Anda di halaman 1dari 7

Polyhedron 29 (2010) 470–476

Contents lists available at ScienceDirect

Polyhedron
journal homepage: www.elsevier.com/locate/poly

Equilibrium and kinetic studies of reactions of [MnN(H2O)(CN)4]2 with


monodentate ligands and the molecular structure of [MnN(NCS)(CN)4]3
Hendrik J. van der Westhuizen a,1, Andreas Roodt b,*, Reinout Meijboom a,*
a
Department of Chemistry, University of Johannesburg, P.O. Box 524, Auckland Park 2006, South Africa
b
Department of Chemistry, University of the Free State, P.O. Box 339, Bloemfontein 9300, South Africa

a r t i c l e i n f o a b s t r a c t

Article history: The dissolution of [(CH3)N]2Na[MnN(CN)5]H2O in water results in the dissociation of the labile trans-CN
Available online 23 June 2009 ligand to form [MnN(H2O)(CN)4]2(aq). The formed aqua complex reacts rapidly with a number of nucle-
ophiles such as NCS and N3. Here we report the single crystal X-ray structure of the reaction products
This article is dedicated to Professor
of as well as the equilibrium and kinetic behaviour of [MnN(H2O)(CN)4]2 in solution.
Stephen S. Basson on the occasion of his
retirement.
Ó 2009 Elsevier Ltd. All rights reserved.

Keywords:
Manganese
Nitrido complex
X-ray structure
Kinetics

1. Introduction With the exception of crystallographic studies of (Ph4P)2


[MnN- (CN)4]2H2O [7], (Ph4P)2[MnN(py)(CN)4]pyH2O [7] and
In-depth studies have shown that the iso-electronic [MX(H2O)- [Rh(en)3]- [MnN(CN)5]H2O [8], hardly any studies of manganese
(CN)4]n complexes (M = Mo(IV), W(IV), Tc(V), Re(V); X = O2, N) nitridotetracyanide complexes were reported. Results from these
are all labile towards monodentate ligand substitution [1–3]. The structure determinations showed that the complexes have very
rates of substitution depend on the pH of the solution, the type short terminal nitrido–metal bonds (1.5–1.6 Å). This indicates
of terminal ligand, the central metal atom and, to lesser extend, the significant p-bond character of the metal–nitrogen bond;
on the type of incoming ligand. the nitrido ligand being the strongest p-donor known to date.
Detailed crystallographic and kinetic studies of the various The Mn„N bond in the five-coordinated [MnN(CN)4]2 anion of
complexes [MO2(CN)4](n+2) at different pH indicated that one of 1.507(2) Å is the shortest nitrido–metal bond reported to date [7].
the O ligands can be protonated to form [MO(H2O)(CN)4]n [4]. With the above in mind we investigated the behaviour of
In contrast to the oxotetracyano complexes, however, a relatively [MnN(CN)5]2 with respect to the anation reaction with CN, NCS,
small amount of nitridotetracyano complexes have been isolated. N3 and OH. Here, we report the synthesis of [MnN(NCS)(CN)4]3,
The protonation behaviour observed for the nitrido complexes 3a, [MnN(OH)(CN)4]3, 3b, [MnN(N3)(CN)4]3, 3c, as well as the
[MN(H2O)(CN)4]n, (M = Re(V) and Os(VI)) [5,6] showed similar equilibrium and kinetic evaluation of the reactions involved.
behaviour to the oxo complexes, with the exception that the parent
[MN(O)(CN)4]n complex was never observed. These kinetic stud- 2. Experimental
ies have shown that the complexes [MN(H2O)(CN)4]n, (M =
Re(V) and Os(VI)) can undergo anation reactions with monoden- 2.1. General
tate ligands like CN and N3 [5]. These kinetic studies have also
shown that the aqua complexes can be deprotonated to form the All chemicals were obtained from Merck or Sigma–Aldrich and
hydroxo complexes and that these are relatively inert towards sub- used as obtained. [(CH3)4N]2Na[MnN(CN)5]H2O, 2a, (kmax = 287,
stitution reactions. 370(sh), 493 nm) and (PPh4)2[MnN(CN)4], 2b, (kmax = 351,
495 nm) were prepared with a slight modification to the published
procedure [7]. The UV–Vis spectra were recorded on either a Var-
ian Cary 50 or Varian Cary 100 spectrophotometer. The IR spectra
* Corresponding authors. Tel.: +27 11 559 2367; fax: +27 11 559 2819.
were recorded on a Bruker Equinox 55 FT-IR spectrometer and ana-
E-mail address: rmeijboom@uj.ac.za (R. Meijboom).
1
Present address: Sasol Technology, P.O. Box 1, Sasolburg 1947, South Africa.
lysed with Bruker Opus-NT software.

0277-5387/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.poly.2009.06.030
H.J. van der Westhuizen et al. / Polyhedron 29 (2010) 470–476 471

2.2. Syntheses of the complexes were performed under pseudo first-order conditions in a carbon-
ate/hydrogen carbonate buffer medium with the ionic strength ad-
2.2.1. [Rh(en)3][MnN(NCS)(CN)4] H2O, 3a H2O justed to l = 1.0 M using KNO3. No reaction between KNO3 and
A solution of 2a (0.06 g, 0.15 mmol) in aqueous 1 M NaNCS [MnN(H2O)(CN)4]2 could be detected. All pH measurements were
(2 cm3) was added to a solution of [Rh(en)3]Cl33H2O (0.120 g, recorded on a Hanna pH211 bench-top pH meter equipped with a
0.27 mmol) dissolved in aqueous 1 M NaNCS (1.5 cm3). This solu- HI1043B refillable combination glass pH electrode using standard
tion was filtered and left at ambient temperature to give red crys- buffer solutions for calibration. The UV/Visible spectral and absor-
tals suitable for X-ray determination after ca. 3 days. Yield: bance changes for the fast kinetic reactions were measured on a Hi
0.0729 g (89%). IR(KBr)/cm1: m(Mn„N) 1056, m(C„N, NCS ) 2068, Tech SF61DX2 Stopped Flow System with a thermostatted
m(C„N) 2126, m(CS) 874. SHU61DX sample handling unit and an attached Jalabu MPV
thermostatted water bath (accurate within ±0.05 °C). The rate con-
2.2.2. [Rh(en)3][MnN(OH)(CN)4] 3b stants were determined using the KinetAsyst 3.10 [9] software
A solution of 2a (0.075 g, 0.19 mmol) in aqueous 1 M NaOH package supplied with the stopped-flow instrument. The UV/Vis
(2 cm3) was added to a solution of [Rh(en)3]Cl33H2O (0.130 g, spectral and absorbance changes for the slow kinetic reactions
0.29 mmol) dissolved in aqueous 1 M NaOH (1.5 cm3). The remain- were measured on Varian Cary 50 Conc and Varian 100 spectro-
ing solution was filtered and left at ambient temperature to give photometers with 1.000 ± 0.001 cm path length tandem quartz
purple crystals of the title compound after ca. 4 days. Yield: cells. All the spectrophotometers were equipped with constant
0.085 g (95%). IR(KBr)/cm1: m(Mn„N) 1002, m(C„N) 2117. temperature cell holders (accurate within 0.1 °C) and Jalabu MPV
thermostatted water baths (accurate within 0.05 °C) fitted with
2.2.3. [Rh(en)3][MnN(N3)(CN)4] 3c circulators and all the temperatures are reported to ±0.1 °C accu-
A solution of 2a (0.06 g, 0.15 mmol) in aqueous 1 M NaN3 racy. Least-squares analyses were performed using MicroMath Sci-
(2 cm3) was added to a solution of [Rh(en)3]Cl33H2O (0.120 g, entist [10] on the absorption versus time data obtained from the
0.27 mmol) dissolved in aqueous 1 M NaN3 (1.5 cm3). The resulting kinetics runs and the spectral changes, respectively.
solution was filtered and left at ambient temperature to give or-
ange crystals after ca. 2 days of the title compound. Yield: 2.4. Crystallography
0.0709 g (92%). IR(KBr)/cm1: m(Mn„N) 1057, m(N„N, N3 ) 2043,
m(C„N) 2125. Intensity data were collected on a Bruker SMART 1K CCD area
detector diffractometer (University of the Witwatersrand) with
2.3. Kinetics graphite monochromated Mo Ka radiation (50 kV, 30 mA) at
273(2) K. The collection method involved x-scans of width 0.5°.
All reagents and chemicals were of analytical grade and double Data reduction was carried out using the program SAINT+ [11] and
distilled water was used in all experiments. All the kinetic runs absorption corrections were made using the program SADABS [11].

Table 1
Crystal data and structure refinement for (PPh4)2[MnN(CN)4] H2O, 2b and [Rh(en)3][MnN(NCS)(CN)4]H2O, 3a.

2b 3a
Empirical formula C52H44N5O2P2Mn C11H26N12OSMnRh
Formula weight 887.80 532.25
T (K) 273.0(2) 273.0(2)
Wavelength (Å) 0.71073 0.71073
Crystal system triclinic orthorhombic
Space group P1 P212121
Unit cell dimensions
a (Å) 12.103(2) 8.749(2)
b (Å) 12.337(2) 12.904(3)
c (Å) 16.781(2) 17.854(4)
a (°) 97.511(3) 90
b (°) 108.993(3) 90
c (°) 90.560(3) 90
V (Å3) 2345.1(6) 2015.8(8)
Z 2 4
Dexp (Mg m3) 1.24 1.76
Dcalc (Mg m3) 1.257 1.754
Absorption coefficient (mm1) 0.395 1.578
F(0 0 0) 924 1080
Crystal size (mm3) 0.44  0.28  0.26 0.50  0.24  0.20
h Range for data collection (°) 1.67–26.00 1.95–28.31
Index ranges 14 6 h 6 14, 9 6 k 6 15, 20 6 l 6 20 11 6 h 6 11, 17 6 k 6 14, 23 6 l 6 22
Reflections collected/unique/observed 13 960/9125/5979 12 757/4904/4667
Independent reflections [Rint] [0.0274] [0.0214]
Completeness to 2h (°, %) 26.00, 99.0 28.31, 99.1
Absorption correction semi-empirical from equivalents semi-empirical from equivalents
Maximum and minimum transmission 0.902, 0.876 0.729, 0.641
Refinement method full-matrix least-squares on F2 full-matrix least-squares on F2
Data/restraints/parameters 9125/0/571 4904/0/250
Goodness-of-fit (GOF) on F2 1.037 1.148
Final R indices [I > 2r(I)] R1 = 0.0456, wR2 = 0.0963 R1 = 0.0283, wR2 = 0.0611
R indices (all data) R1 = 0.0849, wR2 = 0.1094 R1 = 0.0312, wR2 = 0.0619
Largest difference in peak and hole (e Å3) 0.261 and 0.286 0.941 and 0.744
472 H.J. van der Westhuizen et al. / Polyhedron 29 (2010) 470–476

The structures were solved by the Patterson method and refine- quencies for the N3, NCS and the OH ligands are 1057, 1056
ments through full-matrix least-squares cycles using the SHELXL-97 and 1002 cm-1, respectively. These m(Mn„N) values indicate that
P
[12] software package with (/Fo//Fc/)2 being minimized. Non- the trans-influences (and probably the bond strengths) decrease
hydrogen atoms were first refined isotropically, followed by aniso- in order of H2O (or no ligand)  N3 > NCS > py [7] > CN > OH.
tropic refinement. The hydrogen atoms of the water molecules of The m(CN) stretching frequencies of all reported compounds are
crystallization were determined geometrically from difference around 2125 cm1, except for the OH complex 3b which shows
Fourier maps. The program DIAMOND [13] was used to produce a considerable shift to 2117 cm1 for the m(CN). The shift of the
molecular diagrams of each of the respective complexes. The den- m(CN) in the hydroxide complex 3b is a clear sign of the inert nat-
sities of the crystals were determined by flotation in benzene/ ure of the hydroxide complexes (see Section 3.3).
diiodomethane. Crystal data and details of refinement are given
in Table 1. 3.2. Molecular structure

3. Results and discussion Since the structures of the aqua complexes of the various iso-
electronic metal tetracyanides are known, we attempted to crystal-
3.1. Synthesis lise the elusive [MnN(H2O)(CN)4]2 anion. Unfortunately the ques-
tion of the existence of the [MnN(H2O)(CN)4]2 in solution as well
It is somewhat surprising that the nitridotetracyano complexes as in the solid state remains unanswered, since the crystals ob-
of Mn(V) [7,8] have received little attention, whereas the corre- tained from aqueous solutions after numerous careful crystalliza-
sponding analogues of Tc(V), Re(V) and Os(VI) are known. Thus, tions contained only the [MnN(CN)4]2 anion. Here, we report
[TcN(OH2)(CN)4]2 [14], [ReN(OH2)(CN)4]2 [15], [ReN(CN)5]3 the redetermination of this anion (figure not shown). For the
[16], [ReN(N3)(CN)4]3 [17] and [OsN(OH)(CN)4]2 [18] have been description of this structure we refer to the original publication [7].
synthesized and the PPh4+ or AsPh4+ salts of these anions have been A molecular diagram showing the numbering scheme of the ti-
crystallographically characterized. tle compound [Rh(en)3][MnN(NCS)(CN)4] H2O, 3a, is presented in
Fairly simple synthetic routes are available for the synthesis of Fig. 1, with selected bond lengths and angles in Table 2 [NB: extend
the complexes [MN(CN)4]n (M = Tc, Re, Os). It seems, however, Table 2 to include much more bond distances, angles and torsion
that only a limited number of synthetic routes are available for angles]. Comparative bond distances and angles for selected re-
the synthesis of the nitridotetracyano complexes of manganese(V) lated compounds are given in Table 3. Compound 3a crystallizes
[19]. We followed the route reported a decade ago which synthe- in the orthorhombic space group, P212121, with Z = 4. The chiral
sized of the nitridotetracyano complexes of Mn(V) by ligand sub- center of the molecule is situated on the cation and the conforma-
stitution of the salen ligand of [MnVN(salen)], 1, by CN in tion of the cation was determined as Kkkd [20]. The rhodium metal
aqueous solutions of NaCN [7] (Scheme 1). center of the trivalent cation is octahedrally coordinated to the six
The IR stretching frequencies of the products show a behaviour nitrogen atoms of three 1,2-diaminoethane ligands. The average
similar to that of the corresponding Re complexes. The m(Mn„N)
stretching frequency for the parent complex [MnN(CN)4]2 is
1105 cm1 (1094 reported for solution [7]). This is significantly dif-
ferent from the value of 1011 cm1 obtained for the [MnN(CN)5]3
anion [8]. The m(Mn„N) stretching frequency is sensitive to the
trans-influence of entering ligands. The m(Mn„N) stretching fre-

Fig. 1. Molecular diagram showing the numbering scheme and displacement


ellipsoids (30% probability) of the [MnN(NCS)(CN)4]3 anion. The cations and
solvent molecules are omitted for clarity.

Table 2
Selected interatomic bond distances (Å) and angles (°) for 3a. Expand table more with
bond distance, angles and torsion angles.

Distances
Mn–N(10) 1.540(3) Mn–N(5) 2.298(3)
N(5)–C(5) 1.154(5) C(5)–S 1.664(4)
Angles
N(10)–Mn–N(5) 174.60(13) N(5)–C(5)–S 177.6(3)
Scheme 1. Synthesis of compounds.
H.J. van der Westhuizen et al. / Polyhedron 29 (2010) 470–476 473

Table 3
Comparison of structural parameters (Å, °) in related [MNX(CN)4]n (M = Mn, Tc, Re, Os; X = NCS, CN) anions.

Anion M„N (Å) M–Ca (Å) M–Lb (Å) Dc (Å) m(M„N) (cm1) Ref.
3
[MnN(NCS)(CN)4] 1.540(3) 2.006(3) 2.298(3) 0.196(2) 1056 TW
[MnN(CN)4]2 1.497(2) 1.970(3) 0.435(1) 1105 TW
[MnN(CN)4]2 1.507(2) 1.983(2) 0.436 1094 [7]
[MnN(py)(CN)4]2 1.525(4) 1.997(4) 2.472(4) 0.246(2) 1041 [7]
[MnN(CN)5]3 1.499(8) 1.990(6) 2.243(7) 0.222(1) 1011 [8]
[TcN(H2O)(CN)4]2 1.596(10) 2.11(2) 2.559(9) 0.35 1100 [14]
[ReN(CN)5]3 1.68(1) 2.12(1) 2.39(1) 0.31 1040 [16]
[ReN(H2O)(CN)4]2 1.64(1) 2.11(1) 2.496(7) 0.35 1060 [15]
[ReN(N3)(CN)4]3 1.65(2) 2.11(1) 2.36(2) 0.34 1074 [17]
[OsN(OH)(CN)4]2 1.606(5) 2.068(9) 2.123(5) 0.26 1050 [18]

Refer also here to Refs. [2,3].


TW = this work.
a
Equatorial M–CN bonds.
b
trans to M„N.
c
Displacement of metal from the best plane formed by the four CN ligands.

Rh–N bond distance of 2.087(3) Å corresponds well with the same of the nitrogen atoms of an equatorial cyano ligand. The hydrogen-
bond distances found in the [Rh(en)3]3+ cation of the [Rh(en)3] bonding between a water molecule of crystallization and an
[MnN(CN)5]H2O complex [8] [average Rh–N = 2.079(5) Å]. The equatorial cyano ligand seems to be a frequently occurring phe-
N–Rh–N bond angles range from 82.9(1)° to 94.9(1)° and are nomenon for this type of transition metal complex. The water of
comparable to those found for the [Rh(en)3]3+ cation of the crystallization links two anions in a zig-zagging pattern, with the
[Rh(en)3][MnN(CN)5]H2O complex [8]. cations suspended between three anions forming one bend (see
The Mn(V) metal center is octahedrally coordinated to the nitr- Fig. 2).
ido ligand, four cyano ligands and the thiocyanate ligand trans to Interestingly, the two water molecules of crystallization in the
the nitrido ligand. The Mn„N bond distance of 1.540(3) Å is longer [MnN(CN)4]2 anion, 2b, are not located near the vicinity of the va-
than the corresponding bond distance of 1.499(8) Å in the cant sixth coordination site of [MnN(CN)4]2, but instead they form
[MnN(CN)5]3 complex [8]. The average Mn–C bond distance be- weak intermolecular hydrogen-bonding contacts with two nitro-
tween the central metal atom and the carbon atoms of the cyano gen atoms [O1 WN2 = 2.984(3) Å, O2 WN4 = 3.054(4) Å] of
ligands of 2.006(3) Å and the average C„N bond distance of two of the cyano ligands. Here, the anions and the water molecules
1.156(4) Å are comparable to the same bond distances found in form an infinite chain structure in the solid state. In contrast to the
[MnN(CN)4]2 of 1.983(2) and 1.160(2) Å, respectively [7]. The structure of 3a, the nitrido ligand is not involved in any hydrogen-
average Mn–C–N bond angle is 175.5(3)° and is nearly linear as ex- bonding.
pected for this type of moiety. The Mn–N bond distance of A root-mean-square (r.m.s.) calculation is one way to compare
2.298(3) Å between the Mn(V) and the thiocyanate ligand is similar structures. The calculated r.m.s. deviations between the an-
slightly longer than the Mn–C bond distance of 2.243(7) Å between ions [MnN(CN)4]2, 2b, and [MnN(NCS)(CN)4]3, 3a, is calculated to
the Mn(V) and the cyano ligand coordinated trans to the nitrido in be 0.226 Å. This small r.m.s. errors confirms that the general orien-
the [MnN(CN)5]3 anion [8]. The manganese(V) metal center can tation is similar in the two anions. In addition, the differences be-
be considered as a hard acid due to the high formal charge on tween the two compounds can be seen from the overlay of these
the metal atom and will preferably bind to the hard base part of two structures (Fig. 3). It is clear from Fig. 3 that the main differ-
the thiocyanate ion, i.e. the nitrogen atom. The N(5)–C(5) bond dis- ences between the anions is the change from the central Mn atom
tance is 1.154(5) Å and the C(5)–S bond distance is 1.664(4) Å. The out of the plane formed by the four CN ligands, as indicated by the
Mn–N(5)–C(5) angle is 158.3(3)° and the N(5)–C(5)–S angle is larger N–Mn–CN angles in the [MnN(CN)4]2 anion.
177.6(3)°. Similar results were reported for the [ReO(NCS)(CN)4]2
complex [21].
The Mn(V) atom is displaced by 0.196(2) Å from the best plane 3.3. Kinetics
formed by the four carbon atoms of the cyano ligands. This large
distortion of the coordination polyhedron is also reflected in the Results from previous kinetic studies of the reactions of
deviations from the expected rectangular values for the N(10)– [MN(H2O)(CN)4]n [M = Re(V), Os(VI)] with monodentate ligands
Mn–C and C–Mn–N(5) bond angles which range from 99.26(15)° have shown that only the aqua ligand is substituted in these com-
to 93.68(14)° and 88.31(13)° to 81.52(13)°, respectively. The dis- plexes [5]. These studies have also shown that the hydroxo com-
placement of the Mn atom towards the nitrido ligand along the plex is formed upon deprotonation of the aqua complex and that
N(10)„Mn–N(5) axis forces the cis-bonded cyano ligands and the hydroxo complex is inert towards monodentate substitution
the N(5) atom of the NCS ligand closer together as indicated by reactions. A dissociative mechanism for the substitution of the
smaller than 90° C–Mn–N(5) bond angles. A similar displacement aqua ligand in [MN(H2O)(CN)4]n by monodentate nucleophiles
was found for the [MnN(CN)5]3 anion [8]. was proposed.
The [Rh(en)3]3+ cation forms three weak intermolecular hydro- Previous research has shown that the [MN(H2O)(CN)4]n aqua
gen-bonding contacts [N(12)N(4) = 3.017(4) Å, N(31)N(3) = complexes [M = Re(V); Os(VI)] are the only complexes reactive to-
3.309(4) Å and N(32)N(10) = 3.223(5) Å] with two nitrogen wards substitution reactions and that the hydroxo ligand in
atoms of the equatorial cyano ligands and the axial nitrido ligand. [MN(OH)(CN)](n+1) is not substituted by entering nucleophiles,
Normally the nitrido ligand does not form any intermolecular con- i.e., the hydroxo complex is inert towards substitution reactions
tacts, except in the case when potassium was the counter ion in the [5]. These observations enabled the construction of the reaction
[ReN(CN)4]2 complex. In the latter case infinite chains of Re- scheme as represented in Scheme 2 for the substitution reactions
NRe„NRe were observed. The water molecule of crystallization with monodentate nucleophiles as well as the deprotonation reac-
forms a weak hydrogen-bond [O(1 W)N(2) = 2.893(4) Å] with one tions of the [M(H2O)(CN)4]n complexes in solution.
474 H.J. van der Westhuizen et al. / Polyhedron 29 (2010) 470–476

Scheme 2. The dissociation and solution behaviour of the [MnN(CN)5]3 complex


in aqueous medium.

For the nucleophilic reactions of the Re(V) and Os(VI) analogues,


the aqua compound [MN(H2O)(CN)4]n could be used [5,6]. How-
ever, for the Mn(V) analogues, the pentacyano complex
[(CH3)N]2Na[MnN(CN)5]H2O was used. The trans-bonded cyano li-
gand dissociates in solution (even at [CN] 0.3 M) to form the
aqua complex (reaction A, Scheme 2) in situ, and this aqua complex
reacts with monodentate, or even bidentate, nucleophiles.
Due to the high concentration of the aqua ligand in solution (ca.
55 M) it is presumed that the [MnN(H2O)(CN)4]2 anion exists in
solution after the dissociation of the trans-bonded ligand (CN)
or that the aqua ligand is at least in the very near vicinity of the va-
cant coordination site. Unfortunately, the molecular structure of
the [MnN(H2O)(CN)4]2 complex has not been determined as of
yet, and thus, the crucial question of whether the aqua ligand is
bonded trans to the nitrido ligand remains unanswered for now,
although it is assumed to be.
Unfortunately, the corresponding acid dissociation constant
(pKa1) of the aqua complex could also not be determined due to
the relatively small difference in the UV–Vis absorbance spectra
of the aqua and hydroxo complexes (see reaction B, Scheme 2).
Fig. 2. Hydrogen-bonding pattern for 3a.
However, results from the kinetic studies of the monodentate sub-
stitution reactions of the [MN(H2O)(CN)4]n complexes [M = Re(V),
Os(VI)] have shown that the hydroxo complexes do not react with
monodentate nucleophiles [5]. Therefore, the value of the forward
reaction between the [MN(OH)(CN)4]3 and monodentate ligands
(k2) is assumed to be zero within experimental error (reaction D,
Scheme 2). The same tendency is expected for the substitution
reactions of the [MnN(OH)(CN)4]3 with monodentate
nucleophiles.
The UV–Vis spectrum of the reaction between
[MnN(H2O)(CN)4]2 and added CN ions showed only one maxi-
mum in the spectrum at k = 285 nm. In the spectrophotometric
determination of the stability constant of the [MnN(CN)5]3 com-
plex the concentration of the CN was varied from 0.005 to
0.10 M. The plot of absorbance versus concentration of the entering
ligand (Fig. 2) shows that under these conditions [MnN(H2O)-
(CN)4]2 reacts with CN in a 1:1 ratio which is in agreement with
the formulation of the product as [MnN(CN)5]3 [7].
The equilibrium constant for the reaction between
[MnN(H2O)(CN)4]2 and CN was determined by a least-squares
fit of the absorbance versus [CN] to Eq. (1) and is illustrated in
Fig. 4. This equation is derived using Beer0 s law, mass balance as
well as the definition of the stability constant, K1, for the reaction
A in Scheme 2. The stability constant for the [MnN(CN)5]3 com-
Fig. 3. Overlay of [MnN(CN)4]2, 2b (hashed) and [MnN(NCS)(CN)4]3, 3a (solid). plex was determined as 168(20) M1.
H.J. van der Westhuizen et al. / Polyhedron 29 (2010) 470–476 475

second- to third-row element suggests that the metal–aqua bond is


stronger in the third-row elements. Conversely, the aqua substitu-
tion reactions for the first-row elements are expected to be even
faster in a dissociative mechanism, i.e., when the rate is dependent
on the dissociation of the aqua ligand. These results also indicate
that the rate of the reverse reaction (k1) or the dissociation of
the CN ligand has a significant contribution in the overall process.
The very large difference found for the values of k1 for the com-
plexes of the second- and third-row and possibly the first-row ele-
ments may also point towards a dissociative mechanism for the
substitution reactions. A weaker metal–aqua bond for the second
and possibly the first-row elements is also evident from the higher
pKa values for the second- and third-row complexes. A weaker me-
tal–oxygen bond will strengthen the oxygen–hydrogen-bonds in
the aqua ligand with a resulting higher pKa value.
Fig. 4. Plot of absorbance vs. [CN] at 25.2 °C, [MnN]T = 5.05  104 M, l = 1.0 M
Upon comparing the values of the k1 in Table 4 for the reactions
(NaCl), pH 11.01 and k = 285 nm.
of the oxo and nitrido complexes of Re(V) an even more spectacu-
lar effect of bond strength on the relative reactivity of these com-
AM þ AML K 1 ½CN 
Aobs ¼ ð1Þ plexes is observed. The nitrido complex with the weak metal–aqua
1 þ K 1 ½CN  bond [Re–OH2 = 2.496(7) Å] [15] reacts about 106 times faster than
the oxo with the much stronger metal–aqua bond [Re–
In the above equation, AM and AML represents the absorbance of OH2 = 2.142(7) Å] [28]. The effect of the Re–aqua bond strengths
the [MnN(H2O)(CN)4]2 and [MnN(CN)5]3 complexes, Aobs, the ob- in the oxo and nitrido complexes is also evidenced by the very
served absorbance and K1 the equilibrium constant of the reaction, large increase in the pKa value of 1.4 for the oxo complex to about
respectively. 12 for the nitrido complex. This very high pKa value for the nitrido
The forward reactions of the [MnN(H2O)(CN)4]2 with mono- complex points to a very weakly bonded aqua ligand and this is in
dentate ligands (CN, NCS and N3) were extremely rapid. Unfor- agreement with the two above-mentioned Re–aqua bond distances
tunately, these reaction rates of the substitution reactions could found in the crystal structure determinations of the oxo and nitrido
not be determined, even on our third generation stopped-flow complexes. The large dependency of the rate of substitution on the
spectrophotometer. This spectrophotometer has a mixing time of Re–aqua distance, as well as the high reactivity of the nitrido com-
less than 1 ms before the measurement of the absorbance versus plex also suggests a dissociative mechanism.
time data starts. When the standard equation for the calculation Thus it is expected that the [MnN(H2O)(CN)4]2 complex will
of the half-life of pseudo first-order reactions is used, a value larger have a higher pKa value compared to the [ReN(H2O)(CN)4]2 com-
than 1000 s1 is obtained for the observed rate constant, kobs, at plex, due to the fact that upon going from the third to the second-
[CN] = 0.05 M. If the forward rate constant is calculated using row and then the first-row elements the metal–aqua bond is ex-
the minimum value of the observed rate constant and the concen- pected to weaken significantly. The [MnN(H2O)(CN)4]2 complex
tration of the entering ligand, a value of k1 > 20 000 M1 s1 is ob- should also have a much higher reactivity compared to the [Re-
tained, which indicates that the substitution reactions of the N(H2O)(CN)]2 complex and the substitution reactions of the aqua
[MnN(H2O)(CN)4]2 complex are extremely rapid. When compar- ligand in [MnN(H2O)(CN)4]2 is expected to proceed via a dissocia-
ing the stability constants for the known monodentate substituted tive mechanism.
[MO(L)(CN)4]n complexes (for M = Mo(IV), L = CN, N3, NCS:
95(5), 1.5(1), 0.2(1) M1 and M = W(IV), L = CN, N3, NCS: 4. Conclusions
1.1(1)  103, 4.8(11), 2.0(1) M1) it can be predicted that the stabil-
ity constants (reaction C, Scheme 1) for the [MnN(NCS)(CN)4]3 The dissolution of [(CH3)N]2Na[MnN(CN)5]H2O in water re-
and [MnN(N3)(CN)4]3 complexes will be smaller than the stability sulted in the dissociation of the labile trans-CN ligand to form
constant for [MnN(CN)5]3 [K1 = 168(20) M1]. [MnN(H2O)(CN)4]2(aq). The formed aqua complex was reacted
Comparable kinetic studies of the reactions between with X = OH, NCS and N3 to form the [MnN(X)(CN)4]3 com-
[MX(H2O)(CN)4]n (M = Mo(IV) [22], W(IV) [23–25], Tc(V) [26], plexes. Despite several attempts the parent [MnN(H2O)(CN)4]2
Re(V) [27] and Os(VI); X = O, N) and different monodentate ligands anion could not be crystallized, and a redetermination of the
are shown in Table 4. The increase in reaction rate upon going from [MnN(CN)4]2 anion resulted. The single crystal X-ray structure

Table 4
Kinetic data for the reaction of [MX(H2O)(CN)4]n complexes with monodentate ligands (L) at 25 °C.

Anion L pKa1 k1 (M1 s1) K1 (M1) DS– (J K1 mol1) Ref.


[MoO(H2O)(CN)4] CNa 10.0 116(2) 95(5) 53(14) [22]
[WO(H2O)(CN)4] CN 7.8 1.15(3) 1.1(1)  103 12(20) [23]
NCS 2.9(1) 2.0(1) 8(9) [24]
N3c 4.2(1) 4.8(11) 10(8) [25]
[TcO(H2O)(CN)4]3 NCS 2.9 22.2(3) 54(2) 9(12) [26]
[ReO(H2O)(CN)4] NCS 1.4 3.48(4)  103 87(7) 46(20) [27]
[MnN(H2O)(CN)4]2 CN 13 >20 000 168(20) TW
[ReN(H2O)(CN)4]2 CNb 11.7 7.2(4)  103 600(100) 40(6) [5a]
[OsN(H2O)(CN)4] N3 7.43 1.89(7) 189(8) 45(3) [5b]
a
At 20 °C.
b
At 15 °C.
c
DV– = 10.6(5) cm3 mol1.
476 H.J. van der Westhuizen et al. / Polyhedron 29 (2010) 470–476

of [MnN(NCS)(CN)4]3 shows an octahedral coordination around [6] T.N. Mtshali, W. Purcell, H.G. Visser, S.S. Basson, Polyhedron 25 (2006) 2415.
[7] J. Bendix, K. Meyer, T. Weyhermüller, E. Bill, N. Metzler-Nolte, K. Wieghardt,
the Mn atom, with the nitrido ligand trans to the thiocyanate. Ini-
Inorg. Chem. 37 (1998) 1767.
tial kinetic investigations indicated that the reaction of [8] J. Bendix, R.J. Deeth, T. Weyhermüller, E. Bill, K. Wieghardt, Inorg. Chem. 39
[MnN(H2O)- (CN)4]2(aq) with NCS and N3 were too fast to ob- (2000) 930.
tain suitable data. The kinetic investigation with CN revealed a [9] KinetAsyst 3.10 Software Package, Hi-Tech Ltd., Salisbury, UK, 2002.
[10] MicroMath Scientist Software for Windows, Version 2.01, Micromath Inc.,
stability constant of 168(20) M1 for [MnN(CN)5]3. 1995.
[11] Bruker SAINT+. Version 6.02 Includes XPREP and SADABS, Bruker AXS Inc., Madison,
Acknowledgements WI, USA, 1999.
[12] G.M. Sheldrick, SHELX-97, Program for the Refinement of Crystal Structures,
University of Göttingen, Germany, 1997.
Financial assistance from the Research Funds of the Universities [13] K. Brandenburg, DIAMOND (Ver. 2.1e), Crystal Impact GbR, Bonn, Germany, 2001.
of Johannesburg and the Free State, SASOL, THRIP, and the NRF is [14] J. Baldas, J.F. Boas, S.F. Colmanet, M.F. Mackay, Inorg. Chim. Acta 170 (1990)
233.
gratefully acknowledged. Part of this material is based on work [15] W. Purcell, I.M. Potgieter, L.J. Damoense, J.G. Leipoldt, Transition Met. Chem. 17
supported by the South African National Research Foundation. (1992) 387.
Opinions, findings, conclusions or recommendations expressed in [16] W. Purcell, I.M. Potgieter, L.J. Damoense, J.G. Leipoldt, Transition Met. Chem. 16
(1991) 473.
this material are those of the author and do not necessarily reflect [17] W. Purcell, L.J. Damoense, J.G. Leipoldt, Inorg. Chim. Acta 195 (1992) 217.
the views of the NRF. [18] H.J. van der Westhuizen, S.S. Basson, W. Purcell, Transition Met. Chem. 19
(1994) 582.
[19] (a) S.I. Arshankov, A.L. Poznjak, Z. Anorg. Allg. Chem. 481 (1981) 201;
Appendix A. Supplementary data (b) J.W. Buchler, C. Dreher, K.-L. Lay, Z. Naturforsch., B: Anorg. Chem. Org.
Chem. 37B (1982) 1155;
CCDC 730581 and 730582 contain the supplementary crystallo- (c) C.L. Hill, F.J. Hollander, J. Am. Chem. Soc. 104 (1982) 7318;
(d) J.W. Buchler, C. Dreher, K.-L. Lay, Y.J.A. Lee, W.R. Scheidt, Inorg. Chem. 22
graphic data for 2b and 3a. These data can be obtained free of (1983) 888;
charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html, or (e) J. Du Bois, C.S. Tomooka, J. Hong, E.M. Carreira, M.W. Day, Angew. Chem.,
from the Cambridge Crystallographic Data Centre, 12 Union Road, Int. Ed. Engl. 36 (1997) 1645;
(f) A. Niemann, U. Bossek, G. Haselhorst, K. Wieghardt, B. Nuber, Inorg. Chem.
Cambridge CB2 1EZ, UK; fax: (+44) 1223-336-033; or e-mail: de- 35 (1996) 906;
posit@ccdc.cam.ac.uk. Supplementary data associated with this (g) J. Du Bois, C.S. Tomooka, J. Hong, E.M. Carreira, Acc. Chem. Res. 30 (1997)
article can be found, in the online version, at doi:10.1016/ 364.
[20] K.F. Purcell, J.C. Kotz, Inorganic Chemistry, W.B. Sanders Company, 1977.
j.poly.2009.06.030.
[21] W. Purcell, A. Roodt, S.S. Basson, J.G. Leipoldt, Transition Met. Chem. 14 (1989)
369.
References [22] (a) P.R. Robinson, E.O. Schlemper, R.K. Murmann, Inorg. Chem. 14 (1975)
2035;
[1] J.G. Leipoldt, S.S. Basson, A. Roodt, Adv. Inorg. Chem. 40 (1993) 297. (b) J.P. Smit, W. Purcell, A. Roodt, J.G. Leipoldt, Polyhedron 12 (1993) 2271.
[2] A. Roodt, A. Abou-Hamdan, H.P. Engelbrecht, A.E. Merbach, in: A.G. Sykes, (Ed), [23] (a) E. Heijmo, A. Kanas, A. Samotus, Bull. Acad. Polon. Sci. Ser. Sci. Chim. 21
Advances in Inorganic Chemistry, vol. 49, 1999; published 2000, p. 59. (1973) 311;
[3] J.M. Botha, A. Roodt, Metal-Based Drugs, vol. 2008, Article ID 745989, 2008, 9 (b) J.P. Smit, Ph.D. Thesis, University of the Free State, Bloemfontein, South
pages, doi:10.1155/2008/745989. Africa, 1995.
[4] (a) V.W. Day, J.L. Hoard, J. Am. Chem. Soc. 90 (1968) 3374; [24] A. Roodt, J.G. Leipoldt, S.S. Basson, I.M. Potgieter, Transition Met. Chem. 13
(b) S.J. Lippard, B.J. Russ, Inorg. Chem. 6 (1967) 1943; (1988) 336.
(c) K. Weighardt, G. Backes-Dahmann, W.J. Holzbach, W.J. Swiridoff, J. Weiss, [25] J.G. Leipoldt, R. van Eldik, S.S. Basson, A. Roodt, Inorg. Chem. 25 (1986) 4639.
Z. Anorg. Allg. Chem. 499 (1983) 44. [26] A. Roodt, J.G. Leipoldt, E.A. Deutsch, J.C. Sullivan, Inorg. Chem. 31 (1992) 1080.
[5] (a) L.J. Damoense, W. Purcell, J.G. Leipoldt, Transition Met. Chem. 19 (1994) [27] W. Purcell, A. Roodt, S.S. Basson, J.G. Leipoldt, Transition Met. Chem. 14 (1989)
619; 224.
(b) H.J. van der Westhuizen, S.S. Basson, J.G. Leipoldt, W. Purcell, Polyhedron [28] W. Purcell, A. Roodt, S.S. Basson, J.G. Leipoldt, Transition Met. Chem. 15 (1990)
13 (1994) 717. 239.

Anda mungkin juga menyukai