Next Article in Journal
Redox-Active Profile Characterization of Remirea maritima Extracts and Its Cytotoxic Effect in Mouse Fibroblasts (L929) and Melanoma (B16F10) Cells
Next Article in Special Issue
Isomerization of Internal Alkynes to Iridium(III) Allene Complexes via C–H Bond Activation: Expanded Substrate Scope, and Progress towards a Catalytic Methodology
Previous Article in Journal
Antibacterial Activity and Mechanism of Action of Sulfone Derivatives Containing 1,3,4-Oxadiazole Moieties on Rice Bacterial Leaf Blight
Previous Article in Special Issue
Chlorination of (Phebox)Ir(mesityl)(OAc) by Thionyl Chloride
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iridium-Catalysed ortho-Directed Deuterium Labelling of Aromatic Esters—An Experimental and Theoretical Study on Directing Group Chemoselectivity

Department of Pure & Applied Chemistry, University of Strathclyde, WestCHEM, 295 Cathedral Street, Glasgow G1 1XL, UK
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Current address: School of Chemistry, University of Edinburgh, EastCHEM, David Brewster Road, Edinburgh EH9 3FJ, UK.
Molecules 2015, 20(7), 11676-11698; https://doi.org/10.3390/molecules200711676
Submission received: 2 June 2015 / Revised: 17 June 2015 / Accepted: 19 June 2015 / Published: 25 June 2015
(This article belongs to the Special Issue C-H Bond Activation and Functionalization)

Abstract

:
Herein we report a combined experimental and theoretical study on the deuterium labelling of benzoate ester derivatives, utilizing our developed iridium N-heterocyclic carbene/phosphine catalysts. A range of benzoate esters were screened, including derivatives with electron-donating and -withdrawing groups in the para- position. The substrate scope, in terms of the alkoxy group, was studied and the nature of the catalyst counter-ion was shown to have a profound effect on the efficiency of isotope exchange. Finally, the observed chemoselectivity was rationalized by rate studies and theoretical calculations, and this insight was applied to the selective labelling of benzoate esters bearing a second directing group.

Graphical Abstract

1. Introduction

The ability to incorporate an isotopic label into a biologically-active molecule is of profound importance in the drug discovery process. Such a radioactive “tag” or “label” can be used to provide vital information on a compound’s absorption, distribution, metabolism, excretion, and toxicological (ADMET) properties. As a result of these uses, isotopic labelling is the gold standard method by which early stage drug discovery processes can be optimised.
Research into deuterium (2H or D) and tritium (3H or T) labelling is more substantial than for other isotopes, and has been developed on a number of fronts over the past 60 years [1,2,3,4,5,6,7,8,9,10]. Further to this, key developments in synthetic strategies and analytical techniques for tritium labelling over the past three decades now makes this the preferred technique for many ADMET studies [5]. In one particularly active branch of such labelling research, hydrogen isotope exchange (HIE) is employed to deliver either deuterium or radioactive tritium to pharmaceutical drug candidates in one synthetic step. As well as circumventing the requirement for isotopically-enriched starting materials in preparing tritiated drug candidates [1,5], HIE can also provide analogous deuterated compounds for use as internal standards for mass spectrometry [11,12], for kinetic isotope studies [13,14], and for the alteration of reaction pathways in total synthesis [15].
Over recent years, research in our laboratory has focused on the development of iridium(I) N-heterocyclic carbene (NHC)-ligated precatalysts of the type 1, and their application in HIE via ortho-directed C–H activation protocols (24 via 3, Scheme 1) [6,16,17,18,19,20,21,22]. Despite the growing list of compatible directing groups, we [23] and others [24,25,26,27] have found these developed C–H activation methods less applicable in the labelling of aromatic esters under ambient conditions. Herein, we report the extension of our methodology to encompass the labelling of weakly coordinating benzoate ester derivatives under mild conditions.
Scheme 1. General method for Ir-catalysed ortho-HIE via a C–H activation pathway.
Scheme 1. General method for Ir-catalysed ortho-HIE via a C–H activation pathway.
Molecules 20 11676 g001

2. Results and Discussion

2.1. Catalyst Screening and Comparisons with Crabtree’s Catalyst

Until recently, Crabtree’s catalyst, [(COD)Ir(PCy3)(py)]PF6, 5, was the most widely applied iridium-based HIE catalyst for labelling applications within an industrial setting [24,28]. As such, any studies which evaluate our developed catalysts in the labelling of aromatic esters should also compare them against the ability of 5 to mediate the same catalytic labelling reactions under identical conditions [19,29]. To this end, and to initiate this research programme, the labelling of a series of para-substituted ethyl benzoate derivatives 6 was examined, using our standard labelling protocol (5 mol % [Ir], 1 atm D2, 1 h) with Crabtree’s catalyst 5 (Scheme 2, blue bars) and our developed catalyst systems 1b and 1a (Scheme 2, red and green bars, respectively). With the exception of the p-chloro and p-methoxy esters (6d and 6e), Crabtree’s catalyst 5, proved relatively ineffective in the deuteration of these ester substrates, with incorporations as low as 10% being observed with the electron-withdrawing p-CF3 ester 6c. On assessing the larger and more electron-rich variant of our catalyst series, 1b, a significant improvement in labelling esters 6c (X = CF3) and 6d (X = Cl) was observed, whereas the other esters 6a, 6b and 6e were labelled less efficiently relative to Crabtree’s catalyst. Only on employing our more Lewis acidic catalyst, 1a, did we observe the most efficient and encouraging improvement in ester labelling across all examples tested, with the exception of the parent ethyl benzoate 6a. We hypothesize that the more flexible Lewis acidity of 1a vs. 5 or 1b partially diminishes the importance of the ester coordination event and negative Hammett σp values,[30] and simultaneously enhances the effect of positive σm in relation to a more facile C–H activation event. For example, compare the order of substrate reactivity for catalyst 5 (OMe > Cl > Me > H > CF3) vs. 1a (OMe–Cl > CF3 > Me > H). In the absence of more detailed reaction monitoring, we acknowledge that the observed results cannot be directly related to the kinetically-derived Hammett values. Nonetheless, the hypothesis remains feasible.
Scheme 2. Comparative labelling of ethyl benzoates 6 using catalysts 5, 1b, and 1a.
Scheme 2. Comparative labelling of ethyl benzoates 6 using catalysts 5, 1b, and 1a.
Molecules 20 11676 g002

2.2. Applicable Substrate Scope with Catalyst 1a

2.2.1. Para-Substituted Methyl Benzoates

Using the most efficient of the three catalysts tested, catalyst 1a, we investigated the possibility of labelling methyl esters under the same conditions employed for the ethyl analogues 6 (Table 1). While methyl esters 8d and 8e were labelled with high levels of deuterium incorporation, the other methyl esters in the series proved more capricious. Specifically, substrates 8a, 8b and 8c were repeated multiple times, with individual deuterium incorporations ranging from 27% to 52% (see Experimental Section for full details).
Table 1. Labelling of methyl benzoates 8 using catalyst 1a.
Molecules 20 11676 i001
Table 1. Labelling of methyl benzoates 8 using catalyst 1a.
Molecules 20 11676 i001
EntryXSubstrate%D a
1H8a52
2CH38b42
3CF38c32
4Cl8d93
5OMe8e89
a %D incorporation is the average of two runs and was determined by 1H-NMR spectroscopy.

2.2.2. Scope of O-Alkyl Substitution with Electron-Rich Benzoate Esters

Beyond methyl and ethyl benzoates, we also examined the applicability of larger O-alkyl ester substituents using our labelling method (Table 2). We pursued this line of enquiry for p-methoxybenzoate derivatives only, in order to minimise potential substrate coordination issues associated with the aryl substituent. Whilst the n-propyl, 2,2,2-trifluoroethyl, and tert-butyl benzoate derivatives 10a, 10b and 10c unfortunately proved to be less applicable, iso-propyl and benzyl esters 10d and 10e could be labelled with appreciable levels of deuterium incorporation.
Table 2. Labelling of electron-rich benzoate esters using catalyst 1a.
Molecules 20 11676 i002
Table 2. Labelling of electron-rich benzoate esters using catalyst 1a.
Molecules 20 11676 i002
EntryRSubstrate%D a
1n-Pr10a28
2CH2CF310b8
3t-Bu10c10
4i-Pr10d73
5Bn10e62
a %D incorporation is the average of two runs and was determined by 1H-NMR spectroscopy.

2.3. Reaction Optimisation for Efficient Labelling of Challenging Benzoate Ester Substrates

2.3.1. Temperature Effects

Due to the application of this hydrogen isotope exchange method in related tritiation chemistries [6], significant effort is usually made to maintain ambient reaction conditions during the optimization of ortho-deuteration processes. Having stated this, the use of slightly raised reaction temperatures need not be completely discounted from such investigations. We therefore revisited the labelling of the most challenging methyl and ethyl benzoate derivatives, using a moderately increased reaction temperature of 40 °C (Scheme 3). Pleasingly, dramatic improvements in deuterium incorporation were observed across all substrates examined, 6ac and 8ac, whilst the short reaction times were maintained (see Experimental Section 3.4).
Scheme 3. Temperature effects on deuterium labelling of previously problematic benzoate esters.
Scheme 3. Temperature effects on deuterium labelling of previously problematic benzoate esters.
Molecules 20 11676 g003

2.3.2. Catalyst Counterion Effects

We recently reported the improved activity and broad-spectrum solubility resulting from replacement of the PF6 counterion in catalyst 1a with tetrakis[bis-3,5-trifluoromethylphenyl]borate, BArF, to give complex 1d [18]. Applying improved catalyst 1d to the labelling of larger ester derivatives 10ae under otherwise identical conditions, significant and more usable levels of deuterium incorporation were observed across all examples (Scheme 4). Importantly, this counter-ion switch demonstrates an alternative means by which ester labelling efficiency can be improved, should ambient temperature conditions be required.
Scheme 4. Exploiting anion effects for improved ester labelling.
Scheme 4. Exploiting anion effects for improved ester labelling.
Molecules 20 11676 g004

2.4. Exploring Chemoselectivity Issues in Ester Labelling

2.4.1. Experimental Observations

From the outset of our studies, it was clear that the main challenge in labelling aromatic esters would be the weak coordinating ability of this functional group. To understand this issue in more detail, we conducted a series of intramolecular competition studies where multiple potential directing groups can compete for coordination (and subsequent C–H activation) at the iridium centre. To this end, we first investigated the labelling of ethyl p-nitrobenzoate, 12, under the optimised ambient reaction conditions. Interestingly, we observed an approximate 1.4:1 selectivity for labelling via the nitro rather than the ester directing group (Scheme 5). This chemoselectivity was eroded entirely on changing the ester to a tertiary amide in 13 (1:1 nitro:amide), and reversed by replacing the ester with a ketone in 14 (3:1 in favour of the ketone).
Scheme 5. Variation in labelling regioselectivity based on directing group chemoselectivity.
Scheme 5. Variation in labelling regioselectivity based on directing group chemoselectivity.
Molecules 20 11676 g005
Focusing on the extreme substrate cases, with substrates 12 and 14, 1 mol % of catalyst 1a was employed to allow reaction rates and labelling selectivities to be monitored over time (Scheme 6). In the case of substrate 12, and the labelling of the positions ortho- to the ester vs. the nitro, the difference in reaction rate, and thus product selectivity, remains largely constant throughout the course of the reaction. Conversely, with substrate 14, the relative rate of labelling ortho- to ketone vs. nitro is higher at lower conversions, and decreases rapidly over time.
Scheme 6. Rate and product selectivity studies for 12 (top) and 14 (bottom).
Scheme 6. Rate and product selectivity studies for 12 (top) and 14 (bottom).
Molecules 20 11676 g006

2.4.2. Theoretical Analysis

In previous C–H activation studies, we rationalised observed directing group chemoselectivity using density functional theory (DFT) calculations (see Supplementary Materials for details) to model the relative energies of the binding conformers and subsequent C–H activation pathways [20]. We have now extended this approach to the analysis of the labelling reactions of 12 and 14 (Scheme 7). In agreement with previous findings, we qualitatively predicted that the most stable binding isomer should also be the most reactive. If Curtin-Hammett kinetics are assumed [31], the calculated ΔΔG (and thus product selectivity) from equilibrium and activation parameters is predicted to be higher for the ketone in 14 vs. the nitro group in 12 (5.5 vs. 1.4 kcal/mol, respectively). However, the current model does not account for the barrier to interconversion of binding conformations (ketone to nitro, ester to nitro, and vice versa for each case). Considering these points in the context of the experimentally-determined product selectivity vs. time (vide supra), only substrate 12 (showing little variation in selectivity over time) appears to show rapid equilibrium between the binding isomers. Conversely, the labelling of 14 via the ketone may be interpreted as being faster than the rate of interconversion between binding conformers as well as possessing a lower barrier to C–H activation.
Scheme 7. Density functional theory (DFT) analyses on the C–H activation step in labelling substrates 12 and 14 with 1a.
Scheme 7. Density functional theory (DFT) analyses on the C–H activation step in labelling substrates 12 and 14 with 1a.
Molecules 20 11676 g007

2.4.3. Practical Exploitation of Directing Group Chemoselectivity

The fundamental analysis of intramolecular directing group chemoselectivity served to show that observed labelling patterns are, in part, dependent on the relative catalyst binding affinities of each directing group. With this new understanding in hand, we questioned if it would be possible to control the direction of labelling using the inherent reactivity of a given multi-functional substrate. Pleasingly, using substrate 14 as a proof-of-concept substrate, minimal optimisation was required to show that judicious choice of catalyst loading and reaction temperature allowed control of the labelling pattern (Scheme 8). Specifically, labelling ortho- to the ketone group could be achieved with a 5 mol % catalyst loading of 1a at room temperature to give 18, whereas the globally-labelled product 19 could be obtained by employing 5 mol % of 1a at 40 °C. In turn, the previously elusive nitro-selective product, 20, was accessed by a retro-labelling strategy (employing H2 in place of D2) conducted on the globally-labelled product 19.
Scheme 8. Condition-dependent flexible access to alternatively deuterated forms of 18.
Scheme 8. Condition-dependent flexible access to alternatively deuterated forms of 18.
Molecules 20 11676 g008

3. Experimental Section

3.1. General Considerations

All reagents were obtained from commercial suppliers (Sigma-Aldrich, Gillingham, UK; Alfa Aesar, Heysham, UK; or Strem, Newton, UK) and used without further purification, unless otherwise stated. Dichloromethane was obtained from a PureSolv SPS-400-5 Solvent Purification System (Innovative Technology, Inc., Galway, Ireland), and deoxygenated by bubbling argon through for a minimum of ten minutes. Thin layer chromatography was carried out using Camlab silica plates coated with fluorescent indicator UV254 (Sigma-Aldrich), and were visualized using a Mineralight UVGL-25 lamp (Fisher Scientific UK Ltd., Loughborough, UK) or developed using vanillin solution. Catalysts 1a [20], 1b [16], and 1d [18] were prepared according to literature procedures. Esters 6c [32], 10a [33], 10b [34], 10c [35], 10d [36], 10e [37], and 13 [38] were prepared according to the corresponding literature procedures. 1H-NMR spectra were recorded on a Bruker DPX 400 spectrometer (Bruker, Coventry, UK) at 400 MHz. Chemical shifts are reported in ppm and coupling constants are reported in Hz. All coupling constants are 3JH–H unless otherwise stated.

3.2. General Procedures

3.2.1. Deuteration of Substrates Using Iridium(I) Complexes 1a, 1b, 1d and 5

A three-necked round bottom flask was fitted with two stopcock side arms and a rubber septum, and then flame-dried. To this flask was added the iridium(I) complex and substrate. The solvent, DCM (2.5 mL, unless stated otherwise), was added, rinsing the inner walls of the flask, and the rubber septum was replaced with a greased glass stopper. The solution was placed under an atmosphere of Ar and stirred whilst being cooled to −78 °C in a dry ice/acetone bath. The flask was evacuated then refilled with argon and this process repeated. Upon a third evacuation, an atmosphere of deuterium gas was introduced to the flask. After sealing the flask, the cold bath was removed and the flask heated in an oil bath to the desired temperature. The reaction mixture was stirred for the allotted reaction time before removing the deuterium atmosphere and replacing with air. The resulting solution was washed with DCM and transferred to a single-necked flask before removing the solvent under reduced pressure. The catalyst was triturated from the remaining residue by addition of diethyl ether (3 × 5 mL). The solution was filtered through a short plug of silica before the solvent was removed in vacuo to deliver the crude product for analysis of the deuterium incorporation.
The level of deuterium incorporation in the substrate was determined by 1H-NMR spectroscopy. The integrals were calibrated against a peak corresponding to a position not expected to be labelled. Equation (1) was then used to calculate the extent of labelling:
%   Deuteration = 100 [ ( residual integral number of labelling sites ) ×   100 ]

3.2.2. Deuteration of Substrates for Rate Studies

A three-necked round bottom flask fitted with one stopcock side arm and two rubber septa was flame-dried. To this flask was added the iridium(I) complex, and substrate. The solvent, DCM (25 mL), was added, rinsing the inner walls of the flask, and one rubber septum was replaced with a greased glass stopper. The solution was placed under an atmosphere of argon and stirred whilst being cooled to −78 °C in a dry ice/acetone bath. The flask was evacuated then refilled with argon and this process repeated. Upon a third evacuation, an atmosphere of deuterium gas was introduced to the flask via a balloon. The balloon was left in place for the duration of the reaction to ensure a continuous supply of deuterium. The cold bath was removed and the flask heated in an oil bath to the desired temperature. The reaction mixture was then stirred for the allotted reaction time. An aliquot (1 mL) of the reaction mixture was removed at intervals throughout the reaction (10, 20, 30, 40, 50, 60 min, 2 h, and 18 h). Each aliquot was transferred to a vial containing diethyl ether. After removal of solvent in vacuo, the residue was analysed by 1H NMR spectroscopy. The integrals were calibrated against a peak corresponding to a position not expected to be labelled. The extent of labelling was determined using Equation (1).

3.3. Labelling Studies

3.3.1. Deuteration of Ethyl benzoate 6a with Complex 5

Molecules 20 11676 i003
1H-NMR (400 MHz, CDCl3): δ 8.05 (d, J = 7.1 Hz, 2H, CH3), 7.54 (t, J = 7.3 Hz, 1H, CH1), 7.43 (t, J = 7.9 Hz, 2H, CH2), 4.38 (q, J = 7.2 Hz, 2H, CH24), 1.39 (t, J = 7.1 Hz, 3H, CH35). Incorporation expected at δ 8.05. Determined against integral at δ 1.39. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 6a, 0.032 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 17%.
Run 2: (a) 6a, 0.032 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 22%

3.3.2. Deuteration of Ethyl 4-methylbenzoate 6b with Complex 5

Molecules 20 11676 i004
1H-NMR (400 MHz, CDCl3): δ 7.91 (d, J = 8.0 Hz, 2H, CH3), 7.21 (d, J = 7.9 Hz, 2H, CH2), 4.34 (q, J = 7.1 Hz, 2H, CH24), 2.38 (s, 3H, CH31) 1.36 (t, J = 7.1 Hz, 3H, CH35). Incorporation expected at δ 7.91. Determined against integral at δ 2.38. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 6b, 0.035 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 36%.
Run 2: (a) 6b, 0.035 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 41%.

3.3.3. Deuteration of Ethyl 4-(trifluoromethyl)benzoate 6c with Complex 5

Molecules 20 11676 i005
1H-NMR (400 MHz, CDCl3): δ 8.12 (d, J = 8.3 Hz, 2H, CH2) 7.67(d, J = 8.3 Hz, 2H, CH1) 4.38 (q, J = 7.2 Hz, 2H, CH23), 1.38 (t, J = 7.2 Hz, 3H, CH34). Incorporation expected at δ 8.12. Determined against integral at δ 1.38. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 6c, 0.044 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 8%
Run 2: (a) 6c, 0.044 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 11%

3.3.4. Deuteration of Ethyl 4-chlorobenzoate 6d with Complex 5

Molecules 20 11676 i006
1H-NMR (400 MHz, CDCl3): δ 7.95 (d, J = 8.8 Hz, 2H, CH2) 7.38 (d, J = 8.8 Hz, 2H, CH1) 4.34 (q, J = 7.3 Hz, 2H, CH23), 1.36 (t, J = 7.3 Hz, CH34). Incorporation expected at δ 7.95. Determined against integral at δ 1.36. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 6d, 0.039 g, 0.215 mmol; (b) 5, 8.04 mg; 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 50%
Run 2: (a) 6d, 0.039 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 59%

3.3.5. Deuteration of Ethyl 4-methoxybenzoate 6e with Complex 5

Molecules 20 11676 i007
1H-NMR (400 MHz, CDCl3): δ 7.98 (d, J = 9.1 Hz, 2H, CH3), 6.89 (d, J = 9.0 Hz, 2H, CH2), 4.30 (q, J = 7.0 Hz, 2H, CH24), 3.83 (s, 3H, CH31), 1.35 (t, 7.1Hz, 3H, CH35). Incorporation expected at δ 7.98. Determined against integral at δ 3.83. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 6e, 0.038 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 89%
Run 2: (a) 6e, 0.038 g, 0.215 mmol; (b) 5, 8.04 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 71%

3.3.6. Deuteration of Esters 6ae Using Catalysts 1b and 1a

For the results relating to catalysts 1b and 1a in Scheme 2, please refer to the spectroscopic data from Section 3.3.1, Section 3.3.2, Section 3.3.3, Section 3.3.4 and Section 3.3.5 for the analysis of the deuterated esters 6ae. As catalyst type and amount used are the only variables changed, General Procedure A was followed with the results tabulated in Table 3.

3.3.7. Deuteration of Methyl Benzoate 8a with Complex 1a

Molecules 20 11676 i008
1H-NMR (400 MHz, CDCl3): δ 8.04 (d, J = 7.0 Hz, 2H, CH3), 7.56 (t, J = 7.4 Hz, 1H, CH1), 7.44 (t, J = 7.9 Hz, 2H, CH2), 3.92 (s, 3H, CH34). Incorporation expected at δ 8.04. Determined against integral at δ 3.92. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 8a, 0.032 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 52%
Run 2: (a) 8a, 0.032 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 51%
Table 3. Deuteration of Esters 6ae.
Table 3. Deuteration of Esters 6ae.
EntrySubstrateCatalyst%D (run 1)%D (run 2)
16a1b (10.5 mg)1010
26b1523
36c6250
46d8593
56e6559
66a1a (10.1 mg)640
76b5383
86c7479
96d9595
106e9696

3.3.8. Deuteration of Methyl 4-methylbenzoate 8b with Complex 1a

Molecules 20 11676 i009
1H-NMR (400 MHz, CDCl3): δ 7.99 (d, J = 8.0 Hz, 2H, CH3), 7.21 (d, J = 8.0 Hz, 2H, CH2), 3.89 (s, 3H, CH34), 2.39 (s, 3H, CH31). Incorporation expected at δ 7.99. Determined against integral at δ 2.39. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 8b, 0.035 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h, d) 25 °C; and (e) 35%
Run 2: (a) 8b, 0.035 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 48%

3.3.9. Deuteration of Methyl 4-(trifluoromethyl)benzoate 8c with Complex 1a

Molecules 20 11676 i010
1H-NMR (400 MHz, CDCl3): δ 8.14 (d, J = 8.8 Hz, 2H, CH2), 7.70 (d, J = 8.8 Hz, 2H, CH1), 3.94 (s, 3H, CH33). Incorporation expected at δ 8.14. Determined against integral at δ 3.94. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 8c, 0.044 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 27%
Run 2: (a) 8c, 0.044 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 36%

3.3.10. Deuteration of Methyl 4-chlorobenzoate 8d with Complex 1a

Molecules 20 11676 i011
1H-NMR (400 MHz, CDCl3): δ 7.93 (d, J = 8.9 Hz, 2H, CH2), 7.36 (d, J = 8.9 Hz, 2H, CH1), 3.90 (s, 3H, CH33). Incorporation expected at δ 7.93. Determined against integral at δ 3.90. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 8d, 0.039 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 91%
Run 2: (a) 8d, 0.039 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 94%

3.3.11. Deuteration of Methyl 4-methoxybenzoate 8e with Complex 1a

Molecules 20 11676 i012
1H-NMR (400 MHz, CDCl3): δ 7.96 (d, J = 9.0 Hz, 2H, CH3), 6.89 (d, J = 9.0 Hz, 2H, CH2), 3.85 (s, 3H, CH34), 3.82 (s, 3H, CH31). Incorporation expected at δ 7.96. Determined against integral at δ 3.82. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 8e, 0.038 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 87%
Run 2: (a) 8e, 0.038 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 90%

3.3.12. Deuteration of n-Propyl 4-methoxybenzoate 10a with Complex 1a

Molecules 20 11676 i013
1H-NMR (400 MHz, CDCl3): δ 8.00 (d, J = 9.2 Hz, 2H, CH3), 6.92 (d, J = 9.2 Hz, 2H, CH2), 4.25 (t, J = 6.8 Hz, 2H, CH4), 3.86 (s, 3H, CH31), 1.82–1.73 (m, 2H, CH25), 1.02 (t, J = 7.3 Hz, 3H, CH36). Incorporation expected at δ 8.00. Determined against integral at δ 3.86. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 10a, 0.042 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 28%
Run 2: (a) 10a, 0.038 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 27%

3.3.13. Deuteration of 2,2,2-Trifluoroethyl 4-methoxybenzoate 10b with Complex 1a

Molecules 20 11676 i014
1H-NMR (400 MHz, CDCl3): δ 8.06 (d, J = 9.3 Hz, 2H, CH3), 6.97 (d, J = 9.3 Hz, 2H, CH2), 4.69 (q, 3JH-F = 8.5 Hz, 2H, CH24), 3.90 (s, 3H, CH31). Incorporation expected at δ 8.06. Determined against integral at δ 3.90. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 10b, 0.050 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 9%
Run 2: (a) 10b, 0.050 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 7%

3.3.14. Deuteration of t-Butyl 4-methoxybenzoate 10c with Complex 1a

Molecules 20 11676 i015
1H-NMR (400 MHz, CDCl3): δ 7.94 (d, J = 8.7 Hz, 2H, CH3), 6.89 (d, J = 8.9 Hz, 2H, CH2), 3.85 (s, 3H, CH31), 1.58 (s, 9H, CH34). Incorporation expected at δ 7.94. Determined against integral at δ 3.85. Following General Procedure A, results are reported as (a) amount of substrate, (b) amount of catalyst, (c) reaction time, (d) reaction temperature, and (e) level of incorporation.
Run 1: (a) 10c, 0.045 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 10%
Run 2: (a) 10c, 0.045 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 9%

3.3.15. Deuteration of Iso-propyl 4-methoxybenzoate 10d with Complex 1a

Molecules 20 11676 i016
1H-NMR (400 MHz, CDCl3): δ 7.97 (d, J = 9.0 Hz, 2H, CH3), 6.89 (d, J = 9.0 Hz, 2H, CH2), 5.20 (septet, J = 6.3 Hz, 1H, CH4), 3.84 (s, 3H, CH31), 1.33 (d, J = 6.5 Hz, 6H, CH35). Incorporation expected at δ 7.97. Determined against integral at δ 3.84. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 10d, 0.042 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 79%
Run 2: (a) 10d, 0.042 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 66%

3.3.16. Deuteration of Benzyl 4-methoxybenzoate 10e with Complex 1a

Molecules 20 11676 i017
1H-NMR (400 MHz, CDCl3): δ 8.04 (d, J = 9.0 Hz, 2H, CH3), 7.45–7.43 (m, 2H, CH5), 7.40–7.37 (m, 2H, CH6), 7.35–7.32 (m, 1H, CH7), 6.92 (d, J = 9.0 Hz, 2H, CH2), 5.34 (s, 2H, CH24), 3.90 (s, 3H, CH31). Incorporation expected at δ 8.04. Determined against integral at δ 3.90. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
Run 1: (a) 10e, 0.052 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 73%
Run 2: (a) 10e, 0.052 g, 0.215 mmol; (b) 1a, 10.9 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 51%

3.4. Temperature Effects

For the results relating to Scheme 3, readers are directed to the spectroscopic data in the relevant parts of Section 3.3 for the analysis of the corresponding deuterated esters 6ac and 8ac. As catalyst type and amount used are the only variables changed, the remaining results are tabulated below. In all cases, 0.215 mmol of substrate was employed with 10.1 mg of catalyst 1a (0.01 mmol, 5 mol %) and the reactions run at 40 °C, otherwise following General Procedure A, and the results shown below in Table 4.
Table 4. Deuteration of Esters 6ac and 8ac at 40 °C.
Table 4. Deuteration of Esters 6ac and 8ac at 40 °C.
EntrySubstrate%D (run 1)%D (run 2)
16a8586
26b8991
36c9596
48a7576
58b9695
68c9293

3.5. Catalyst Counterion Effects

3.5.1. Deuteration of n-Propyl 4-methoxybenzoate 10a with Complex 1d

Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
(a) 10a, 0.042 g, 0.215 mmol; (b) 1d, 18.6 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 89%

3.5.2. Deuteration of 2,2,2-Trifluoroethyl 4-methoxybenzoate 10b with Complex 1d

Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
(a) 10b, 0.050 g, 0.215 mmol; (b) 1d, 18.6 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 24%

3.5.3. Deuteration of t-Butyl 4-methoxybenzoate 10c with Complex 1d

Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
(a) 10c, 0.045 g, 0.215 mmol; (b) 1d, 18.6 mg, 0.01 mmol; (c) 1 h, (d) 25 °C; and (e) 41%

3.5.4. Deuteration of iso-Propyl 4-methoxybenzoate 10d with Complex 1d

Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
(a) 10d, 0.042 g 0.215 mmol; (b) 1d, 18.6 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 94%.

3.5.5. Deuteration of Benzyl 4-methoxybenzoate 10e with Complex 1d

Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation.
(a) 10e, 0.052 g, 0.215 mmol; (b) 1d, 18.6 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 89%.

3.6. Chemoselectivity Studies

3.6.1. Deuteration of Ethyl 4-nitrobenzoate 12 with Complex 1a

Molecules 20 11676 i018
1H-NMR (400 MHz, CDCl3): δ 8.27 (d, J = 9.0 Hz, 2H, CH1), 8.19 (d, J = 8.3 Hz, 2H, CH2), 4.42 (q, J = 7.5 Hz, 2H, CH23), 1.41 (t, J = 7.5 Hz, 3H, CH34). Incorporation expected at δ 8.27 and 8.19. Determined against integral at δ 1.41. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation at position 1 and position 2.
Run 1: (a) 12, 0.042 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 1: 72%, 2: 49%
Run 2: (a) 12, 0.042 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 1: 72%, 2: 52%

3.6.2. Deuteration of N,N-Diethyl 4-nitrobenzamide 13 with Complex 1a

Molecules 20 11676 i019
1H-NMR (400 MHz, CDCl3): δ 8.28 (d, J = 8.9 Hz, 2H, CH1), 7.57 (d, J = 8.9 Hz, 2H, CH2), 3.17–3.53 (m, 4H, CH23), 1.25 (br s, 6H, CH34). Incorporation expected at δ 8.28 and 7.57. Determined against integral at δ 1.25. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation at position 1 and position 2.
Run 1: (a) 13, 0.047 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 1: 89%, 2: 93%.
Run 2: (a) 13, 0.047 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h, d) 25 °C, and (e) 1: 94%, 2: 93%.

3.6.3. Deuteration of 4-Nitroacetophenone 14 with Complex 1a

Molecules 20 11676 i020
1H-NMR (400 MHz, CDCl3): δ 8.32 (d, J = 8.9 Hz, 2H, CH1), 8.11 (d, J = 8.9 Hz, 2H, CH2), 2.69 (s, 1H, CH33). Incorporation expected at δ 8.32 and 8.11. Determined against integral at δ 2.69. Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation at position 1 and position 2.
Run 1: (a) 14, 0.0467 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 1: 28%, 2: 97%
Run 2: (a) 14, 0.0467 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 25 °C; and (e) 1: 33%, 2: 85%

3.7. Rate Studies

3.7.1. Deuteration of Ethyl 4-nitrobenzoate 12 with Complex 1a

Molecules 20 11676 i021
1H-NMR (400 MHz, CDCl3): δ 8.27 (d, J = 9.0 Hz, 2H, CH1), 8.19 (d, J = 8.3 Hz, 2H, CH2). 4.42 (q, J = 7.5 Hz, 2H, CH23), 1.41 (t, J = 7.5 Hz, 3H, CH34). Incorporation expected at 1: δ 8.27 and 2: δ 8.19. Determined against integral at δ 1.41. Following General Procedure B, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation at position 1 and 2 at each time interval.
Labelling at Position 1: (a) 12, 0.399 g, 2.15 mmol; (b) 1a, 2.17 mg, 0.0215 mmol; (c) 18 h; (d) 25 °C; and (e) 21% (10 min), 37% (20 min), 47% (30 min), 56% (40 min), 61% (50 min), 64% (1 h), 73% (2 h), and 76% (18 h).
Labelling at Position 2: (a) 12, 0.399 g, 2.15 mmol; (b) 1a, 2.17 mg, 0.0215 mmol; (c) 18 h; (d) 25 °C; and (e) 2: 14% (10 min), 24% (20 min), 33% (30 min), 37% (40 min), 41% (50 min), 43% (1 h), 52% (2 h), and 56% (18 h).

3.7.2. Deuteration of 4-Nitroacetophenone 14 with Complex 1a

Molecules 20 11676 i022
1H-NMR (400 MHz, CDCl3): δ 8.32 (d, J = 8.9 Hz, 2H, CH1), 8.11 (d, J = 8.9 Hz, 2H, CH2), 2.69 (s, 1H, CH33). Incorporation expected at δ 8.32 and 8.11. Determined against integral at δ 2.69. Following General Procedure B, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation at position 1 and 2 at each time interval.
Labelling at Position 1: (a) 14, 0.355 g, 2.15 mmol; (b) 1a, 2.12 mg, 0.021 mmol; (c) 18 h; (d) 25 °C; and (e) 8% (10 min), 13% (20 min), 19% (30 min), 26%(40 min), 29% (50 min), 33% (1 h), 65% (2 h), 69% (18 h).
Labelling at Position 2: (a) 14, 0.0467 g, 2.15 mmol; (b) 1a, 2.12 mg, 0.01 mmol; (c) 18 h; (d) 25 °C; and (e) 54% (10 min), 75% (20 min), 82% (30 min), 84% (40 min), 85% (50 min), 85% (1 h), 91% (2 h), 95% (18 h).

3.8. Practical Exploitation of Directing Group Chemoselectivity

Molecules 20 11676 i023

3.8.1. Deuteration of 4-Nitroacetophenone 14 with Complex 1a

1H-NMR (400 MHz, CDCl3): δ 8.32 (d, J = 8.9 Hz, 2H, CH1), 8.11 (d, J = 8.9 Hz, 2H, CH2), 2.69 (s, 1H, CH33). Incorporation expected at δ 8.32 and 8.11. Determined against integral at δ 2.69. See the details within Section 3.6.3.

3.8.2. Deuteration of 4-Nitroacetophenone 14 with Complex 1a

Following General Procedure A, results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time, (d) reaction temperature, and (e) level of incorporation at position 1 and position 2.
(a) 14, 0.035 g, 0.215 mmol; (b) 1a, 10.1 mg, 0.01 mmol; (c) 1 h; (d) 40 °C; and (e) 1: 97%, 2: 97%

3.8.3. Hydrogenation of 4-Nitroacetophenone-d4 19 with Complex 1a

Following General Procedure A (deuterium gas was replaced with hydrogen gas), results are reported as (a) amount of substrate; (b) amount of catalyst; (c) reaction time; (d) reaction temperature; and (e) level of incorporation of deuterium at position 1 and position 2.
(a) 19, 0.0363 g, 0.215 mmol; (b) 1a, 2.1 mg, 0.0021 mmol; (c) 1 h; (d) 25 °C; and (e) 1: 90%, 2: 12%.

4. Computational Details

DFT [39] was employed to calculate the gas-phase electronic structures and energies for all species involved in H/D exchange reactions. All structures have been optimised with the hybrid meta-GGA exchange correlation functional M06 [40]. The M06 density functional was used in conjunction with the 6-31G(d) basis set for main group non-metal atoms and the Stuttgart RSC [41] effective core potential along with the associated basis set for Ir. The participating transition states (TS) are located at the same level of theory. Harmonic vibrational frequencies are calculated at the same level of theory to characterize respective minima (reactants, intermediates, and products with no imaginary frequency) and first order saddle points (TSs with one imaginary frequency). The validity of using the 6-31G(d) basis set has previously been checked by comparative single point energy calculations employing the def2-TZVP basis set for all atoms on similar H/D exchange systems [20]. All calculations using the M06 functional have been performed using Gaussian 09 quantum chemistry program package (version A.02). Calculations were first carried out in the gas phase before reoptimising each structure at the same level of theory, implementing the Polarizable Continuum Model (PCM) for DCM as the solvent [42]. All coordinates provided are listed in Cartesian format, with charge and multiplicity of each system given at the top of the coordinate list (i.e., 0 1 = neutral singlet; 1 1 = 1 + charged singlet).

5. Conclusions

In summary, we have divulged novel iridium-catalysed methods for the ortho-deuteration of benzoate esters by the application of complexes emerging from our laboratory, possessing a bulky NHC/phosphine combination. Inherent variability in reproducibly labelling ester substrates to useable levels of d-incorporation was solved by two methods: (i) a mild increase in reaction temperature; and (ii) a switch in the catalyst anion from PF6 to BArF; this delivered good to excellent levels of deuterium incorporation adjacent to ester directing groups. Kinetic studies on intramolecular directing group chemoselectivity revealed that selectivity vs. time is substrate dependent, showing the possibility that different levels of binding conformer equilibria are possible. Supporting DFT analyses of the systems studied experimentally support previous findings that suggested the most stable binding conformer is also the most reactive. We have demonstrated that such knowledge can be exploited experimentally and, as such, we have shown that different modes of regioselective labelling is possible in a multifunctional substrate by simple variation of the reaction conditions. Overall, we believe that these methods further enhance the applicable substrate scope and wider utility of the iridium complexes at the centre of this study.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/20/07/11676/s1.

Acknowledgments

The authors would like to thank the Carnegie Trust for the Universities of Scotland (M.R.) for funding. Mass spectrometry data were acquired at the EPSRC UK National Mass Spectrometry Facility at Swansea University.

Author Contributions

The project was devised by W.J.K. and M.R. Experimental results were obtained by J.D. and T.J.D.M. Computational analysis was conducted by M.R. with consultation from T.T. The manuscript was prepared by D.M.L., M.R., and W.J.K.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Isin, E.M.; Elmore, C.S.; Nilsson, G.N.; Thompson, R.A.; Weidolf, L. Use of radiolabeled compounds in drug metabolism and pharmacokinetic studies. Chem. Res. Toxicol. 2012, 25, 532–542. [Google Scholar] [CrossRef] [PubMed]
  2. Allen, P.H.; Hickey, M.J.; Kingston, L.P.; Wilkinson, D.J. Metal-catalysed isotopic exchange labelling: 30 years of experience in pharmaceutical R&D. J. Label. Compd. Radiopharm. 2010, 53, 731–738. [Google Scholar]
  3. Hesk, D.; Lavey, C.F.; McNamara, P. Tritium labelling of pharmaceuticals by metal-catalysed exchange methods. J. Label. Compd. Radiopharm. 2010, 53, 722–730. [Google Scholar] [CrossRef]
  4. Heys, J.R. Organoiridium complexes for hydrogen isotope exchange labeling. J. Label. Compd. Radiopharm. 2007, 50, 770–778. [Google Scholar] [CrossRef]
  5. Lockley, W.J.S.; McEwen, A.; Cooke, R. Tritium: A coming of age for drug discovery and development ADME studies. J. Label. Compd. Radiopharm. 2012, 55, 235–257. [Google Scholar] [CrossRef]
  6. Nilsson, G.N.; Kerr, W.J. The development and use of novel iridium complexes as catalysts for ortho-directed hydrogen isotope exchange reactions. J. Label. Compd. Radiopharm. 2010, 53, 662–667. [Google Scholar] [CrossRef] [Green Version]
  7. Sawama, Y.; Monguchi, Y.; Sajiki, H. Efficient H-D Exchange Reactions Using Heterogeneous Platinum-Group Metal on Carbon-H2-D2O System. Synlett 2012, 23, 959–972. [Google Scholar] [CrossRef]
  8. Atzrodt, J.; Derdau, V.; Fey, T.; Zimmermann, J. The renaissance of H/D exchange. Angew. Chem. Int. Ed. 2007, 46, 7744–7765. [Google Scholar] [CrossRef] [PubMed]
  9. Lockley, W.J.S. 30 Years with ortho-directed hydrogen isotope exchange labelling. J. Label. Compd. Radiopharm. 2007, 50, 779–788. [Google Scholar] [CrossRef]
  10. Hesk, D.; McNamara, P. Synthesis of isotopically labelled compounds at Schering-Plough, an historical perspective. J. Label. Compd. Radiopharm. 2007, 50, 875–887. [Google Scholar] [CrossRef]
  11. Lin, D.; Chang, W. Chemical derivatization and the selection of deuterated internal standard for quantitative determination—Methamphetamine example. J. Anal. Toxicol. 2000, 24, 275–280. [Google Scholar] [CrossRef] [PubMed]
  12. Atzrodt, J.; Derdau, V. Pd- and Pt-catalyzed H/D exchange methods and their application for internal MS standard preparation from a Sanofi-Aventis perspective. J. Label. Compd. Radiopharm. 2010, 53, 674–685. [Google Scholar] [CrossRef]
  13. Parkin, G. Applications of deuterium isotope effects for probing aspects of reactions involving oxidative addition and reductive elimination of H–H and C–H bonds. J. Label. Compd. Radiopharm. 2007, 50, 1088–1114. [Google Scholar] [CrossRef]
  14. Heinekey, D.M. Transition metal dihydrogen complexes: Isotope effects on reactivity and structure. J. Label. Compd. Radiopharm. 2007, 50, 1063–1071. [Google Scholar] [CrossRef]
  15. Quasdorf, K.W.; Huters, A.D.; Lodewyk, M.W.; Tantillo, D.J.; Garg, N.K. Total synthesis of oxidized welwitindolinones and (−)-N-methylwelwitindolinone C isonitrile. J. Am. Chem. Soc. 2012, 134, 1396–1399. [Google Scholar] [CrossRef] [PubMed]
  16. Brown, J.A.; Irvine, S.; Kennedy, A.R.; Kerr, W.J.; Andersson, S.; Nilsson, G.N. Highly active iridium(I) complexes for catalytic hydrogen isotope exchange. Chem. Commun. 2008, 1115–1118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Cochrane, A.R.; Idziak, C.; Kerr, W.J.; Mondal, B.; Paterson, L.C.; Tuttle, T.; Andersson, S.; Nilsson, G.N. Practically convenient and industrially-aligned methods for iridium-catalysed hydrogen isotope exchange processes. Org. Biomol. Chem. 2014, 12, 3598–3603. [Google Scholar] [CrossRef] [PubMed]
  18. Kennedy, A.R.; Kerr, W.J.; Moir, R.; Reid, M. Anion effects to deliver enhanced iridium catalysts for hydrogen isotope exchange processes. Org. Biomol. Chem. 2014, 12, 7927–7931. [Google Scholar] [CrossRef] [PubMed]
  19. Atzrodt, J.; Derdau, V.; Kerr, W.J.; Reid, M.; Rojahn, P.; Weck, R. Expanded applicability of iridium(I) NHC/phosphine catalysts in hydrogen isotope exchange processes with pharmaceutically-relevant heterocycles. Tetrahedron 2015, 71, 1924–1929. [Google Scholar] [CrossRef]
  20. Kerr, W.J.; Reid, M.; Tuttle, T. Iridium-Catalyzed C–H Activation and Deuteration of Primary Sulfonamides: An Experimental and Computational Study. ACS Catal. 2015, 5, 402–410. [Google Scholar] [CrossRef]
  21. Cochrane, A.R.; Irvine, S.; Kerr, W.J.; Reid, M.; Andersson, S.; Nilsson, G.N. Application of Neutral Iridium(I) N-Heterocyclic Carbene Complexes in ortho-Directed Hydrogen Isotope Exchange. J. Label. Compd. Radiopharm. 2013, 56, 451–454. [Google Scholar] [CrossRef] [PubMed]
  22. Kerr, W.J.; Mudd, R.J.; Paterson, L.C.; Brown, J.A. Iridium(I)-Catalyzed Regioselective C–H Activation and Hydrogen-Isotope Exchange of Non-aromatic Unsaturated Functionality. Chem. Eur. J. 2014, 20, 14604–14607. [Google Scholar] [CrossRef] [PubMed]
  23. Cross, P.W.C.; Ellames, G.J.; Gibson, J.S.; Herbert, J.M.; Kerr, W.J.; McNeill, A.H.; Mathers, T.W. Conditions for deuterium exchange mediated by iridium complexes formed in situ. Tetrahedron 2003, 59, 3349–3358. [Google Scholar] [CrossRef]
  24. Ellames, G.; Gibson, J.; Herbert, J.; McNeill, A. The scope and limitations of deuteration mediated by Crabtree’s catalyst. Tetrahedron 2001, 57, 9487–9497. [Google Scholar] [CrossRef]
  25. Heys, J.R.; Shu, A.Y.L.; Senderoff, S.G.; Phillips, N.M. Deuterium exchange labelling of substituted aromatics using [IrH2(Me2CO)2(PPh3)2BF4. J. Label. Compd. Radiopharm. 1993, 33, 431–438. [Google Scholar] [CrossRef]
  26. Shu, A.; Chen, W.; Heys, J. Organoiridium catalyzed hydrogen isotope exchange: ligand effects on catalyst activity and regioselectivity. J. Organomet. Chem. 1996, 524, 87–93. [Google Scholar] [CrossRef]
  27. Piola, L.; Fernández-Salas, J.A.; Manzini, S.; Nolan, S.P. Regioselective ruthenium catalysed H–D exchange using D2O as the deuterium source. Org. Biomol. Chem. 2014, 12, 8683–8688. [Google Scholar] [CrossRef] [PubMed]
  28. Hesk, D.; Das, P.R.; Evans, B. Deuteration of Acetanilides and Other Substituted Aromatics Using [Ir(COD)(Cy3P)Py)PF6 as Catalyst. J. Label. Compd. Radiopharm. 1995, 36, 497–502. [Google Scholar] [CrossRef]
  29. Brown, J.A.; Cochrane, A.R.; Irvine, S.; Kerr, W.J.; Mondal, B.; Parkinson, J.A.; Paterson, L.C.; Reid, M.; Tuttle, T.; Andersson, S.; et al. The Synthesis of Highly Active Iridium(I) Complexes and their Application in Catalytic Hydrogen Isotope Exchange. Adv. Synth. Catal. 2014, 356, 3551–3562. [Google Scholar] [CrossRef]
  30. Hansch, C.; Leo, A.; Taft, R.W. A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev. 1991, 91, 165–195. [Google Scholar] [CrossRef]
  31. Seeman, J.I. Effect of conformational change on reactivity in organic chemistry. Evaluations, applications, and extensions of Curtin-Hammett Winstein-Holness kinetics. Chem. Rev. 1983, 83, 83–134. [Google Scholar] [CrossRef]
  32. Shang, R.; Fu, Y.; Li, J.B.; Zhang, S.L.; Guo, Q.X.; Liu, L. Synthesis of Aromatic Esters via Pd-Catalyzed Decarboxylative Coupling of Potassium Oxalate Monoesters with Aryl Bromides and Chlorides. J. Am. Chem. Soc. 2009, 131, 5738–5379. [Google Scholar]
  33. Moller, H.; Thimm, H.J. Cosmetic Preparations with Alkoxybenzoic Acid Esters as Inflammation Inhibitors and Method. U.S. Patent 4,136,165; filed 22 April 1977, and issued 23 January 1979,
  34. Mori, N.; Togo, H. Facile oxidative conversion of alcohols to esters usingmolecular iodine. Tetrahedron 2005, 61, 5915–5925. [Google Scholar] [CrossRef]
  35. Jackson, R.W. A mild and selective method for the cleavage of tert-butyl esters. Tetrahedron Lett. 2001, 42, 5163–5165. [Google Scholar] [CrossRef]
  36. Kuh, L.P. Ionization of Organic Esters in Sulfuric Acid. II. Alkyl Oxygen Fission. J. Am. Chem. Soc. 1949, 71, 1575–1577. [Google Scholar]
  37. Rivero-Cruz, B.; Rivero-Cruz, I.; Rodríguez-Sotres, R.; Mata, R. Effect of natural and synthetic benzyl benzoates on calmodulin. Phytochemistry 2007, 68, 1147–1155. [Google Scholar] [CrossRef] [PubMed]
  38. Kaur, G.; Narayanan, V.L.; Risbood, P.A.; Hollingshead, M.G.; Stinson, S.F.; Varma, R.K.; Sausville, E.A. Synthesis, structure–activity relationship, and p210bcr−abl protein tyrosine kinase activity of novel AG 957 analogs. Bioorg. Med. Chem. 2005, 13, 1749–1761. [Google Scholar] [CrossRef] [PubMed]
  39. Kohn, W.; Sham, L. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133–1138. [Google Scholar] [CrossRef]
  40. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other function. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  41. Andrae, D.; Häußerman, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  42. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093. [Google Scholar] [CrossRef] [PubMed]
  • Sample Availability: Not available.

Share and Cite

MDPI and ACS Style

Devlin, J.; Kerr, W.J.; Lindsay, D.M.; McCabe, T.J.D.; Reid, M.; Tuttle, T. Iridium-Catalysed ortho-Directed Deuterium Labelling of Aromatic Esters—An Experimental and Theoretical Study on Directing Group Chemoselectivity. Molecules 2015, 20, 11676-11698. https://doi.org/10.3390/molecules200711676

AMA Style

Devlin J, Kerr WJ, Lindsay DM, McCabe TJD, Reid M, Tuttle T. Iridium-Catalysed ortho-Directed Deuterium Labelling of Aromatic Esters—An Experimental and Theoretical Study on Directing Group Chemoselectivity. Molecules. 2015; 20(7):11676-11698. https://doi.org/10.3390/molecules200711676

Chicago/Turabian Style

Devlin, Jennifer, William. J. Kerr, David M. Lindsay, Timothy J. D. McCabe, Marc Reid, and Tell Tuttle. 2015. "Iridium-Catalysed ortho-Directed Deuterium Labelling of Aromatic Esters—An Experimental and Theoretical Study on Directing Group Chemoselectivity" Molecules 20, no. 7: 11676-11698. https://doi.org/10.3390/molecules200711676

Article Metrics

Back to TopTop