainless steel vessel and the reaction was
immediately sealed. The pressure began to rise slowly, at which point heating was
turned on and the reaction was allowed to proceed with stirring for 72 hours at 60 °C.
The pressure rose to roughly 7.5 psi over this period of time. After the reaction was
brought to room temperature, the pressure was vented, the volatiles were removed
under reduced pressure, and the product distribution was assayed via 1H NMR
spectroscopy. The ration of 3:4 skewed to favor the formation of 4 by roughly 10 %
(40:60).

Observation of (iPrPNNNP)Mn(CO)2(Br) (5).


In a 25 mL Schlenk flask, 0.100 g of 4 was dissolved in 5 mL of DMSO-d6. The
Schlenk flask was moved to a Schlenk line and heated to 105 °C while slightly open to
vacuum. The reaction was stirred for four hours, over the course of which the color
went from near colorless to yellow. A 1H NMR spectrum of the reaction mixture was
obtained, showing a clean partial conversion to the desired complex 5. 1H NMR
(DMSO-d6): 8.34 (s, 2H, N-H), 7.23 (t, 1H, p-Py, J= 8.0 Hz), 6.25 (d, 2H, m-Py, J= 8.0
Hz), 3.28 (m, 2H, iPr-CH), 3.28 (m, 2H, iPr-CH), 1.40-1.30 (m, 18H, iPr-CH3), 1.13
(m, 6H, iPr-CH3), (the isopropylmethyl groups of 4 and 5 overlap significantly). 31P
NMR (DMSO-d6): 133.57. 13C{1H} NMR (DMSO-d6): 161.18, 138.30, 97.23, 27.96,
26.52, 18.82, 18.57, 18.05, 17.13.

Synthesis of (iPrPNNNP)Mn(Cl)2 (6).


A 100 mL round bottom flask was charged with 2.0 g (5.858 mmol) of
iPr
PNNNP, 0.735 g of MnCl2 (5.850 mmol) and roughly 30 mL of THF. The reaction
was allowed to stir overnight. Over that time original pink spheres of MnCl2 were
replaced with a flocculent white precipitate. White product was further precipitated
upon the addition of 40 mL of N-hexane. The precipitate was collected on a glass frit,
rinsed with 2 x 30 mL of hexane, and dried under reduced pressure. This yielded 2.35 g
(87 %) grams of (iPrPNNNP)Mn(Cl)2 (6) as an off-white powder. Single crystals
suitable for X-ray diffraction were grown from the mother liquor of 6 by slow
evaporation under reduced pressure. Analysis for C17H33Cl2MnN3P2: Calculated: C =
43.70%, H = 7.12%, N = 8.99%. Found: C = 45.54%, H = 6.97%, N = 8.48%. 1H NMR
(THF-d8): 12.2 (1140 Hz). µeff (Evans, DMSO-d6, 23 °C, µB): 5.77. THF was
consistently present in the 1H NMR spectrum of the complex, indicating strong non-
covalent interactions between the solvent and the complex. We attribute the
unsuccessful elemental analysis results to the presence of THF in the samples.

Observation of (iPrPNNNP)3Mn2 (7).


A 100 mL round bottom flask was charged with 0.500 g (1.070 mmol) of 6 and
35 mL of fluorobenzene. NaHMDS (0.400g, 2.181 mmol) was added portion-wise over
the course of five minutes. The white slurry turned yellow and the solution became
clear as the starting material reacted. After the addition was complete, the reaction was
allowed to stir for an hour, at which time the solution was filtered over Celite and the
volatiles were removed. The residue was taken into fluorobenzene and layered with
hexane. This yielded 0.135 g (12 % based on the amount of ligand present) of a yellow
powder identified as (iPrPNNNP)3Mn2 (7). More product could be gathered by
resubmitting the residue to the same crystallization method. Crystals were grown from
a concentrated ether solution of 7 allowed to slowly evaporate over the course of a
weekend. Complex 7 is NMR silent in either THF-d8 or benzene-d6. The limited
solubility and lack of spectroscopic handles of the material hindered attempts to acquire
adequate characterization of complex 7.

Synthesis of (iPrPONOP)Mn(CO)2(NO2) (8).


A fluorobenzene (5 mL) solution of 0.150 g (0.281 mmol) of
iPr
( PONOP)Mn(CO)2Br (1) was stirred overnight with 0.044 g (0.282 mmol) of silver
nitrite. The light yellow solution was filtered through Celite and cooled to -30 °C
before being layered with 5 mL of N-hexane. After several days pale yellow crystals
had formed that were isolated by decantation and washing with N-hexane before being
placed under reduced pressure to remove volatiles. This gave 0.115 g (81 %) of an off-
white solid identified as (iPrPONOP)Mn(CO)2(NO2). Analysis for C19H31MnN2O6P2:
Calculated: C = 45.61%, H = 6.25%, N = 5.60%. Found: C = 46.78%, H = 6.37%, N =
5.42%. 1H NMR (C6D6): 6.75 (t, 1H, J= 8 Hz), 6.19 (d, 2H, J= 8 Hz), 2.75 (tdd, 2H, J=
13.6, 9.1, 6.4 Hz, iPr-CH), 2.57 (dh, 2H, J= 14.2, 7.2 Hz, iPr-CH), 1.26 (m, 12H, iPr-
CH3), 1.18 (q, 6H, J= 7.4 Hz, iPr-CH3), 1.08 (q, 6H, J= 6.8 Hz, iPr-CH3); 31P NMR
(C6D6): 232.92; 13C{1H} NMR (C6D6): 224.86, 163.58 (t, J= 5.3), 141.67, 102.39,
30.88 (t, J= 6.8 Hz), 28.40 (t, J= 10 Hz), 17.65 (t, J= 4.1 Hz), 17.30 (t, J= 3.5 Hz),
16.97, 15.66. Only one carbonyl resonance was located. IR (KBr, ν CO): 1878 and
1955 cm-1.

Synthesis of (iPrPNHPMn)(CO)2(NO2) (9).


A fluorobenzene (5 mL) solution of 0.150 g (0.302 mmol) of
iPr H
( PN P)Mn(CO)2Br (2) was stirred overnight with 0.047 g (0.303 mmol) of silver
nitrite. The light yellow solution was filtered through Celite and cooled to -30 °C
before being layered with 5 mL of N-hexane. After several days pale yellow crystals
had formed that were isolated by decantation and washing with N-hexane before being
placed under reduced pressure to remove volatiles. This gave 0.095 g (69 %) of an off-
white solid identified as (iPrPNHPMn)(CO)2(NO2). Single crystals suitable for X-ray
diffraction studies were grown in the same manner. Analysis for C18H37MnN2O4P2:
Calculated: C = 46.76%, H = 8.07%, N = 6.06%. Found: C = 44.80%, H = 7.18%, N =
5.68%. 1H NMR (C6D6): 5.55 (br s, 1H, N-H), 2.67 (br s, 2H), 2.22-1.95 (m, 6H), 1.66
(br s, 2H), 1.49 (m, 6H, iPr-CH3), 1.34 (m, 6H, iPr-CH3), 1.20 (m, 6H, iPr-CH3), 1.06
(br s, 2H), 0.92 (m, 6H, iPr-CH3). 31P NMR (C6D6): 86.40. 13C{1H} NMR (C6D6):
229.15, 52.12 (t, J= 5.1 Hz), 27.82 (t, J= 8.8 Hz), 26.05 (t, J= 6.6 Hz), 23.89 (t, J= 9.3
Hz), 20.11, 19.84, 18.66, 17.55. Only one carbonyl resonance was located. IR (KBr, ν
CO): 1844 and 1928 cm-1.

Synthesis of [(iPrPNNN)PMn(CO)3][(NO2)] (10).


A slurry of 0.200 g (0.357 mmol) of [(iPrPNNNP)Mn(CO)3][Br] (4) in 3 mL
DMF was stirred with 0.056 g (0.360 mmol) of silver nitrite overnight. The solution
was filtered through Celite, which was then rinsed with 2 mL of THF. The solution was
layered with 3 mL diethyl ether at room temperature. After the solutions had mixed, a
white powder had formed. The mother liquor was decanted and placed at -30 °C. The
remaining powder was washed with 2 mL of THF and 10 mL of diethyl ether and
volatiles were removed under reduced pressure, leaving 0.140 g (75 %) of a white
powder identified as [iPrPNNNPMn(CO)3][(NO2)]. Analysis for C20H33MnN4O5P2:
Calculated: C = 45.64%, H = 6.32%, N = 10.64%. Found: C = 44.60%, H = 6.03%, N =
9.51%. 1H NMR (DMSO-d6): 9.22 (br. s, 2H, N-H), 7.47 (t, 1H, p-Py, J= 8.0 Hz), 6.37
(d, 2H, m-Py, J= 8.0 Hz), 2.83-2.68 (m, 4H, iPr-CH), 1.40 (dd, 12H, iPr-CH3, J= 15.7,
6.4 Hz), 1.27 (dd, 12H, iPr-CH3, J= 16.6, 7.3 Hz). 31P NMR (DMSO-d6): 133.51.
13
C{1H} NMR (DMSO-d6): 221.01, 215.12, 160.86 (t, J= 7 Hz), 140.88, 99.11 (t, J=
3.1 Hz), 30.48 (t, J= 13.2 Hz), 17.92, 17.80. IR (KBr, ν CO, NO2): br. 1996-1826 cm-1,
1689 cm-1. THF was consistently present in the 13C NMR spectrum of the complex,
even after prolonged exposure to reduced pressure.

Synthesis of [(iPrPONOP)Fe(NO)(CO)][FeCl3NO] (12).


A 20 mL scintillation vial was loaded with 0.100 g (0.201 mmol) of
iPr
PONOPFe(CO)Cl2 (11) and 5 mL of THF. To this stirring slurry was added AgNO2
(0.031 g, 0.202 mmol) as a solid. The reaction immediately turned dark red/brown in
color. The reaction was allowed to stir for one hour, at which time the solution was
filtered over Celite and placed at -30 °C. When the solution was cool, hexane (~ 7 mL)
was carefully layered on top of the THF solution. The layered solution was allowed to
sit undisturbed for one week, after which time dark clumps of crystals had formed on
the bottom of the vial. After decanting the mother liquor and removing volatiles under
reduced pressure, 0.045 g (70 %) of [(iPrPONOP)Fe(NO)(CO)][FeCl3NO] were
recovered. Cutting and separating a fragment from a cluster of the recovered crystalline
solid resulted in a single crystal suitable for X-ray diffraction. Analysis for
C19H31Cl2Fe2N2O5P2: Calculated: C = 33.29%, H = 4.81%, N = 6.33%. Found: C =
33.88%, H = 4.93%, N = 4.05%. 1H NMR (THF-d8): 7.75 (br. s, 1H, 29 Hz), 6.69 (br.
s, 2H, 42 Hz), 2.50 (br. s, 4H, 65 Hz), 1.31 (br. s, 24H, 67 Hz). µeff (Evans, THF-d8, 23
°C, µB): 3.80. IR (KBr, ν NO, CO): 1732, 1780, 1948 cm-1. The low nitrogen content
for the EA was consistent over several attempts. The IR and X-ray confirm one nitrosyl
per iron and only one carbonyl present.

3. Results and Discussion

3.1 Synthesis of PNP-manganese dicarbonyl bromide complexes

Few low-spin synthons for manganese are readily available except for
Mn(CO)5Br. Rather than perform alkali metal reductions of the PNP-supported
manganese dihalides in the presence of CO, these initial studies utilized the
commercially available Mn(I) precursor. Ligation of a meridional PNP chelate will
likely yield a species that maintains the low-spin ground state of the starting manganese
complex. The two initial ligands used in this study, iPrPONOP and iPrPNHP, were both
successfully chelated onto the manganese center with concurrent loss of three
equivalents of CO to yield the complexes (iPrPONOP)Mn(CO)2Br and
(iPrPNHP)Mn(CO)2Br, complexes 1 and 2, respectively (Scheme 1, a and b).

a) b)
Figure 2. The solid-state structures of 1 (a) and 2 (b) are shown at 30 % probability ellipsoids. Hydrogen
atoms, except for N-H, have been removed for clarity.

(iPrPONOP)Mn(CO)2Br, 1, was isolated as an orange powder that exhibits a 1H-


NMR spectrum consistent with a diamagnetic complex. The resonances of the proton
NMR spectrum are also indicative of a top-bottom inequivalent complex, with two sets
of isopropyl resonances. The carbonyl resonances appeared at 228.09 and 225.74 ppm
in the 13C NMR spectrum, and the two intense bands at 1866 and 1946 cm-1 are
assigned to the asymmetric and symmetric CO stretches, respectively. A solid-state
structure was obtained (Figure 2) in order to verify the structure of the complex, as
well as to build a library of ligand metrics with which to compare similar metal
complexes.

Scheme 1. The reactivity of iPrPONOP (a) iPrPNHP (b) and iPrPNNNP (c) with Mn(CO)5Br is summarized.
Number notation and isolated yields are given in bold under the compound.

(iPrPNHP)Mn(CO)2Br (2) was isolated as the syn complex, where the N-H is syn
to the bromide ligand. The 1H NMR spectrum indicated that complex 2 is a diamagnetic
molecule that contains four isopropyl methyl resonances as well as the N-H resonance
at 2.96 ppm. The 13C NMR spectrum of 2 has carbonyl resonances at 232.72 and
229.62 ppm. Infrared spectroscopy in KBr displays two strong bands assigned to the
carbonyl stretches at 1916 and 1824 cm-1. These stretches are lower in energy than
those found in 1 as is expected for the more electron donating (iPrPNHP) ligand.

a) b)
Figure 3. The solid-state structures of 3 (a) and 4 (b) are shown at 30 % probability ellipsoids. Hydrogen
atoms, except for N-H, have been removed for clarity.

Upon reaction of Mn(CO)5Br with the Kirschner-reported chelate iPrPNNNP 6d


in fluorobenzene, two manganese complexes are formed in a roughly 50:50 ratio. 1H
NMR investigations of the reaction mixture show resonances that are consistent with a
left/right inequivalent Cs molecule 3, and a C2v symmetric molecule 4 (Scheme 2). The
similar solubility properties of these complexes in ethers and hydrocarbon solvents
initially complicated the separation of the two compounds. However, extraction of the
mixture of 3 and 4 into DMF completely dissolved 3, leaving 4 as a spectroscopically
pure, off-white powder. The DMF could be removed with mild heating (~ 50 °C) under
vacuum, and the residue washed with THF to afford the purified 3. Carbonyl stretching
frequencies, 13C NMR shifts and selected metrical parameters of 3 and 4 are
summarized for direct comparison (Table 1). Single-crystals suitable for X-ray
diffraction were grown and solid-state structures were obtained to unambiguously
verify the identity of 3 and 4 (Figure 3).

Table 1. Selected properties of the two products, 3 and 4, that result from the reaction of Mn(CO)5Br and
iPr
PNNNP.
Distance (Å), Angles(°) 3 4*
Mn(1)-P(1) 2.266 (1) 2.263 (1)
Mn(1)-C(6) 2.003 (1) -
Mn(1)-N(2) 2.019 (1) 2.059 (4)
P(1)-N(1) - 1.689 (3)
P(1)-N(3) 1.708 (1) -
C(6)-N(1) 1.407 (2) -
P(1)-Mn(1)-C(6) 160.71 (4) -
P(1)-Mn(1)-P(1) - 163.76
N(2)-Mn(1)-C(11) - 177.7
N(2)-Mn(1)-C(15) 177.46 (6) -
C(13)-Mn(1)-C(14) 167.55 (6) -
C(10)-Mn(1)-C(12) - 176.4
δ (N-H) 1H NMR (ppm) 10.10, 8.48 9.09
δ (CO) 13C NMR (ppm) 225.75, 223.74, 217.53 221.01, 215.18,
*The solid-state structure of 4 contains a mirror plane through the center of the molecule, and thus only
has one manganese-phosphorus distance.

Scheme 2. A previously reported formation of an amide bond from a manganese carbonyl complex.15
Complex 3 arises from the formal loss of iPr2PBr and amide formation from the
remaining pyridine “NH” fragment and CO. However, iPr2Br is not observed in the 31P
NMR spectrum of the reaction mixture,19 but several unidentified phosphorus signals
were detected, ranging from approximately -50 ppm to 220 ppm in the 31P NMR
spectrum. Similar transformations, a carbonyl insertion into amides to form the metal-
formamide moiety, have been observed previously in the reaction of dimanganese
nonacarbonyl acetonitrile (Mn2(CO)9MeCN) with thioamides (Scheme 2). 20 This
reaction proceeds in low yields (10% isolated yield with thioacetamide and 7% with
thiobenzamide). Both N-H’s remain in the complex, observed both by 1H NMR
spectroscopy and in the solid-state structure (Table 1). Three separate carbonyl carbon
resonances were observed by 13C NMR spectroscopy for 3, one each for the two CO
ligands and a third for the formed formate moiety. The bond length of P(1)-N(1) of
1.708(1) Å is consistent with other N-P bonds that have the N-H bond intact.

Scheme 3. The product distribution from the attempted decarbonylation of compound 4 is shown.

The other 50% of the yield from the complexation of iPrPNNNP with
Mn(CO)5Br is preserved as the expected [iPrPNNNPMn(CO)3][Br] 4. Unlike the
iPr
PONOP and iPrPNHP supported compounds, 4 is isolated as the salt pair and the
manganese retains three carbonyl ligands. The limited solubility can be attributed to the
extended salt lattice that is observed in the solid-state structure that was obtained from a
DMSO solution of 4. Performing the reaction in a closed, stainless steel manifold
altered the product ratio of 3:4 to 40:60, indicating that select alteration of the reaction
conditions may yield either 3 or 4 preferentially. In this case, retaining the released CO
in the reaction mixture increased the formation of 4. Heating complex 4 in a DMSO
solution under vacuum at 100 °C for 5 hours resulted in the loss of one carbonyl group
and the formation of the bromine-coordinated manganese complex 5 in roughly 60%
yield, with the remaining material persisting as the starting complex (Scheme 3).
Further heating under reduced pressure generated only slightly more pure 5, but always
contaminated with appreciable quantities of 4. This reactivity is reminiscent of a
previously reported inability to fully remove a carbonyl ligand from a PNP manganese
complex.12
The bifurcated product mixture arising from iPrPNNNP and Mn(CO)5Br led us
to explore the chelation of MnCl2 with iPrPNNNP to form a single manganese starting
material. Not surprisingly, the coordination proceeded smoothly and in high yield to
form the off-white flocculent powder of iPrPNNNPMnCl2 (6; Scheme 4). Complex 6
possesses an uninformative 1H NMR spectrum, with a single broad (1140 Hz peak-
width at half-height) resonance located at 12.2 ppm. An Evans method measurement of
6 confirmed the high-spin Mn(II) oxidation state of the metal with a µeff of 5.77 µβ,
close to the expected spin only value of 5 unpaired electrons (5.92 µβ). Other tridentate,
meridional-chelating ligands coordinated to MnCl2 have also been reported to maintain
the high-spin state.21 Single-crystal X-ray diffraction confirmed the structure of 6.

Scheme 4. The reactivity of MnCl2 with PrPNNNP and subsequent dehydrohalogenation reactivity are
summarized.

a) b)
Figure 4. The solid-state structures of 6 (a) and 7 (b) are shown at 30 % probability ellipsoids. Hydrogen
atoms, except for N-H, have been removed for clarity. A THF molecule (for 6) and a diethyl ether
molecule and isopropyl methyl groups (for 7) have also been removed for clarity.

The dehydrohalogenation of 6 was performed to free the manganese


coordination sites of chloride (Scheme 4). In THF or the non-coordinating solvents
toluene or fluorobenzene, the dehydrohalogenation resulted in the formation of a
bimetallic, three ligand compound 7 (Figure 4b). Even when the reaction was
performed in the presence of 10 equivalents of trimethylphosphite, the formation of the
bimetallic complex dominates the reactivity. Both manganese centers adopt a pseudo-
tetrahedral geometry and coordinate to three nitrogen atoms, one nitrogen from the
central ligand and two nitrogen atoms from a second ligand, and a phosphorus atom
from the third ligand. The manganese centers are not within bonding distance at 3.433
Å (the covalent radius of manganese is 1.37 Å). The amine hydrogen (N-H) atoms of
the ligand were found and refined in the solid state structure on N(6) and N(9). The
bond distances are consistent with this assignment as N(6)-C(22) and N(9)-C(39) are
roughly 0.02 Å longer than the amide nitrogen atoms. Further evidence for protonation
of N(6) and N(9) is found in the nitrogen phosphorus bond lengths, with N(6)-P(4) and
N(9)-P(6) being longer than N(1)-P(1) and N(3)-P(3) by an average of 0.025 Å and
N(4)-P(3) and N(7)-P(5) by 0.05 Å. This leaves a total of four anionic charges for the
three ligands, which is consistent with two manganese 2+ centers as can be expected
given the starting material’s spin-state and the tetrahedral environment of each of the
two manganese centers. This also explains the lack of NMR signals in the 1H NMR
spectrum.

Table 2. Select metrical parameters of the solid-state structure of 7 are presented.


Distance (Å), Angles(°) 7
Mn(1)-N(1) 2.016 (2)
Mn(1)-N(4) 2.122 (2)
Mn(1)-N(5) 2.198 (2)
Mn(1)-P(5) 2.538 (1)
Mn(1)-Mn(2) 3.433
Mn(2)-N(3) 2.024 (2)
Mn(2)-N(7) 2.134 (2)
Mn(2)-N(8) 2.241 (2)
Mn(2)-P(3) 2.538 (1)
P(1)-N(1) 1.679 (3)
P(2)-N(3) 1.681 (3)
P(3)-N(4) 1.656 (3)
P(4)-N(6) 1.699 (3)
P(5)-N(7) 1.655 (3)
P(6)-N(9) 1.714 (3)
N(1)-C(1) 1.383 (4)
N(3)-C(5) 1.372 (4)
N(4)-C(18) 1.362 (4)
N(6)-C(22) 1.387 (4)
N(7)-C(35) 1.361 (4)
N(9)-C(39) 1.381 (4)
N(1)-Mn(1)-N(5) 118.19 (9)
N(1)-Mn(1)-N(4) 126.25 (9)
N(1)-Mn(1)-P(5) 114.70 (7)
N(4)-Mn(1)-N(5) 61.93 (9)
N(3)-Mn(2)-N(7) 135.80 (9)
N(3)-Mn(2)-N(8) 115.65 (9)
N(3)-Mn(2)-P(3) 114.78 (7)
N(7)-Mn(2)-N(8) 61.21 (8)

3.2 Reactivity with silver nitrite

The synthesis of manganese nitro compounds was performed with the use of
stoichiometric quantities of silver nitrite in either THF or fluorobenzene (Scheme 5).
The reaction of 1 with silver nitrite proceeded smoothly with concomitant formation of
a grey precipitate, AgBr, to yield (iPrPONOP)Mn(CO)2(NO2), (8). Compound 8 was
isolated as an off-white solid. Upon prolonged sitting at room temperature in an inert
atmosphere, a purple discoloration concurrent with a broadening of the signals in the 1H
NMR spectrum. This decomposition occurred to only a small extent in the solid state
and 8 was easily purified by recrystallization. Spectroscopic characterization of the
purified 8 revealed the coordination geometry as similar to that of 1, with two distinct
isopropyl resonances identified in both 1H and 13C NMR spectra. The carbonyl
stretching frequencies of 1878 and 1955 cm-1 are also similar to the values observed for
1. A solid-state structure confirmed the identity of 8 (Figure 5a).
The same reaction was performed using complex 2. The nitration occurred in a
similar manner and yielded the pale yellow nitro manganese complex
(iPrPNHP)Mn(CO)2(NO2), (9). NMR investigations led to the assignment of 9 as the
nitro-complex analogous to the bromide 2. Complex 9 exhibited four isopropyl methyl
resonances in the 1H NMR spectrum, and the N-H resonance was located at 5.55 ppm,
far downfield compared to the N-H shift of 2 that appears at 2.35 ppm. This shift most
likely arises from the hydrogen bonding of the N-H proton to the nitro group of the
manganese. This is consistent with previously described hydrogen bonding properties
of the N-H of the chelate to a second ligand on the metal; a stronger hydrogen bond
results in a further downfield resonance in the 1H NMR spectrum.7b In this case, the
nitro-group of 9 is forming a stronger hydrogen bond than the bromide of 2 to the N-H
of the chelate, hence the more downfield shift of 5.55 ppm. Obtaining a 13C NMR
spectrum of 9 revealed a single carbonyl resonance at 229.15 ppm and a pattern
associated with the ligand similar to that observed with 2. A solid-state structure of 9
(Figure 5b) reveals a close-contact hydrogen bond between the proton of the chelate
nitrogen and the nitro group on the manganese.

Scheme 5. A summary of the nitration of 1 (a), 2 (b), 4 (c), and 11 (d) to yield 8, 9, 10, and 12,
respectively, is shown.
The addition of silver nitrite to the salt pair 4 yielded the salt pair 10 (Scheme
5c). When the reaction was performed in THF, a grey precipitate formed, but filtration
through Celite yielded negligible quantities of product. When the reaction was carried
out in DMF, 10 could be isolated in good yield. There was no evidence for the
displacement or oxidation of carbonyl by the nitrite moiety. Complex 10 was nearly
identical to 4 by NMR and IR spectroscopies and is most likely the salt pair
[iPrPNNNPMn(CO)3][NO2]. Attempts to grow single crystals were met with mild
success; thin plates were isolated, but the single-crystals were weakly diffracting and
only connectivity was obtained from the experiment (SI), confirming the assignment of
the structure of 10.

a) b)

c) d)

Figure 5. a) The solid-state structure of the nitration product 8, (iPrPONOP)Mn(CO)2(NO2), is shown at


30 % probability ellipsoids. Hydrogen atoms have been omitted for clarity; b) The solid-state structure of
the nitration product 9, (iPrPNHP)Mn(CO)2(NO2), is shown at 30 % probability ellipsoids. Hydrogen
atoms except for the N-H have been omitted for clarity; c) The cationic portion of solid-state structure of
12, [(iPrPONOP)Fe(CO)(NO)]+, is shown at 30 % probability ellipsoids. Hydrogen atoms have been
omitted for clarity; d) The anionic portion of solid-state structure of 12, [FeCl3(NO)]-, is shown at 30 %
probability ellipsoids. Disorder has been omitted for clarity.

Previously reported iron complexes that are iso-electronic with these manganese
complexes have been known for some time. The reactivity of the iron complexes with
silver nitrite would give a direct comparison to the results obtained with the manganese,
and (iPrPONOP)Fe(CO)Cl2,5b 11, was chosen as the iron analogue for nitration. The
reaction of 11 with silver nitrite produced a dark red/brown crystalline solid identified
as [(iPrPONOP)Fe(CO)(NO)]+ using NMR and IR spectroscopies, and single crystal X-
ray diffraction (Figure 5c). The 1H NMR spectrum of 12 showed four broad resonances
attributed to the iPrPONOP ligand. A solid-state structure was collected to provide more
information on the connectivity of the complex. The anion of 12 was revealed to be a
trichloro-iron nitrosyl, [FeCl3(NO)]- (Figure 5d), a moiety first observed by Connelly
in 1976,22 which has been sparsely reported23 up until the most recent report by Lippard
whereby bis(imidazole) iron is used in nitrosyl detection. 24 The carbonyl stretch of
[(iPrPONOP)Fe(CO)(NO)]+ was assigned to the strong band at 1948 cm-1, blue-shifted
compared to the previously reported carbonyl stretches of the neutral complex
(iPrPONOP)Fe(CO)2 reported at 1824 and 1876 cm-1.5b This expected strengthening of
the CO bonds arises from the cationic iron center of 12 having less electron density to
back-donate into the π* orbitals of the carbonyl ligands compared the neutral
(iPrPONOP)Fe(CO)2. The nitrosyl stretch attributed to the [FeCl3(NO)]- of 1790 cm-1 is
similar to other “FeX3NO” (X= Cl, Br; υ NO = 1802 cm-1 for both in CH2Cl2) anions
reported in the literature.15 The nitrosyl stretch attributed to [(iPrPONOP)Fe(CO)(NO)]+
of 1732 cm-1 is comparable to another TBP iron compound containing both a CO and
an NO ligand. The groups of Roustan and Berke have reported complexes of the type
L2(NO)(CO)FeX, where L = monodentate phosphines and X = hydride or chloride. 25
Cationic [L3(NO)(CO)Fe]+ compounds, reported by Johnson, 26 have NO stretches
between 1780 cm-1 and 1697 cm-1 and CO stretches between 1975 cm-1 and 1915 cm-1.
These values are consistent with the values obtained for 12. The µeff of 3.8 µB is also
consistent with reported values for the [FeCl3(NO)]- moiety.22
The oxidation of a variety of organic substrates was attempted with 8 and 9 with
no success. A stoichiometric quantity of the olefins 1-hexene or 1,3-cyclooctadiene was
added to THF solutions containing either 8 or 9, which were subsequently heated to 70
°C. There was no change in the 1H NMR spectrum of either starting material after
heating overnight. Both 9,10-dihydroanthracene and 5 equivalents of phenylsilane with
either 8 or 9 produced no change in the starting nitro-manganese complexes. This series
of experiments confirms the stability of the nitro-moiety in these compounds and their
lack of reactivity under various conditions.

4. Summary/Conclusions
A series of PNP chelates were added to Mn(CO)5Br in an attempt to gain access
to low spin manganese(I) starting materials. Both iPrPONOP and iPrPNHP ligands were
successfully chelated to the manganese center with the concurrent loss of three
equivalents of CO to yield the (PNP)Mn(CO)2Br complexes. Upon reaction of
iPr
PNNNP with Mn(CO)5Br, two complexes were formed that could be isolated in pure
form. A carbonyl ligand from [(iPrPNNNP)Mn(CO)3][Br] could be removed to give the
neutral complex (iPrPNNNP)Mn(CO)2Br, although not in pure form. iPrPNNNP was
capable of yielding three different manganese complexes for further study.
Coordination of MnCl2 with iPrPNNNP gave the high-spin manganese (II) chloride
complex. The product resulting from dehydrohalogenation of 6 is dimanganese,
trisligand complex. Structural varieties similar to 6 remain of interest to us and we
continue to pursue the synthesis reactivity of such systems.
The nitration of the manganese complexes was achieved with the use of silver
nitrite to furnish the nitro-manganese complexes in high yields. Unfortunately, no
reactivity with the coordinated nitro group was observed with a variety of organic
substrates. Although these compounds do not display the desired reactivity, several of
the manganese materials presented here show promise in other chemistry. We continue
to research the reactivity of PNP-supported manganese complexes and will present our
findings in due course.

Acknowledgements
We wish to acknowledge a LANL director’s funded post-doctoral fellowship for AMT
and also acknowledge LANL science campaign 5 for partial support.

Appendix A. Supplementary data

CCDC: 1449437 (Compound 1), 1449435 (Compound 2), 1449436 (Compound


3), 1449430 (Compound 4), 1449432 (Compound 6), 1449429 (Compound 7), 1449434
(Compound 8), 1449431 (Compound 9), and 1449433 (Compound 12) are available
free of charge from http://www.ccdc.cam.ac.uk/conts/retrieving.html, or from the
Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK;
fax: (+44) 1223-336-033; or e-mail: deposit@ccdc.cam.ac.uk. Supplementary
information including 13C NMR data for all diamagnetic compounds, 1H NMR spectra
for select compounds, and tabulated crystal data can be found at
1
a) D. W. Christianson, Prog. Biophys. molec. Biol. 67 (1997) 217; b) K. Wieghardt, Angew. Chem. Int. Ed. Engl.
28 (1989) 1153.
2
a) V. J. Just, C. E. M. Stevenson, L. Bowater, A. Tanner, D. M. Lawson, and S. Bornemann, The Journal of
Biological Chemistry 279(19), (2004), 19867; b) R. Anand, P. C. Dorrestein, C. Kinsland, T. P. Begley, and S. E.
Ealick, Biochemistry 41 (2002) 7659; c) A. Tanner, L. Bowater, S. A. Fairhurst, and S. Bornemann, The Journal of
Biological Chemistry 276(47), (2001), 43627; d) A. Tanner and S. Bornemann, Journal of Bacteriology, 182 (2000)
5271.
3
a) C. Zhang, P. Sricastava, K. Ellis-Guardiola, and J. C. Lewis, Tetrahedron 70 (2014) 4245; b) E. P. Talsi and K.
P. Bryliakov, Coord. Chem. Rev. 256 (2012) 1418; c) E. M. McGarigle and D. G. Gilheany, 105 (2005) 1563; d) S.
Mukhopadhyay, S. K. Mandalm S. Bhaduri, and W. H. Armstrong, Chem. Rev. 104, (2004) 3981.
4
a) Y. Kashiwame, M. Watanabe, K. Araki, S. Kuwata, and T. Ikariya, Bull. Chem. Soc. Jpn. 84 (2011) 251; b) D.
A. Freedman, S. Kruger, C. Roosa, and C. Wymer, Inorg. Chem. 45 (2006) 9558; c) D. S. Bohle, C.-H. Hung, A. K.
Powell, B. D. Smith, and S. Wocadlo, Inorg. Chem. 36 (1997) 268; d) M. A. Andrews, T. C.-T. Chang, and C.-W. F.
Cheng, Organometallics, 4 (1985) 268; e) J. A. Connor and P. I. Riley, J. Chem. Soc., Dalton Trans. (1979) 1231.
5
a) S. Mazza, R. Scopelliti, and X. Hu, Organometallics 34 (2015) 1538; b) W. S. W. DeRieax, A. Wong, and Y.
Schrodi, J. Organomet. Chem. 772-773 (2014) 60.
6
a) M. Glatz, C. Holzhacker, B. Bichler, M. Mastalir, B. Stöger, K. Mereiter, M. Weil, L. F. Veiros, N. C. Mösch-
Zanetti and K. Kirchner, Eur. J. Inorg. Chem. 30 (2015) 5053; b) B. Bichler, C. Holzhacker, B. Stöger, M.
Puchberger, L. F. Veiros, and K. Kirchner, Organometallics 32 (2013) 4114; c) D. Benito-Garagorri, L. G. Alves, L.
F. Veiros, C. M. Standfest-Hauser, S. Tanaka, K. Mereiter, and K. Kirchner, Organometallics 29 (2010) 4932; d) D.
Benito-Garagorri, E. Becker, J. Wiedermann, W. Lackner, M. Pollak, K. Mereiter, J. Kisala, and K. Kirchner,
Organometallics 29 (2006) 1900.
7
a) E. A. Bielinski, M. Förster, Y. Zhang, W. H. Bernskoetter, N. Hazari, and M. C. Holthausen, ACS Catal. 5
(2015) 2404; b) Y. Zhang, A. D. MacIntosh, J. L. Wong, E. A. Bielinski, P. G. Williard, B. Q. Mercado, N. Hazari,
and W. H. Bernskoetter, Chem. Sci. 6 (2015) 4291; c) E. A. Bielinski, P. O. Lagaditis, Y. Zhang, B. Q. Mercado, C.
Würtele, W. H. Bernskoetter, N. Hazari, and S. Schneider, J. Am. Chem. Soc. 136 (2014) 10234; d) I. Koehne, T. J.
Schmeier, E. A. Bielinski, C. J. Pan, P. O. Lagaditis, W. H. Bernskoetter, M. K. Takase, C. Würtele, N. Hazari, and
S. Schneider, Inorg. Chem. 53 (2014) 2133; e) E. Alberico, P. Sponholz, C. Cordes, M. Nielsen, H.-J. Drexler, W.
Baumann, H. Junge, and M. Beller, Angew. Chem. Int. Ed. 52 (2013) 14162.
8
S. Chakraborty, H. Dai, P. Bhattacharya, N. T. Fairweather, M. S. Gibson, J. A. Krause, and H. Guan, J. Am.
Chem. Soc. 136 (2014) 7869.
9
S. Chakraborty, W. W. Brennessel, and W. D. Jones, J. Am. Chem. Soc. 136 (2014) 8564.
10
S. Chakraborty, P. O. Lagaditis, M. Förster, E. A. Bielinski, N. Hazari, M. C. Holthausen, W. D. Jones, and S.
Schneider, ACS Catal. 4 (2014) 3994.
11
C. Bornschein, S. Werkmeister, B. Wendt, H. Jiao, E. Alberico, W. Baumann, H. Junge, K. Junge, and M. Beller,
Nat. Commun. 5 (2014) 4111.
12
a) D. Bacciu, C.-H. Chen, P. Surawatanawong, B. M. Foxman, and O. V. Ozerov, Inorg. Chem. 49 (2010) 5328; b)
A. T. Radosevich, J. G. Melnick, S. A. Stoian, D. Bacciu, C.-H. Chen, B. M. Foxman, O. V. Ozerov, and D. G.
Nocera, Inorg. Chem. 48 (2009) 9214.
13
a) C. Ghosh, T. K. Mukhopadhyay, M. Flores, T. L. Groy, and R. J. Trovitch, Inorg. Chem. 54 (2015) 10398; b)
Mukhopadhyay, M. Flores, T. L. Groy, and R. J. Trovitch, J. Am. Chem. Soc. 136 (2014) 882.
14
H. Salem, L.J.W. Shimon, Y. Diskin-Posner, G. Leitus, Y. Ben-David, and D. Milstein,
Organometallics 28 (2009) 4791.
15
Bruker Instrument Service, Version 2010.1.0.0; Bruker AXS Inc.:Madison, WI, USA, 2010.
16
a) SAINT Plus, Data Reduction Software, Version 7.68A; Bruker AXS Inc.: Madison, WI, USA, 2009; b) G. M.
Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr. A64 (2008) 112−122.
17
a) G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Adv. C71 (2015) 3−8; b) APEX2, Data Refinement
Software, Version 2010.1-2; Bruker AXS Inc.: Madison, WI, USA, 2010.
18
a) Olex2 1.2 (compiled 2014.06.27 svn.r2953 for OlexSys, GUI svn.r4855); b) O. V. Dolomanov, L. J. Bourhis, R.
J. Gildea, J. A. K. Howard, and H. Puschmann, J. Appl. Cryst. 42 (2009) 339-341.
19
A. Hinke and W. Kuchen, Phosphorus and Sulfur and the Related Elements, 15 (1983) 93.
20
a) M. I. Hossain, S. Ghosh, G. Hogarth, and S. E. Kabir, J. Organomet. Chem. 737 (2013) 53; b) D. M. Chipman
and R. A. Jacobson, Inorganica Chimica Acta 1:3 (1967) 393.
21
D. Reardon, G. Aharnian, S. Gambarotta, and G. P. A. Yap, Organometallics 21 (2002) 786; b) D. A. Edwards, M.
F. Mahon, W. R. Martin, K. C. Molloy, J. Chem. Soc., Dalton Trans. 11 (1990) 3161.
22
N. G. Connelly and C. Gardner, J. Chem. Soc., Dalton Trans. (1976) 1525.
23
a) C. T. Tran and E. Kim, Inorg. Chem. 51 (2012) 10086; b) H. Kalyvas and D. Coucouvanis, Polyhedron 26
(2007) 4765; c) S. Brownstein and B. Irish, Polyhedron 7 (1988) 97; d) M. Steimann, U. Nagel, R. Grenz, and W.
Beck, J. Organomet. Chem. 247 (1983) 171.
24
E. Victor, S. Kim, and S. J. Lippard, Inorg. Chem. 53 (2014) 12809.
25
a) J. L. Roustan and A. Forgues, J. Organomet. Chem. 184 (1980) C13; b) J. L. Roustan, J. Y. Merour, and A.
Forgues, J. Organomet. Chem. 184 (1980) C23; c) M. Cygler, F. R. Ahmed, A. Forgues, and J. L. Roustan, Inorg.
Chem. 22 (1983) 1026; d) M. Jänicke, H.-U. Hund, and H. Berke, Chem. Ber. 124 (1991) 719-724.
26
B. F. G. Johnson and J. A. Segal, J. C. S. Dalton 12 (1972) 1268.

Vous aimerez peut-être aussi