Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Jun 13.
Published in final edited form as: Inorg Chem. 2014 Jan 8;53(2):1144–1155. doi: 10.1021/ic4026987

A New Series of Complexes Possessing Rare “Tertiary” Sulfonamide Nitrogen-to-Metal Bonds of Normal Length: fac-[Re(CO)3(N(SO2R)dien)]PF6 Complexes with Hydrophilic Sulfonamide Ligands

Pramuditha L Abhayawardhana 1, Patricia A Marzilli 1, Frank R Fronczek 1, Luigi G Marzilli 1,
PMCID: PMC4465231  NIHMSID: NIHMS554741  PMID: 24400928

Abstract

Tertiary sulfonamide nitrogen-to-metal bonds of normal length are very rare. We recently discovered such a bond in one class of fac-[Re(CO)3(N(SO2R)(CH2Z)2)]n complexes (Z = 2-pyridyl) with N(SO2R)dpa ligands derived from di-(2-picolyl)amine (N(H)dpa). fac-[M(CO)3(N(SO2R)(CH2Z)2)]n agents (M = 186/188Re, 99mTc) could find use as radiopharmaceutical bioconjugates when R is a targeting moiety. However, the planar, electron-withdrawing 2-pyridyl groups of N(SO2R)dpa destabilize the ligand to base and create relatively rigid chelate rings, raising the possibility that the rare M– N(sulfonamide) bond is an artifact of a restricted geometry. Also, the hydrophobic 2-pyridyl groups could cause undesirable accumulation in the liver, limiting future use in radiopharmaceuticals. Our goal is to identify a robust, hydrophilic, and flexible N(CH2Z)2 chelate framework. New C2-symmetric ligands, N(SO2R)(CH2Z)2 with (Z = CH2NH2; R = Me, dmb, or tol), were prepared by treating N(H)dien(Boc)2, a protected diethylenetriamine (N(H)dien) derivative, with methanesulfonyl chloride (MeSO2Cl), 3,5-dimethylbenzenesulfonyl chloride (dmbSO2Cl), and 4-methylbenzenesulfonyl chloride (tolSO2Cl). Treatment of fac-[Re(CO)3(H2O)3]+ with these ligands, designated as N(SO2R)dien, afforded new fac-[Re(CO)3(N(SO2R)dien)]PF6 complexes. Comparing the fac-[Re(CO)3(N(SO2Me)dien)]PF6 and fac-[Re(CO)3(N(SO2Me)dpa)]PF6 complexes, we find that the ReI–N(sulfonamide) bonds are normal in length and statistically identical and that the methyl 13C NMR signal has an unusually upfield shift compared to that in the free ligand. We attribute this unusual upfield shift to the fact that the sulfonamide N undergoes an sp2-to-sp3 rehybridization upon coordination to ReI in both complexes. Thus, the sulfonamide N of N(SO2R)dien ligands is a good donor, even though the chelate rings are conformationally flexible. Addition of the strongly basic and potentially monodentate ligand, 4-dimethylaminopyridine, did not affect the fac-[Re(CO)3(N(SO2tol)dien)]PF6 complex, even after several weeks. This complex is also stable to heat in aqueous solution. These results indicate that N(SO2R)dien ligands form fac-[Re(CO)3(N(SO2R)dien)]PF6 complexes sufficiently robust to be utilized for radiopharmaceutical development.

Introduction

Many fac-[99mTc(CO)3L]n imaging agents with facially coordinated tridentate ligands (L) have been studied1-8 because of the convenient generation of the fac-[99mTc(CO)3(H2O)3]+ precursor.9,10 Some of these imaging agents have exhibited satisfactory results in human volunteers and in early patient studies.4,7,8 Such fac-[99mTc(CO)3L]n agents are more robust and have better pharmacokinetic properties than agents with bidentate ligands.11 The γ-emitting 99mTc radionuclide has ideal nuclear properties12,13 for diagnostic applications in nuclear medicine.1,14-17

The development of 99mTc radiopharmaceutical agents benefits from an understanding of the chemistry of their Re analogues. The discovery of a straightforward preparation of the fac-[Re(CO)3(H2O)3]+ precursor18 has led to significantly improved aqueous synthetic methods for fac-[M(CO)3L]n agents (M = various isotopes of Tc and Re).3,16,17,19-21 fac-[Re(CO)3L]n complexes serve as excellent structural models for fac-[99mTc(CO)3L]n imaging agents.4,22-27 Moreover, fac-[186/188Re(CO)3L]n agents themselves are emerging as promising radiopharmaceuticals, owing to their potential usefulness in radiotherapy.1,12,20,28

New types of tridentate ligands and ligand conjugation methods will expand the likelihood of developing useful new agents with the fac-[MI(CO)3]+ core (M = 99mTc, 186/188Re).10,29,30 Meeting such goals requires the identification of suitable linker systems with high stability, small size, and core ligands having a tridentate donor framework that does not increase the number of isomers.31-34 Symmetrical linear tridentate ligands with linkage at the center donor are thus suitable candidates because the generation of racemic or diastereoisomeric mixtures of radiopharmaceuticals can be avoided.35 Recent studies have reported promising biomedical properties for fac-[Re(CO)3L]n complexes bearing a tridentate L with three N donors having a substituent replacing the proton at the central sp3 N.16,29,36,37 However, the conjugation of biologically important groups in these complexes was limited to groups attached via an N–C bond to the central N of ligands with the N(CH2Z)2 tridentate ligand framework. Z is commonly an N donor (e.g., 2-pyridyl)16,36 or a carboxyl group.3 The use of a biologically compatible linking group that did not create an N–C bond would greatly increase the chances of discovering useful agents.

Molecules containing a sulfonamide represent a very important class of biologically active molecules with a wide variety of applications.39-45 Therefore, we previously set out to explore conjugation that utilizes an N–S bond with the central N being the sulfonamide N.35 The reaction of various sulfonyl chlorides (RSO2Cl) with di-(2-picolyl)amine (N(H)dpa) afforded N(SO2R)dpa ligands.35 [Note that we use the normal convention: an N (italic N) designates a substituent location on nitrogen in the name of a compound. In this article, N designates a substituent located on the central or anchoring nitrogen atom of a tridentate ligand.] These N(SO2R)dpa ligands readily added to fac-[Re(CO)3(H2O)3]+ to form fac-[Re(CO)3(N(SO2R)dpa)]PF6 (or BF4) complexes. We learned that these were the only examples of structurally characterized bearing an N-bound, open-chain tertiary sulfonamide linkage with a normal M–N bond length. These fac-[Re(CO)3(N(SO2R)dpa)]PF6 complexes were the first examples of such structurally characterized complexes with a neutral tertiary sulfonamide donor bound not only to the fac-[ReI(CO)3]+ core but to any metal center.35 Although the results in that report appeared to serve as proof of principle, the tridentate framework, N(CH2Z)2 with Z = 2-pyridyl, is relatively rigid and the resulting geometric constraints could possibly account for the observation of the M–N(sulfonamide) bond of normal length, as found in cases when complicated ligand ring structures fix the bond lengths.35 Thus, the study of fac-[Re(CO)3(N(SO2R)(CH2Z)2)]n complexes with more flexible chelate rings is of fundamental importance to coordination chemistry.

The new fac-[Re(CO)3(N(SO2R)dpa)]X complexes revealed the feasibility of having a tridentate ligand anchored by a central tertiary sulfonamide N. However, the two 2-pyridyl rings in a potential fac-[M(CO)3(N(SO2R)dpa)]n imaging agent are hydrophobic, an undesirable property in an imaging agent because it is expected to promote liver uptake.12 Also, the 2-pyridyl groups are electron withdrawing, facilitating decomposition of coordinated N(SO2R)dpa ligands by strong base.35

With the goals of exploring fundamental coordination chemistry of sulfonamides and of identifying ligands for use in imaging agents, we have now explored a more hydrophilic and more flexible ligand system that is suitable for our new conjugation method. Here we employ a prototypical triamine ligand framework based on diethylenetriamine (N(H)dien).46 In the new N(SO2R)dien ligands, the aromatic 2-pyridyl groups of the dpa moiety are replaced with hydrophilic –CH2NH2 groups. The new ligands are stable to base and have the advantage of being small in size, a feature considered to be desirable for bioconjugates.32,34 All of the new complexes discussed below have the facial geometry, and thus from this point onward we omit the fac- designation when discussing specific compounds.

Experimental Section

Starting Materials

Methanesulfonyl chloride (MeSO2Cl), 3,5-dimethylbenzenesulfonyl chloride (dmbSO2Cl), 4-methylbenzenesulfonyl chloride (tolSO2Cl), N,N,N-triethylamine, trifluoroacetic acid (TFA), N(H)dien, 2-(tert-butoxycarbonyloxyimino)-2-(phenylacetonitrile), 4-dimethylaminopyridine, and Re2(CO)10 were used as received from Aldrich. Aqueous [Re(CO)3(H2O)3]OTf (OTf = trifluoromethanesulfonate) was prepared by a known method.18

NMR Measurements

1H NMR and 13C NMR spectra were recorded on a 400 MHz Bruker spectrometer. Peak positions are relative to TMS or to solvent residual peak, with TMS as reference. All NMR data were processed with TopSpin and MestReNova software. NMR data not presented in the Experimental Section can be found in the Results Section or in Supporting Information.

Mass Spectrometric Measurements

High resolution mass spectra were recorded on a Bruker Ultraflex MALDI TOF TOF mass spectrometer and an Agilent 6210 ESI TOF LCMS mass spectrometer.

X-ray Data Collection and Structure Determination

Intensity data were collected at low temperature on a Bruker Kappa Apex-II DUO CCD diffractometer fitted with an Oxford Cryostream cooler with graphite-monochromated Mo Kα (λ = 0.71073 Å) radiation. Data reduction included absorption corrections by the multiscan method, with SADABS.47 The structures were determined by direct methods and difference Fourier techniques and refined by full-matrix least squares using SHELXL-97.48 All non-hydrogen atoms were refined anisotropically. All H atoms were visible in difference maps, but were placed in idealized positions, except for N–H hydrogen atoms, for which coordinates were refined. A torsional parameter was refined for each methyl group. Compound 6 has two independent formula units in the asymmetric unit.

Synthesis of N,N″-Bis(tert-butoxycarbonyl)diethylenetriamine (N(H)dien(Boc)2)

The N(H)dien(Boc)2 ligand was prepared in 96% yield by a known method.49 1H NMR signals (ppm) in CDCl3: 4.89 (br s, 1H, NH), 3.20 (m, 4H, 2CH2), 2.72 (t, 4H, 2CH2), 1.44 (s, 18H, 6CH3). These 1H NMR chemical shifts matched the previously reported values.49

General Synthesis of N(SO2R)dien

The following general procedure was employed to obtain the [N(SO2Me)dienH2](CF3CO2)2 (1) and N(SO2R)dien (R = dmb (2), R = tol (3)) ligands: a solution of the sulfonyl chloride (2 mmol) in 30 mL of dioxane was added dropwise over 2 h to a solution of N(H)dien(Boc)2 (0.61 g, 2 mmol) and triethylamine (0.28 mL, 2 mmol) in dioxane (100 mL) at room temperature. The reaction mixture was stirred at room temperature for 20 h and filtered to remove any precipitate. The solvent was completely removed by rotary evaporation, water (50 mL) was added to the resulting oil, and the product was extracted into CH2Cl2 (2 × 25 mL). The CH2Cl2 extracts were combined and washed again with water (2 × 25 mL) at pH ∼6. The organic layer was dried with anhydrous Na2SO4, and the CH2Cl2 was removed by rotary evaporation to yield an off-white solid. The Boc-protected diensulfonamide, N(SO2R)dien(Boc)2, was purified if necessary by column chromatography (silica gel column and a mixture of ethyl acetate/hexane). After characterization by 1H NMR spectroscopy, this N(SO2R)dien(Boc)2 product was then deprotected by dropwise addition of trifluoroacetic acid (0.16–0.24 mL, 2–3 mmol) to a CH2Cl2 solution (at ∼0 °C) of the compound (5 mL, 1 mmol). The reaction mixture was allowed to warm to room temperature, stirred for 16 h at room temperature, and then filtered; the filtrate was taken to dryness by rotary evaporation. Water at pH ∼8–9 (50 mL) was added to the residue, and the product was extracted into CHCl3 (2 × 25 mL). The combined CHCl3 extracts were dried over anhydrous Na2SO4, and the solvent was removed by rotary evaporation, leaving a white/off-white powder or, when R = Me, a yellow oil.

General Synthesis of [Re(CO)3(N(SO2R)dien)]PF6

A solution of the N(SO2R)dien ligand (0.1 mmol) in methanol (2 mL) was added to an aqueous solution of [Re(CO)3(H2O)3]OTf (5 mL, 0.1 mmol). Methanol (1–2 mL) was added to dissolve any precipitate that formed. The pH of the reaction mixture was adjusted to ∼6–7 with 0.5 M NaOH if necessary, and the clear reaction mixture was heated at reflux for 24 h. An excess (0.16 g, 1 mmol) of NaPF6 was added to the clear solution, and the precipitate that formed within ∼30 min was collected on a filter, washed with water, and air dried.

Synthesis of [N(SO2Me)dienH2](CF3CO2)2 ([1H2](CF3CO2)2)

The use of MeSO2Cl (0.16 mL, 2 mmol) in the general method described above yielded crude N(SO2Me)dien(Boc)2 as a brown oil (0.65 g, 85% yield). 1H NMR signals (ppm) in DMSO-d6: 6.92 (br s, 2H, 2NH), 3.14 (m, 4H, 2CH2), 3.06 (m, 4H, 2CH2), 2.89 (s, 3H, CH3), 1.37 (s, 18H, 6CH3).

The deprotection process described in the general method above afforded N(SO2Me)dien (1) as a pale-yellow, oily substance (yield: 0.004 g, 18%). 1H NMR signals (ppm) in CDCl3: 3.25 (t, J = 6.3 Hz, 4H, 2C(5/6)H2), 2.93 (s, 3H, CH3), 2.90 (t, J = 6.4 Hz, 4H, 2C(4/7)H2); in acetonitrile-d3: 3.14 (t, J = 6.5 Hz, 4H, 2C(5/6)H2), 2.86 (s, 3H, CH3), 2.76 (t, J = 6.5 Hz, 4H, 2C(4/7)H2); in DMSO-d6: 3.08 (broad t, 4H, 2C(5/6)H2), 2.91 (s, 3H, CH3), 2.67 (broad t, 4H, 2C(4/7)H2). 13C NMR signals (ppm) in CDCl3: 51.59 (C5/6), 40.70 (C4/7), 37.60 (CH3).

Compared to ligands with other R groups, isolation for R = Me produced a low yield, owing to the more hydrophilic nature of N(SO2Me)dien. To increase the amount of material available for synthesis of the complex, the general deprotection process was modified by adding a slight excess of trifluoroacetic acid to obtain the protonated ligand [N(SO2Me)dienH2](CF3CO2)2 ([1H2](CF3CO2)2). A mixture of N(SO2Me)dien(Boc)2 (0.38 g, 1 mmol) and trifluoroacetic acid (0.24 mL, 3 mmol) was stirred at room temperature for 16 h, and the resulting mixture was filtered; the white precipitate that was collected on a filter paper was washed several times with CH2Cl2 and air dried. The precipitate was then dissolved in methanol (3 mL), and the solution was filtered. CH2Cl2 (-10 mL) was added to the filtrate until cloudiness was observed. After 24 h the undisturbed solution yielded thin, colorless, needle-like crystals of the trifluoroacetate salt of the protonated ligand [N(SO2Me)dienH2](CF3CO2)2 ([1H2](CF3CO2)2) (0.29 g, 72% yield). 1H NMR signals (ppm) in DMSO-d6: 7.81 (br s, 6H, 2+NH3), 3.37 (t, J = 6.2 Hz, 4H, 2CH2), 3.04 (s, 3H, CH3), 3.00 (t, J = 6.3 Hz, 4H, 2CH2). ESI-MS: m/z calcd for C5H15O2N3S [M + H]+: 182.0958; found 182.0957.

Synthesis of N(SO2dmb)dien (2)

The use of dmbSO2Cl (0.40 g, 2 mmol) according to the general synthetic procedure above yielded a pale yellow oil, which was purified by dissolving the oil in a minimum (∼1 mL) of ethyl acetate and loading the solution onto a silica gel column. A 1:5 mixture of ethyl acetate:hexane was used to elute the remaining dmbSO2Cl starting material (UV-vis). The product (UV-vis) was then eluted with a 2:3 (v:v) mixture of ethyl acetate/hexane. Thin-layer chromatography was used to determine the progress of separation. Removal of solvent by rotary evaporation yielded N(SO2dmb)dien(Boc)2 as a pale yellow powder (0.87 g, 92% yield). 1H NMR signals (ppm) in CDCl3: 7.65 (s, 2H, H2/6), 7.21 (s, H, H4), 5.19 (br s, 2H, NH), 3.32 (m, 4H, 2CH2), 3.18 (m, 4H, 2CH2), 2.38 (s, 6H, 2CH3), 1.45 (s, 18H, 6CH3).

Deprotection of N(SO2dmb)dien(Boc)2 (0.47 g, 1 mmol) with trifluoroacetic acid (0.16 mL, 2 mmol), by the procedure described above, afforded N(SO2dmb)dien (2) as a white powder (0.26 g, 98% yield). 1H NMR signals (ppm) in CDCl3: 7.42 (s, 2H, H2/6), 7.20 (s, H, H4), 3.15 (t, J = 6.4 Hz 4H, 2C(5/6)H2), 2.91 (t, J = 6.4 Hz, 4H, 2C(4/7)H2), 2.38 (s, 6H, 2CH3). ESI-MS: m/z calcd for C12H21O2N3S [M + H]+: 272.1427; found 272.1431.

Synthesis of N(SO2tol)dien (3)

The general procedure using tolSO2Cl (0.44 g, 2 mmol) yielded N(SO2tol)dien(Boc)2 as a pale yellow oil (0.78 g, 86% yield). 1H NMR signals (ppm) in CDCl3: 7.67 (d, J = 8.3 Hz, 2H, H2/6), 7.31 (d, J = 8.1 Hz, 2H, H3/5), 5.18 (br s, 2H, NH), 3.32 (m, 4H, 2CH2), 3.17 (m, 4H, 2CH2), 2.43 (s, 3H, CH3), 1.45 (s, 18H, 6CH3).

Deprotection of N(SO2tol)dien(Boc)2 (0.45 g, 1 mmol) with trifluoroacetic acid (0.16 mL, 2 mmol) afforded N(SO2tol)dien (3) as a white powder (0.23 g, 92% yield). 1H NMR signals (ppm) in CDCl3: 7.70 (d, J = 8.2 Hz, 2H, H2/6), 7.31 (d, J = 8.0 Hz, 2H, H3/5), 3.15 (t, J = 6.4 Hz, 4H, 2C(5/6)H2), 2.91 (t, J = 6.4 Hz, 4H, 2C(4/7)H2), 2.43 (s, 3H, CH3); in acetonitrile-d3: 7.69 (d, J = 8.4 Hz, 2H, H2/6), 7.39 (d, J = 7.9 Hz, 2H, H3/5), 3.07 (t, J = 6.4 Hz, 4H, 2C(5/6)H2), 2.79 (t, J = 6.5 Hz, 4H, 2C(4/7)H2), 2.41 (s, 3H, CH3). ESI-MS: m/z calcd for C11H19O2N3S [M + H]+: 258.1271; found 258.1277.

Synthesis of [Re(CO)3(N(SO2Me)dien)]PF6 (4)

The general method above, with the [1H2](CF3CO2)2 ligand (0.04 g, 0.1 mmol) and [Re(CO)3(H2O)3]OTf (5 mL, 0.1 mmol), afforded [Re(CO)3(N(SO2Me)dien)]PF6 (4) as a white crystalline precipitate (0.036 g, 60% yield) after the addition of NaPF6 (∼16 mg). (The pH of the acidic reaction mixture was adjusted to ∼7 with 0.5 M NaOH before the solution was heated at reflux.) 1H NMR signals (ppm) in DMSO-d6: 5.65 (m, 2H, NH), 4.27 (m, 2H, NH), 3.57 (s, 3H, CH3), 3.45 (m, 2H, CH2), 3.22 (m, 2H, CH2), 3.10 (m, 2H, CH2), 2.97 (m, 4H, 2CH2). 1H NMR signals (ppm) of the R group in acetone-d6: 3.62 (s, 3H, CH3); in acetonitrile-d3: 3.33 (s, 3H, CH3). Chelate ring 1H NMR signals for 4-6 in these solvents are reported in the Results Section. 13C NMR signals (ppm) in DMSO-d6: 55.91 (C5/6), 44.86 (C4/7), 32.50 (CH3). 13C NMR signals (ppm) in acetone-d6: 55.11 (C5/6), 44.46 (C4/7), 30.97(CH3). MALDI-TOF: m/z calcd for C8H15O5N3SRe [M+]: 452.029; found: 452.158.

Synthesis of [Re(CO)3(N(SO2dmb)dien)]PF6 (5)

The use of the general procedure with N(SO2dmb)dien (2) (0.03 g, 0.1 mmol) afforded [Re(CO)3(N(SO2dmb)dien)]PF6 (5) as a white precipitate (0.047 g, 69% yield) after the addition of NaPF6 (∼16 mg). 1H NMR signals (ppm) in DMSO-d6: 7.72 (s, 2H, H2/6), 7.57 (s, H, H4), 5.72 (m, 2H, NH), 4.28 (m, 2H, NH), 3.41 (m, 2H, CH2), 3.14 (m, 2H, CH2), 3.00 (m, 2H, CH2), 2.65 (m, 2H, CH2), 2.42 (s, 6H, 2CH3). 1H NMR signals (ppm) of the R group in acetone-d6: 7.77 (s, 2H, H2/6), 7.60 (s, H, H4), 2.46 (s, 6H, 2CH3); in acetonitrile-d3: 7.65 (s, 2H, H2/6), 7.55 (s, H, H4), 2.44 (s, 6H, 2CH3). 13C NMR signals (ppm) in acetone-d6: 142.30 (dmb ring C1), 139.54 (dmb ring C4), 130.93 (dmb ring C2/6), 126.71(dmb ring C3/5), 57.83 (C5/6), 46.93(C4/7), 22.09 (2CH3). MALDI-TOF: m/z calcd for C15H21O5N3SRe [M+]: 542.076; found:542.155.

Synthesis of [Re(CO)3(N(SO2tol)dien)]PF6 (6)

The use of N(SO2tol)dien (3) (0.02 g, 0.1 mmol) as described in the general synthesis afforded [Re(CO)3(N(SO2tol)dien)]PF6 (6) as a white precipitate (0.056 g, 84% yield) after the addition of NaPF6 (∼16 mg). Slow evaporation of a solution of the compound in acetone produced colorless, X-ray quality, needle-shaped crystals. 1H NMR signals (ppm) in DMSO-d6: 7.99 (d, J = 8.5 Hz, 2H, H2/6), 7.61 (d, J = 8.1 Hz, 2H, H3/5), 5.70 (m, 2H, NH), 4.28 (m, 2H, NH), 3.40 (m, 2H, CH2), 3.14 (m, 2H, CH2), 3.01 (m, 2H, CH2), 2.66 (m, 2H, CH2), ∼2.5, overlapped (s, 3H, CH3). 1H NMR signals (ppm) of the R group in acetone-d6: 8.05 (d, 2H, H2/6), 7.65 (d, 2H, H3/5), 2.53 (s, 3H, CH3); in acetonitrile-d3: 7.91(d, 2H, H2/6), 7.58 (d, 2H, H3/5), 2.51 (s, 3H, CH3). 13C NMR signals (ppm) in DMSO-d6: 147.80 (tol ring C4), 132.07 (tol ring C2/6), 131.00 (tol ring C3/5), 125.98 (tol ring C1), 56.20 (C5/6), 44.80 (C4/7), 21.69 (CH3). 13C NMR signals (ppm) in acetone-d6: 148.88 (tol ring C4), 132.70 (tol ring C2/6), 131.58 (tol ring C3/5), 127.05 (tol ring C1), 56.94(C5/6), 46.00 (C4/7), 21.74 (CH3). MALDI-TOF: m/z calcd for C14H19O5N3SRe [M +]: 528.060;found: 528.001.

Challenge Reactions of [Re(CO)3(N(SO2tol)dien)]PF6 (6)

Two 5 mM solutions of 6 in DMSO-d6 (500 μL) were treated separately with 10 molar equiv (0.6 μL, 50 mM) or 2 molar equiv (0.12 μL, 10 mM) of diethylenetriamine and monitored over time by 1H NMR spectroscopy. Another challenge experiment was carried out with a 5 mM solution of 6 in DMSO-d6 by first adding 1 molar equiv of 4-dimethylaminopyridine (4-Me2Npy) and then increasing the 4-Me2Npy to 10 molar equiv.

Results and Discussion

Synthesis of N(SO2R)dien and [Re(CO)3(N(SO2R)dien)]PF6

N(SO2Me)dien (1), N(SO2dmb)dien (2), and N(SO2tol)dien (3) were synthesized by coupling N(H)dien with the respective sulfonyl chloride (Figure 1). In this case, unlike in the synthesis of N(SO2R)dpa ligands,35 additional steps are required to synthesize the N(SO2R)dien ligands (13), owing to the possibility of attack by the terminal N(H)dien amino groups on the sulfur atom of the sulfonyl chloride. Thus, the N(H)dien terminal amine groups were protected with Boc groups. The products were obtained in good yield and purity, as indicated by 1H NMR spectral data. Reaction of the deprotected ligands with an aqueous solution of [Re(CO)3(H2O)3]OTf afforded [Re(CO)3(N(SO2Me)dien)]PF6 (4), [Re(CO)3(N(SO2dmb)dien)]PF6 (5), and [Re(CO)3(N(SO2tol)dien)]PF6 (6) in 60–84% yields. All three complexes were characterized by 1H and 13C NMR spectroscopy and by mass spectrometry (Experimental Section).

Figure 1.

Figure 1

General reaction scheme for the synthesis of N(SO2R)dien ligands (top) and [Re(CO)3(N(SO2R)dien)]+ complexes (bottom).

X-ray quality crystals of 4 and 6 were grown by slowly cooling a 5 mM solution of the compound in warm water. Crystals of 5 (grown by slow evaporation of a 5 mM acetone solution) suffered from twinning, space-group ambiguity, and disorder problems in the molecule, which prevented refinement within acceptable values. However, the structural characterization of 5 (using the same procedures as for 4 and 6) confirmed that all three N atoms of N(SO2dmb)dien are coordinated to Re.

Structural Results

Crystal data and structural refinement details for [Re(CO)3(N(SO2Me)dien)]PF6 (4) and [Re(CO)3(N(SO2tol)dien)]PF6 (6) are summarized in Table 1, and the ORTEP plots appear in Figure 2 (see Figure 1 for the numbering scheme used to describe the solid-state data). Both 4 and 6 exhibit a pseudo octahedral structure, with the three carbonyl ligands occupying one face; the remaining three coordination sites are occupied by the three nitrogen atoms of the tridentate ligand.

Table 1. Crystal Data and Structure Refinement for [Re(CO)3(N(SO2Me)dien)]PF6 (4) and [Re(CO)3(N(SO2tol)dien)]PF6 (6).

4 6
empirical formula C8H15N3O5ReS·PF6 C14H19N3O5ReS PF6
fw 596.46 672.55
crystal system monoclinic monoclinic
space group P21/c P21/c
unit cell dimensions
a (Å) 12.822(3) 26.0233(14)
b (Å) 8.8089(19) 13.0607(7)
c (Å) 14.026(3) 12.2147(7)
β(deg) 94.373(10) 95.831(3)
V3) 1579.6(6) 4130.1(4)
T (K) 100 100
Z 4 8
ρcalc (Mg/m3) 2.508 2.163
abs coeff (mm-1) 8.02 6.15
2θmax (°) 70 71.2
R indicesa 0.018 0.029
wR2 = [I> 2σ(I)]b 0.044 0.055
data collected, Rint 28379, 0.034 51479, 0.033
data/param 6953/239 18802/585
a

R = (Σ║Fo│ - │Fc║)/Σ │Fo│;

b

wR2= [Σ[w(Fo2 - Fc2)2]/Σ [w(Fo2)2]]1/2 in which and w = 1/[σ2(Fo2) + (dP)2 + (eP)] and P = (Fo2 + 2Fc2)/3.

Figure 2.

Figure 2

ORTEP plots of the cations in [Re(CO)3(N(SO2Me)dien)]PF6 (4) (left) and Re(CO)3(N(SO2tol)dien)]PF6 (6) (right). Thermal ellipsoids are drawn with 50% probability.

Selected bond distances and angles of complexes 4 and 6 are presented in Table 2. The Re–C bond distance of the CO group trans to the sulfonamide group is not significantly different from those of the two cis Re–CO bonds, indicating the absence of any trans influence. All Re–N bond distances in 6 are generally similar to the Re–N(sp3) distances observed in [Re(CO)3L]+ complexes with prototypical NNN donor ligands, which range from ∼2.21 to 2.29 Å as the bulk of substituents on the N atoms increases.46,50 However, the distances from Re to the central N2 (2.2763(14) Å in 4 and 2.2686(19) Å in 6) are significantly longer than the Re-N1 distances (2.2377(14) Å for 4 and 2.221(2) Å for 6) and Re–N3 distances (2.2072(15) Å for 4 and 2.216(2) Å for 6) (Table 2). These Re–N2 distances in 4 and 6 are significantly longer than the Re–N2 distance in [Re(CO)3(N(H)dien)]PF646 (2.201(3) Å), but they are comparable to the Re–N2 distance of [Re(CO)3(N(Me)dien)]PF6.46 The Re–N2 distance in the latter complex (2.250(4) Å) is not statistically different from the Re–N2 distance in [Re(CO)3(N(SO2tol)dien)]PF6 (6, Table 2), suggesting that the bond lengthening in 6 is caused chiefly by steric effects, not electronic effects. Furthermore, the good overlap of the three donor N atoms observed when the three carbonyl carbon (C1, C2 and C3) and Re atoms in the structure of 6 are superimposed with the corresponding atoms in [Re(CO)3(N(Me)dien)]PF646 (Supporting Information, Figure S1) demonstrates the lack of any large effect of having a tertiary sulfonamide nitrogen instead of a typical sp3 nitrogen serve as the central anchor of the tridentate ligand.

Table 2. Selected Bond Distances (Å) and Angles (deg) for [Re(CO)3(N(SO2Me)dien)]PF6 (4) and [Re(CO)3(N(SO2tol)dien)]PF6 (6).

4 6 4 6
bond distance bond angle
Re–N1 2.2377(14) 2.221(2) N1–Re–N2 77.07(5) 77.54(7)
Re–N2 2.2763(14) 2.2686(19) N1–Re–N3 86.53(6) 85.83(8)
Re–N3 2.2072(15) 2.216(2) N2–Re–N3 77.80(5) 77.86(7)
Re–C1 1.9128(16) 1.913(2) C2–Re–N1 95.72(6) 92.70(9)
Re–C2 1.9109(18) 1.911(2) C3–Re–N1 92.31(6) 94.99(9)
Re–C3 1.9136(19) 1.917(3) C2–Re–N3 96.60(7) 98.29(9)
S1–O5 1.4273(13) 1.4280(18) C1–Re–N3 93.12(6) 91.92(9)
S1–O4 1.4324(13) 1.4339(17) Re–N2–S1 111.07(7) 111.35(9)
S1–N2 1.7434(13) 1.759(2) Re–N2–C5 105.93(9) 106.14(13)
Re–N2–C6 110.01(10) 109.86(13)
non-bonded distance S1–N2–C5 109.13(10) 108.82(14)
N1–N3 3.046(2) 3.021(3) S1–N2–C6 108.52(10) 108.18(13)
N1–N2 2.812(2) 2.812(3) C5–N2–C6 112.19(13) 112.52(18)
N2–N3 2.816(2) 2.818(3) C1–Re–N2 99.98(6) 100.55(8)
C3–Re–N2 97.91(6) 96.81(8)
deviation of N2a 0.513(1) 0.514(2) C1–Re–C2 87.25(7) 89.08(10)
C3–Re–C2 87.62(7) 87.29(10)
C1–Re–C3 87.83(7) 87.10(10)
a

Distance to the sulfonamide N from the plane defined by the S and the two C atoms attached to N2 (average of two independent molecules for 6).

The lengthening of the Re–N2 bond in 4 and 6 versus [Re(CO)3(N(H)dien)]PF6 is similar to that observed in [Re(CO)3(N(SO2R)dpa)]PF6 complexes (e.g., R = Me and tol) versus the parent [Re(CO)3(N(H)dpa)]Br complex.30,35 The similarity of the central Re–N bond distances of the analogous [Re(CO)3(N(SO2R)dpa)]PF6 complexes (average = ∼2.277 Å) and in 4 and 6 (average = ∼2.272 Å) indicates that the donating ability of the sulfonamide N is similar for both ligands.35 Furthermore, all three donor N atoms show good overlap when the C1, C2, C3, and Re atoms in [Re(CO)3(N(SO2tol)dien)]PF6 and [Re(CO)3(N(SO2tol)dpa)]PF6 are superimposed (Figure 3). The central Re–N bond distance of the [Re(CO)3(N(SO2Me)dpa)]PF6 complex (2.2826(16) Å) is not significantly different from this distance in 4. Thus, the normal length of the Re–N(sulfonamide) bonds in [Re(CO)3(N(SO2R)dpa)]PF6 complexes is not a consequence of the rigidity of the 5-membered dpa chelate rings.

Figure 3.

Figure 3

Overlay of Re, C1, C2, and C3 atoms of [Re(CO)3(N(SO2tol)dien)]PF6 (6) (blue) and [Re(CO)3(N(SO2tol)dpa)]PF6 35 (magenta) complexes. (The SO2R groups have been omitted for clarity).

The sulfonamide nitrogen of the N(SO2tol)dpa ligand in a PdII complex is not bound to PdII and all bond angles around the sulfonamide nitrogen of the N(SO2tol)dpa ligand in the bidentate binding mode are close to 120°, indicating sp2 hybridization for this N atom.52 In contrast, in 4 and 6, all bond angles around the sulfonamide N (N2) are ∼109° (Table 2). This observation, which is consistent with data for [Re(CO)3(N(SO2R)dpa)]PF6 complexes,35 indicates that the sulfonamide nitrogen changes from sp2 to sp3 hybridization upon tridentate binding to ReI.35 This conclusion is confirmed by the fact that the C5–N2 (average = 1.511 Å) and C6-N2 (average = 1.512 Å) bond distances are longer than an average C–N(sp2) distance (∼1.28 Å).53 Data for all of the bond angles and distances involving the sulfonamide N indicate that it is sp3 hybridized.

In [Re(CO)3(N(SO2R)dpa)]PF6 and other complexes that have metal-bound tertiary sulfonamide groups, the distance to the sulfonamide N from the plane defined by the S and the two C atoms attached to N typically ranges from 0.47–0.52 Å.35 Such values are larger than those for the corresponding distance in complexes with an unbound sulfonamide N atom (0.06–0.26 Å).35 In the new complexes, [Re(CO)3(N(SO2Me)dien)]PF6 (4) and [Re(CO)3(N(SO2tol)dien)]PF6 (6), this distance (Table 2) is in the range characteristic of a M–N bond.

Most of the known complexes with tertiary sulfonamide ligands exhibit relatively long M–N(sulfonamide) bond distances and short S–N distances, features attributed to the resonance contribution of the lone pair on the sulfonamide N and to the poor electron donation of the sulfonamide N to the metal.54,55 Thus, both the comparatively short Re–N(sulfonamide) bond distance [2.2763(14) for 4 and 2.2686(19) Å for 6] and the long S–N bond distance in these complexes [1.7434(13) for 4 and 1.759(2) Å) for 6] indicate that the sulfonamide N is a relatively strong donor.

Only one other structurally characterized complex in which a diethylenetriamine derivative is bound to a metal (LaIII) via a sulfonamide bond through the N has been reported;56 the sulfonamide group in that complex, however, is part of a cyclic heptadentate ligand. Thus, the sulfonamide complexes reported here are the first structurally characterized metal complexes in which the metal is coordinated by the central tertiary sulfonamide N in a linear, highly flexible tridentate ligand.

We have previously discussed at length chelate ring pucker λ or δ chirality and the consequences on structure and NMR spectra of [Re(CO)3(polyamine)]X complexes.38,46,50 The cation of [Re(CO)3(N(SO2Me)dien)]PF6 (4) and one of the cations of [Re(CO)3(N(SO2tol)dien)]PF6 (6) (the one shown in Figure 2) have five-membered chelate rings of different chirality (λ or δ). However, the other independent cation in the asymmetric unit of 6 (not shown) has chelate rings of the same chirality. The relative frequency of observing the same or different chiralities in the two chelate rings with a central tertiary sulfonamide N donor (4 and 6) suggests that structures in which the chirality of the rings differs may be slightly more stable than structures having rings of the same chirality. Thus, our finding agrees with cases in which the central N is a classical sp3 N donor, namely that examples in which the chirality of the rings in a given structure differs (λδ or δλ) occur more frequently than those in which the chirality is the same (λλ or δδ).46

NMR Spectroscopy

The N(SO2R)dien ligands and the [Re(CO)3(N(SO2R)dien)]PF6 complexes reported here were characterized by 1H NMR (16) and 13C NMR (1, 36) spectroscopy, usually in one or more of several solvents (CDCl3, acetone-d6, acetonitrile-d3 and DMSO-d6). NMR signals were assigned by analyzing the splitting pattern, integration, and data from 2D NMR experiments.

The two methylene 1H NMR signals in N(SO2R)dien ligands (13) are triplets integrating to four protons in CDCl3 (Experimental Section and Figure S2). Selected data were obtained in acetonitrile-d3 or in DMSO-d6 (Experimental Section and Figure S3). The C(5/6)H2 triplet for 2 (R = dmb) and for 3 (R = tol) is slightly upfield from the C(5/6)H2 triplet for 1 (R = Me) in CDCl3 (Figure S2). Shift values for the C(4/7)H2 triplet are similar for all three ligands. These observations are attributed to the anisotropic sulfonamide aromatic groups, which are close to the C(5/6)H2 groups but far from the C(4/7)H2 groups. Similar results were obtained for 1 and 3 in acetonitrile-d3. The two protons in each methylene group are not sterically similar because the sulfonamide N lies out of the plane defined by the S and the two C atoms attached to N; the apparent magnetic equivalence in the NMR spectra indicates that inversion at the sulfonamide nitrogen leads to time averaging, as discussed previously for the N(SO2R)dpa ligands.35

The spectra of the respective complexes, on the other hand, are consistent with each of the four protons in the chelate ring giving rise to a multiplet. For example, the methylene group 1H NMR signals observed in acetonitrile-d3 changed from two triplets (at 3.07 and 2.79 ppm) for the N(SO2tol)dien ligand (3) to four multiplets [at 3.51, 3.16 (two overlapped), and 2.59 ppm; see Experimental Section and Supporting Information, Figure S4] for [Re(CO)3(N(SO2tol)dien)]PF6 (6). The least amount of overlap of the multiplets for complexes 46 was found in acetone-d6 (Figure 4). Thus, as expected, the apparent magnetic equivalence of methylene protons in the free ligands is lost in 46. Similar changes in signals for methylene protons upon coordination of ligands have been reported for [Re(CO)3(N(SO2R)dpa)]PF6 complexes35 and for [Re(CO)3(N(H)dien)]PF6.46

Figure 4.

Figure 4

1H NMR spectra of [Re(CO)3(N(SO2R)dien)]PF6 complexes (46) in acetone-d6 at 25 °C.

We designate the magnetically distinct protons in −CH2– or −NH2 groups as endo-H or exo-H protons, on the basis of the orientation of the proton either toward (endo) or away (exo) from the carbonyl ligands (Chart 1). Note that the chelate rings in these complexes are puckered, and the corresponding protons in the two chelate rings are not equivalent in the solid state. However, the chelate rings are fluxional, and time averaging leads to only one signal for each type of CH or NH proton, or six 1H NMR signals for the N(C(5/6)H2–C(4/7)H2–NH2)2 tridentate framework moiety.

Chart 1.

Chart 1

Front (left) and side (right) views of [Re(CO)3(N(SO2tol)dien)]PF6 (6), showing the designation of endo- and exo-CH and endo- and exo-NH protons.

1H NMR Assignments

1H NMR signals of the protons in the terminal NH2 groups of previously reported [Re(CO)3(L)]n compounds (L = simple tridentate ligands similar to diethylenetriamine with two terminal −NH2 groups)46,50 are well resolved. The endo-NH signal has a more downfield chemical shift than the exo-NH signal; solvent has access to the exposed endo-NH protons, but solvent access to the exo-NH protons is impeded by the chelate rings.46,50 The signal of the exposed endo-NH protons is shifted downfield by NH-solvent H-bonding. Both the shifts and the characteristic shift separation between these signals (Δ = ∼1.5 ppm in DMSO-d6) are useful for assigning such NH NMR signals.46,50

In this work, two broad 1H NMR signals of [Re(CO)3(N(SO2tol)dien)]PF6 (6) (5 mM) in DMSO-d6 at 4.28 and 5.70 ppm decreased in size when D2O was added, indicating that they are NH signals. The signals are connected by a strong COSY cross-peak, as expected for an NH2 group (Figure S5). The downfield signal at 5.70 ppm and the relatively upfield signal at 4.28 ppm, each integrating to two protons with Δ = ∼1.5 ppm, can be assigned as endo-NH and exo-NH signals, respectively.46,50

Similar Δ values of∼1.5 ppm (DMSO-d6), ∼1.1 ppm (acetone-d6), and ∼1.0 ppm (acetonitrile-d3) were observed between the downfield and upfield NH signals for all three complexes studied here (Table 3). This consistency aided in assigning the endo-NH and exo-NH signals in complexes 46 (Table 3, Figure 4). Although X-ray quality crystals for compound 5 have not been isolated, the 1H NMR patterns and shifts of its chelate ring signals in all solvents used match the corresponding 1H NMR data for the crystallographically characterized compound 6 (Table 3 and Figure 4). These 1H NMR chemical shifts provide strong evidence that 5 has a structure very similar to that of 6. The NH signals of 6 in different solvents (Figure 5) are shifted downfield as the H-bonding ability of the solvent increases. Thus, significant downfield chemical shifts are observed in acetone-d6 and DMSO-d6, in contrast to the relatively upfield chemical shifts observed in acetonitrile-d3 (Table 3 and Figure 5), owing to the weak interactions of that solvent with the NH groups. The highest Δ value (∼1.5 ppm) occurs in DMSO-d6 because it has the greatest H-bonding ability among the three solvents.

Table 3. 1H NMR Shifts (ppm) of exo- and endo-NH and exo- and endo-CH Signals for [Re(CO)3(N(SO2R)dien)]PF6 Complexes in Various Solvents at 25 °C.

Signal R = Me (4) R = dmb (5) R = tol (6)
DMSO-d6
exo-NH 4.27 4.28 4.28
endo-NH 5.65 5.72 5.70
endo-C(5/6)H 3.45 3.41 3.42
exo-C(4/7)H 3.23 3.14 3.14
endo-C(4/7)H 3.10 3.00 3.01
exo-C(5/6)H 2.97 2.65 2.66
acetone-d6
exo-NH 4.45 4.40 4.40
endo-NH 5.54 5.52 5.51
endo-C(5/6)H 3.85 3.81 3.81
exo-C(4/7)H 3.69 3.58 3.55
endo-C(4/7)H ∼3.50 3.50 3.52
exo-C(5/6)H ∼3.50 3.02 3.02
acetonitrile-d3
exo-NH 3.44 3.41 3.37
endo-NH 4.46 4.52 4.51
endo-C(5/6)H 3.50 3.52 3.51
exo-C(4/7)H ∼3.18 ∼3.16 ∼3.16
endo-C(4/7)H ∼3.18 ∼3.16 ∼3.16
exo-C(5/6)H 3.00 2.60 2.59

Figure 5.

Figure 5

1H NMR spectra of [Re(CO)3(N(SO2tol)dien)]PF6 (6), illustrating the relative position of NH signals observed at 25 °C in a) acetonitrile-d3, b) acetone-d6, and c) DMSO-d6.

The endo-NH and exo-NH 1H NMR signals assignments provide a starting point that allows us to exploit our structural data to achieve complete and unambiguous assignment of the ethylene 1H NMR signals, which are rarely well resolved for coordinated ligands having such chelate rings.50 To demonstrate our assignment strategy of assigning the multiplets to a specific endo-CH or exo-CH proton (Chart 1), we use [Re(CO)3(N(SO2tol)dien)]PF6 (6) in acetone-d 6 because of the low degree of overlap of signals for the ethylene group (Figure 4). The atom-numbering system used in this discussion appears in Figure 1. Cross-peaks in COSY spectra (Supporting Information, Figure S5) and ROESY spectra (Figure 6 and Supporting Information, Figure S6) identify a multiplet as a C(4/7)H or C(5/6)H signal. In the solid state, the endo-NH protons have a shorter average distance to the endo-C(4/7)H protons (2.20 Å) than to the exo-C(4/7)H protons (2.71 Å), and the exo-NH protons have a shorter average distance to the exo-C(4/7)H protons (2.22 Å) than to the endo-C(4/7)H protons (2.43 Å). Thus, from NH-CH NOE cross-peak intensities, the CH multiplets at 3.52 ppm and 3.55 ppm are assigned to the endo- and exo-C(4/7)H protons, respectively. The average distance from the exo-C(4/7)H protons is shorter to the exo-C(5/6)H protons (2.30 Å) than to the endo-C(5/6)H protons (2.85 Å). In the ROESY spectrum, a strong NOE cross-peak from the exo-C(4/7)H multiplet to the most upfield multiplet at 3.02 ppm thus assigns the multiplet to the exo-C(5/6)H protons. Similarly, an NOE cross-peak from the endo-C(4/7)H multiplet to the most downfield multiplet at 3.81 ppm assigns this multiplet to the endo-C(5/6)H protons. The tosyl methyl singlet (2.53 ppm) of 6 has an NOE cross-peak to the tosyl doublet at 7.65 ppm, unambiguously assigning it to the tosyl H3/5 protons and the other tosyl doublet to H2/6. The latter is more downfield (8.05 ppm), consistent with the proximity of the H2/6 protons to the sulfonamide group, shown in previous work on the [Re(CO)3(N(SO2tol)dpa)]PF6 analogue to have an electron-withdrawing inductive effect.35 A strong H2/6-to-endo-C(5/6)H NOE cross-peak and a very weak H2/6-to-exo-C(5/6)H cross-peak is consistent with the shorter H2/6-to-endo-C(5/6)H average distance (2.60 Å) compared to the H2/6-to-exo-C(5/6)H average distance (2.81 Å), thus confirming the assignment of the endo-C(5/6)H and exo-C(5/6)H signals (Table 3).

Figure 6.

Figure 6

1H-1H ROESY spectrum of [Re(CO)3(N(SO2tol)dien)]PF6 (6) in acetone-d6 at 25 °C. An expanded version of this figure is in Supporting Information (Figure S6).

1H NMR Shift Interpretation

The exo-C(5/6)H signal is more upfield for 5 and 6 than for 4 in acetone-d6 (Figure 4). In 6 (Figure 2 and Chart 1), the relative average distance from the centroid of the tosyl ring to the exo-C(5/6)H protons (which point toward the center of the ring) is shorter (3.45 Å) than the distance (4.09 Å) to the endo-C(5/6)H protons (Chart 1). Thus, the anisotropic effect of the aromatic ring of the R group explains the unusually upfield shift of the exo-C(5/6)H multiplet at 3.02 ppm in acetone-d6 for 5 and 6. This explanation is supported by the fact that the exo-C(5/6)H signal is not so upfield for 4, a complex in which R lacks an aromatic ring. This same trend, with the methylene endo-CH signal downfield from the exo-CH signal and the latter signal appearing more upfield when R has an aromatic ring (tol) than when R = Me, was reported for the [Re(CO)3(N(SO2R)dpa)]X complexes.35

Coordination of a ligand normally leads to downfield 1H NMR shift changes attributable to the metal inductive effect. Upon coordination to ReI of the free N(SO2R)dien ligands 1 and 3 in acetonitrile-d3, the downfield shift for signals of the C4/7 protons is ∼ 0.4 ppm. For example, the C(4/7)H2 triplet for 1 shifts from 2.76 ppm downfield to ∼3.18 ppm for the overlapping C(4/7)H signals of [Re(CO)3(N(SO2Me)dien)]PF6 (4). A corollary of the relationships of the distance of the exo-C(5/6)H protons to the aromatic groups mentioned above for 6 is that the endo-C(5/6)H protons are farther from R, and the shift of the endo-C(5/6)H signal for 4 and 6 is not influenced by the R group. Hence, compared to the C(5/6)H2 triplets of 1 and 3, the endo-C(5/6)H signals of 4 and 6 are shifted downfield (∼0.4 ppm to ∼3.5 ppm in acetonitrile-d3). Remarkably, upon coordination of N(SO2Me)dien to form 4, the exo-C(5/6)H signal (expected to shift downfield by ∼0.4 ppm owing to inductive effects) is shifted slightly upfield (0.14 ppm) compared to the free ligand triplet, even though R = Me is not anisotropic. Presently there are not enough other complexes with assigned signals for us to interpret the factors influencing shift, but the upfield shift may be caused by the SO2 group anisotropy. Also, although the reaction of free amines with acetone prevents comparison using acetone-d6, in DMSO-d6 (the other solvent rather different from acetonitrile-d3 that we used for the complexes) the same pattern of shift changes was observed upon coordination of 1 to form 4, including the slight upfield shift of the exo-C(5/6)H signal (Figure S3). Thus, solvent effects are probably not causing the unexpected upfield shift.

The methyl 1H NMR signal of N(SO2Me)dien (1) shifted downfield upon coordination of 1 to form [Re(CO)3(N(SO2Me)dien)]PF6 (4) in both acetonitrile-d3 (0.47 ppm) and DMSO-d6 (0.66 ppm) (Experimental Section and Supporting Information, Figure S3). These downfield shifts can be attributed to the ReI inductive effect. We previously attributed a similar downfield shift (∼0.76 ppm) observed for [Re(CO)3(N(SO2Me)dpa)]PF6 to the ReI inductive effect.35 A slightly greater downfield shift (∼1.03 ppm) relative to the free ligand reported for the [Re(CO)3(N(Me)dien)]PF6 complex in DMSO-d646 can be explained by the transmission of the inductive effect through just two bonds (N–C and C–H) versus three bonds in 4 (N–S, S–C, and C–H).

13C NMR Shift Interpretation

Unlike 1H NMR signals, the shifts of 13C NMR signals are not very dependent on solvent effects (e.g., as discussed for the NH signals) or on aromatic ring anisotropic effects (e.g., as discussed for exo-C(5/6)H signals). Instead, 13C NMR signals are influenced more significantly by other factors, with signals of aliphatic carbons most frequently being shifted downfield by metal inductive effects. Analysis of 13C NMR shifts can often provide insight into the effect of a diamagnetic metal center on the ligand. Therefore, we undertook the unambiguous assignment of the 13C NMR signals of selected ligands and complexes as described in Supporting Information. The 13C NMR signals of the chelate ring carbons shift downfield, as expected, by 4 to 5 ppm upon coordination of the.tridentate ligands 1 and 3 to form complexes 4 and 6, and the shifts for the C4/7 and C5/6 signals of 4 are very similar to the shifts for these signals of 6 (data for acetonitrile-d3 are reported in Table S1).

The methyl group 13C NMR signal for the free N(SO2Me)dien ligand (1) appears at 37.87 ppm, but it is shifted considerably upfield for [Re(CO)3(N(SO2Me)dien)]PF6 (4) (33.02 ppm, Table S1 in Supporting Information). The almost 5 ppm upfield change in shift when 1 forms 4 is unusual in that 13C NMR shifts are influenced mainly by through-bond effects, such as the metal inductive effect expected to produce a downfield shift. For example, a significant ∼14 ppm downfield 13C NMR shift was observed for the methyl group of [Re(CO)3(N(Me)dien)]PF6 (56.48 ppm) compared to the methyl group in the free N(Me)dien ligand (42.72 ppm) in acetonitrile-d3.51 Because of the unexpected upfield direction of the shift change of the methyl 13C NMR signal of 4, we measured the 13C NMR shift of the methyl signal of [Re(CO)3(N(SO2Me)dpa)]PF6 in DMSO-d6 (32.88 ppm), a value ∼6.5 ppm upfield compared to the methyl signal in the free N(SO2Me)dpa ligand (39.24 ppm).51 We attribute the observed unusual upfield shift for 4 to the fact that the sulfonamide N undergoes an sp2-to-sp3 rehybridization upon coordination to ReI in both 4 and [Re(CO)3(N(SO2Me)dpa)]PF6.

Challenge Reactions of [Re(CO)3(N(SO2tol)dien)]PF6 (6) in DMSO- d6

No change was observed in the 1H NMR signals of 6 (5 mM), even after six months, indicating that 6 is robust. However, addition of diethylenetriamine to 5 mM solutions of 6 to create solutions that were 50 mM or 10 mM in diethylenetriamine led to the elimination of all peaks for 6 by the next day and after about 6 days (t1/2 ∼28 h), respectively. In both cases, the final 1H NMR spectrum exhibits peaks for the N(SO2tol)dien ligand (3) and for the known [Re(CO)3(N(H)dien)]PF6 complex.46 No signals indicating any intermediates were observed. The complete displacement of the coordinated N(SO2tol)dien ligand of 6 by 10 mM diethylenetriamine establishes that although the Re–N bonds involving the central donor of N(SO2R)dien ligands have normal lengths, the donor ability of the central sulfonamide nitrogen atom is lower than that of the central nitrogen atom of diethylenetriamine.

A related study38 was reported for the neutral Re(CO)3(tmbSO2-dien) complex. The coordinated unsymmetrical monoanionic NNN donor ligand, tmbSO2-dien- (Chart 2), employed in that study has a 2,4,6-trimethylbenzenesulfonyl (tmbSO2) group linked to one of the terminal N atoms of diethylenetriamine. The terminal donor of the free ligand prior to coordination is a secondary sulfonamide of the type (RSO2)R'NH. When 10 molar equiv of diethylenetriamine was added to a solution of Re(CO)3(tmbSO2-dien) (3 mM, DMSO-d6), no coordinated tmbSO2-dien- ligand was displaced, even after the solution was heated at ∼60 °C.38 Thus, the deprotonated monoanionic tmbSO2-dien- ligand is clearly a better ligand than neutral N(SO2tol)dien.

Chart 2.

Chart 2

The tmbSO2-dien- monoanionic ligand.38

If the central tertiary sulfonamide donor in N(SO2tol)dien were coordinated weakly enough to the Re in [Re(CO)3(N(SO2R)dien)]PF6 complexes, a good monodentate ligand might be able to substitute for the coordinated sulfonamide group. Therefore, we conducted a challenge experiment in which the strongly basic, potentially monodentate 4-Me2Npy ligand was added to [Re(CO)3(N(SO2tol)dien)]PF6 (6). No 1H NMR spectral changes were observed over many days, even with a tenfold excess of 4-Me2Npy. Such evidence indicates that the central sulfonamide bond is strong enough to stay coordinated to Re, even in the presence of an excess of a strong monodentate ligand. Also, the results show that the [Re(CO)3(N(SO2tol)dien)]PF6 complex is resistant to strong base, an advantage over [Re(CO)3(N(SO2Me)dpa)]PF6, which decomposes readily in the presence of base.35

Solubility of [Re(CO)3(N(SO2tol)dien)]PF6 (6) in Water

A mixture of 5 mg of [Re(CO)3(N(SO2tol)dien)]PF6 (6) and 450 μL of D2O gave a clear solution when heated in a boiling water bath for 30 min. The 1H NMR spectrum of this solution, which compares favorably with spectra of 6 in other solvents (Figure 4), has four methylene multiplets, whereas the N(SO2tol)dien ligand has two methylene triplets in D2O. These results established that 6 dissolved unchanged to an extent much greater than is required for radiopharmaceuticals and that 6 is robust even in a hot aqueous solution. Furthermore, another experiment conducted using 5 mg of 6 also showed complete dissolution (observed visibly and by 1H NMR spectroscopy) of 6 in a 450 μL:20 μL of D2O:DMSO mixture (95.7%:4.3%). In contrast, similar NMR experiments carried out using the previously reported [Re(CO)3(N(SO2Me)dpa)]PF6 complex35 show that even the [Re(CO)3(N(SO2R)dpa)]PF6 complex having the smallest R group is insoluble in D2O or in D2O:DMSO (95.7%:4.3%).

Summary and Conclusions

From the present results with [Re(CO)3(N(SO2R)dien)]PF6 complexes, in which the chelate rings are less rigid, we conclude that the M–N bond of normal length observed was not a result of the rigid tridentate framework in the [Re(CO)3(N(SO2R)dpa)]X35 complexes. In both series of complexes, the methyl 13C NMR signal of the R = Me member of the series exhibited a very unusual upfield shift for an aliphatic carbon signal upon coordination of the ligand. This result, attributed to the similar sp2-to-sp3 rehybridization of the sulfonamide N upon coordination to ReI in both series of complexes, further establishes that coordination of the sulfonamide N is not influenced by the rigidity of the N(SO2R)dpa chelate rings. The decomposition of the [Re(CO)3(N(SO2R)dpa)]X complexes by base35 led us to hypothesize that base was attacking the coordinated N(SO2R)dpa ligand, most likely by deprotonating the CH2 group, and that the low electrophilicity of the Z = CH2NH2 group of the N(CH2CH2NH2)2 framework would confer stability toward base. The stability of the new [Re(CO)3(N(SO2R)dien)]PF6 complexes toward base supports these hypotheses and allowed us to conduct a challenge reaction study with the basic N(H)dien ligand. N(H)dien replaced the coordinated N(SO2tol)dien ligand in [Re(CO)3(N(SO2tol)dien)]PF6, indicating that the neutral sulfonamide N central donor in N(SO2tol)dien is a somewhat weaker donor than the central traditional sp3 N donor in N(H)dien. Nevertheless, the new [Re(CO)3(N(SO2R)dien)]PF6 complexes are long lived, even in the presence of base, and are relatively robust to heat treatment. As expected, the moderately hydrophilic character of the Z = CH2NH2 group of the N(CH2CH2NH2)2 framework also confers water solubility on the [Re(CO)3(N(SO2R)dien)]PF6 complexes. The aqueous solubility of the new ligands and complexes is much higher than necessary for radiopharmaceutical kit formulation. Also, there are perceived advantages in using small chelate ligands when constructing bioconjugates, and the N(CH2CH2NH2)2 framework is relatively small. Thus, the results obtained here suggest that N(SO2R)dien ligands should be explored in the development of radiopharmaceuticals, including bioconjugates.

Supplementary Material

1_si_001
2_si_002

Acknowledgments

This work was supported in part by the National Institutes of Health (R37 DK038842-22). An upgrade of the diffractometer was made possible by Grant No. LEQSF (2011-2012)-ENH-TR-01, administered by the Louisiana Board of Regents. L. G. M. thanks the RAYMOND F. SCHINAZI INTERNATIONAL EXCHANGE PROGRAMME between the University of Bath, UK, and Emory University, Atlanta, GA, USA, for a Faculty Fellowship.

Footnotes

Supporting Information Available. Crystallographic data for [Re(CO)3(N(SO2Me)dien)]PF6 (4) and [Re(CO)3(N(SO2tol)dien)]PF6 (6), in CIF format; figure illustrating overlay of the Re, C1, C2 and C3 atoms of 6 and [Re(CO)3(N(Me)dien)]PF6; 1H NMR spectra of ligands 13 in CDCl3; 1H NMR spectra of 1 and 4 in DMSO-d6; 1H NMR spectra of [Re(CO)3(N(SO2R)dien)]PF6 complexes (46) in acetonitrile-d3; COSY spectrum of 6 in acetone-d6; ROESY spectrum of 6 in acetone-d6; 13C NMR spectrum of 6 and assignment of 13C NMR signals of 6 in acetone-d6; aromatic region of the HMBC spectrum of 6 in acetone-d6; HSQC spectrum of 6 in acetone-d6; and a comparison of 13C NMR shifts of ligands (1 and 3) and complexes (4 and 6) in acetonitrile-d3. This material is available free of charge via the Internet at http://pubs.acs.org.

References

  • 1.Schibli R, Schubiger AP. Eur J Nucl Med Mol Imaging. 2002;29:1529–1542. doi: 10.1007/s00259-002-0900-8. [DOI] [PubMed] [Google Scholar]
  • 2.Liu S. Chem Soc Rev. 2004;33:445–461. doi: 10.1039/b309961j. [DOI] [PubMed] [Google Scholar]
  • 3.Lipowska M, Marzilli LG, Taylor AT. J Nucl Med. 2009;50:454–460. doi: 10.2967/jnumed.108.058768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Lipowska M, He H, Malveaux E, Xu X, Marzilli LG, Taylor AT. J Nucl Med. 2006;47:1032–1040. [PMC free article] [PubMed] [Google Scholar]
  • 5.Maresca KP, Marquis JC, Hillier SM, Lu G, Femia FJ, Zimmerman CN, Eckelman WC, Joyal JL, Babich J. Biocongugate Chem. 2010;21:1032–1042. doi: 10.1021/bc900517x. [DOI] [PubMed] [Google Scholar]
  • 6.Alberto R, N'Dongo HP, Clericuzio M, Bonetti S, Gabano E, Cassino C, Ravera M, Osella D. Inorg Chim Acta. 2009;362:4785–4790. [Google Scholar]
  • 7.Taylor AT, Lipowska M, Marzilli LG. J Nucl Med. 2010;51:391–396. doi: 10.2967/jnumed.109.070813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Taylor AT, Lipowska M, Cai H. J Nucl Med. 2013;54:578–584. doi: 10.2967/jnumed.112.108357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Alberto R, Schibli R, Egli A, Schubiger AP, Abram U, Kaden TA. J Am Chem Soc. 1998;120:7987–7988. doi: 10.1021/ic980112f. [DOI] [PubMed] [Google Scholar]
  • 10.Alberto R, Ortner K, Wheatley N, Schibli R, Schubiger AP. J Am Chem Soc. 2001;123:3135–3136. doi: 10.1021/ja003932b. [DOI] [PubMed] [Google Scholar]
  • 11.Schibli R, Bella RL, Alberto R, Garcia-Garayoa E, Ortner K, Abram U, Schubiger AP. Bioconjugate Chem. 2000;11:345–351. doi: 10.1021/bc990127h. [DOI] [PubMed] [Google Scholar]
  • 12.Abram U, Alberto R. J Braz Chem Soc. 2006;17:1486–1500. [Google Scholar]
  • 13.Duatti A. Technetium-99m radiopharmaceuticals: status and trends. International Atomic Energy Agency; Vienna: 2009. pp. 7–17. [Google Scholar]
  • 14.Alberto R. Eur J Nucl Med Mol Imaging. 2003;30:1299–1302. doi: 10.1007/s00259-003-1292-0. [DOI] [PubMed] [Google Scholar]
  • 15.Banerjee SR, Maresca KP, Francesconi L, Valliant J, Babich JW, Zubieta J. Nucl Med Biol. 2005;32:1–20. doi: 10.1016/j.nucmedbio.2004.09.001. [DOI] [PubMed] [Google Scholar]
  • 16.Bartholomä M, Valliant J, Maresca KP, Babich J, Zubieta J. Chem Commun (Cambridge, UK) 2009:493–512. doi: 10.1039/b814903h. [DOI] [PubMed] [Google Scholar]
  • 17.Alberto R. Technetium-99m radiopharmaceuticals: status and trends. International Atomic Energy Agency; Vienna: 2009. pp. 19–40. [Google Scholar]
  • 18.He HY, Lipowska M, Xu X, Taylor AT, Carlone M, Marzilli LG. Inorg Chem. 2005;44:5437–5446. doi: 10.1021/ic0501869. [DOI] [PubMed] [Google Scholar]
  • 19.Wei L, Babich JW, Ouellette W, Zubieta J. Inorg Chem. 2006;45:3057–3066. doi: 10.1021/ic0517319. [DOI] [PubMed] [Google Scholar]
  • 20.Desbouis D, Struthers H, Spiwok V, Küster T, Schibli R. J Med Chem. 2008;51:6689–6698. doi: 10.1021/jm800530p. [DOI] [PubMed] [Google Scholar]
  • 21.Lipowska M, He H, Xu X, Taylor AT, Marzilli PA, Marzilli LG. Inorg Chem. 2010;49:3141–3151. doi: 10.1021/ic9017568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Alberto R, Schibli R, Schubiger AP, Abram U, Pietzsch HJ, Johannsen B. J Am Chem Soc. 1999;121:6076–6077. [Google Scholar]
  • 23.He HY, Lipowska M, Xu X, Taylor AT, Marzilli LG. Inorg Chem. 2007;46:3385–3394. doi: 10.1021/ic0619299. [DOI] [PubMed] [Google Scholar]
  • 24.He HY, Lipowska M, Christoforou AM, Marzilli LG, Taylor AT. Nucl Med Biol. 2007;34:709–716. doi: 10.1016/j.nucmedbio.2007.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Kyprianidou P, Tsoukalas C, Chiotellis A, Papagiannopolu D, Raptopolou CP, Terzis A, Pelecanou M, Papadopoulos M, Pirmettis I. Inorg Chim Acta. 2011;370:236–242. [Google Scholar]
  • 26.Boros E, Hafeli UO, Patrick BO, Adam MJ, Orvig C. Bioconjugate Chem. 2009;20:1002–1009. doi: 10.1021/bc900022c. [DOI] [PubMed] [Google Scholar]
  • 27.Klenc J, Lipowska M, Marzilli LG, Taylor AT. Eur J Inorg Chem. 2012:4334–4344. doi: 10.1002/ejic.201200599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Schibli R, Schwarzbach R, Alberto R, Ortner K, Schmalle H, Dumas C, Egli A, Schubiger PA. Bioconjugate Chem. 2002;13:750–756. doi: 10.1021/bc015568r. [DOI] [PubMed] [Google Scholar]
  • 29.Xia J, Wang Y, Li G, Yu J, Yin D. J Radioanal Nucl Chem. 2009;279:245–252. [Google Scholar]
  • 30.Banerjee SR, Levadala MK, Lazarova N, Wei L, Valliant JF, Stephenson KA, Babich JW, Maresca KP, Zubieta J. Inorg Chem. 2002;41:6417–6425. doi: 10.1021/ic020476e. [DOI] [PubMed] [Google Scholar]
  • 31.Pietzsch HJ, Gupta A, Reisgys M, Drews A, Seifert S, Syhre R, Spies H, Alberto R, Abram U, Schubiger PA, Johannsen B. Bioconjugate Chem. 2000;11:414–424. doi: 10.1021/bc990162o. [DOI] [PubMed] [Google Scholar]
  • 32.Schubiger PA, Alberto R, Smith A. Bioconjugate Chem. 1996;7:165–179. doi: 10.1021/bc950097s. [DOI] [PubMed] [Google Scholar]
  • 33.Hom R, Katzenellenbogen J. Nucl Med Biol. 1997;24:485–498. doi: 10.1016/s0969-8051(97)00066-8. [DOI] [PubMed] [Google Scholar]
  • 34.Dilworth JR, Parrot S. Chem Soc Rev. 1998;27:43–55. [Google Scholar]
  • 35.Perera T, Abhayawardhana P, Marzilli PA, Fronczek FR, Marzilli LG. Inorg Chem. 2013;52:2412–2421. doi: 10.1021/ic302180t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Storr T, Fisher CL, Mikata Y, Yano S, Adam MJ, Orvig C. Dalton Trans. 2005:654–655. doi: 10.1039/b416040a. [DOI] [PubMed] [Google Scholar]
  • 37.Alberto R. Top Organomet Chem. 2010;32:219–246. [Google Scholar]
  • 38.Christoforou AM, Fronczek FR, Marzilli PA, Marzilli LG. Inorg Chem. 2007;46:6942–6949. doi: 10.1021/ic700594a. [DOI] [PubMed] [Google Scholar]
  • 39.Supuran CT, Casini A, Scozzafava A. Med Res Rev. 2003;23:535–558. doi: 10.1002/med.10047. [DOI] [PubMed] [Google Scholar]
  • 40.Drew J. Science. 2000;287:1960–1964. doi: 10.1126/science.287.5460.1960. [DOI] [PubMed] [Google Scholar]
  • 41.Epstein ME, Amodio-Groton M, Sadick NS. J Am Acad Dermatol. 1997;37:365–381. doi: 10.1016/s0190-9622(97)70135-x. [DOI] [PubMed] [Google Scholar]
  • 42.Gadad AK, Mahajanshetti CS, Nimbalkar S, Raichurkar A. Eur J Med Chem. 2000;35:853–857. doi: 10.1016/s0223-5234(00)00166-5. [DOI] [PubMed] [Google Scholar]
  • 43.Li JJ, Anderson DG, Burton EG, Cogburn JN, Collins JT, Garland DJ, Gregory SA, Huang HC, Isakson PC, Koboldt CM, Logusch EW, Norton MB, Perkins WE, Reinhard EJ, Seibert K, Veenhuizen AW, Zang Y, Reitz DB. J Med Chem. 1995;38:4570–4578. doi: 10.1021/jm00022a023. [DOI] [PubMed] [Google Scholar]
  • 44.Maren TH. Annu Rev Pharmacol Toxicol. 1976;16:309–327. doi: 10.1146/annurev.pa.16.040176.001521. [DOI] [PubMed] [Google Scholar]
  • 45.Renzi G, Scozzafava A, Supuran CT. Bioorg Med Chem Lett. 2000;10:673–676. doi: 10.1016/s0960-894x(00)00075-5. [DOI] [PubMed] [Google Scholar]
  • 46.Christoforou AM, Marzilli PA, Fronczek FR, Marzilli LG. Inorg Chem. 2007;46:11173–11182. doi: 10.1021/ic701576u. [DOI] [PubMed] [Google Scholar]
  • 47.Otwinowski Z, Minor W. Macromolecular Crystallography, Part A, Methods in Enzymology. Vol. 276. New York Academic Press; New York: 1997. pp. 307–326. [DOI] [PubMed] [Google Scholar]
  • 48.Sheldrick G. SADABS. University of Göttingen; Germany: 2004. [Google Scholar]
  • 49.Lane SR, Veerendra B, Rold TL, Sieckman GL, Hoffman TJ, Jurisson SS, Smith CJ. Nucl Med Biol. 2008;35:263–272. doi: 10.1016/j.nucmedbio.2007.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Perera T, Marzilli PA, Fronczek FR, Marzilli LG. Inorg Chem. 2010;49:5560–5572. doi: 10.1021/ic100386c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Perera T, Abhayawardhana P. Unpublished work. Louisiana State University; Baton Rouge, LA: 2013. [Google Scholar]
  • 52.Canovese L, Visentin F, Chessa G, Uguagliati P, Levi C, Dolmella A, Bandoli G. Organometallics. 2006;25:5355–5365. [Google Scholar]
  • 53.Perera T, Fronczek FR, Marzilli PA, Marzilli LG. Inorg Chem. 2010;49:7035–7045. doi: 10.1021/ic100714m. [DOI] [PubMed] [Google Scholar]
  • 54.Veltzé S, Egdal RK, Johansson FB, Bond AD, McKenzie CJ. Dalton Trans. 2009:10495–10504. doi: 10.1039/b912617a. [DOI] [PubMed] [Google Scholar]
  • 55.Lee W, Tseng H, Kuo T. Dalton Trans. 2007:2563–2570. doi: 10.1039/b701367a. [DOI] [PubMed] [Google Scholar]
  • 56.Gunnlaugsson T, Leonard J, Mulready S, Nieuwenhuyzen M. Tetrahedron. 2004;60:105–113. [Google Scholar]
  • 57.Rodriguez V, Gutierrez-Zorrilla JM, Vitoria P, Luque A, Roman P, Martinez-Ripoll M. Inorg Chim Acta. 1999;290:57–63. [Google Scholar]
  • 58.Orpen AG. Chem Soc Rev. 1993;22:191–197. [Google Scholar]
  • 59.Harris SE, Pascual I, Orpen AG. J Chem Soc, Dalton Trans. 2001:2996–3009. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1_si_001
2_si_002

RESOURCES