Next Article in Journal
Degradative Effect of Nattokinase on Spike Protein of SARS-CoV-2
Next Article in Special Issue
Synthesis of CF3-Containing Spiro-[Indene-Proline] Derivatives via Rh(III)-Catalyzed C-H Activation/Annulation
Previous Article in Journal
In Vitro Evaluation of In Silico Screening Approaches in Search for Selective ACE2 Binding Chemical Probes
Previous Article in Special Issue
First Trifluoromethylated Phenanthrolinediamides: Synthesis, Structure, Stereodynamics and Complexation with Ln(III)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Synthesis of New α-hydroxy β-fluoro/β-trifluoromethyl and Unsaturated Phosphonates from Carbohydrate-Derived Building Blocks via Pudovik and Horner–Wadsworth–Emmons Reactions †

by
Monika Bilska-Markowska
1,
Wojciech Jankowski
1,
Marcin Hoffmann
1 and
Marcin Kaźmierczak
1,2,*
1
Faculty of Chemistry, Adam Mickiewicz University, Uniwersytetu Poznańskiego 8, 61-614 Poznań, Poland
2
Centre for Advanced Technologies, Adam Mickiewicz University, Uniwersytetu Poznańskiego 10, 61-614 Poznań, Poland
*
Author to whom correspondence should be addressed.
Dedication to Professor Hanna Wójtowicz-Rajchel on the occasion of her retirement.
Molecules 2022, 27(17), 5404; https://doi.org/10.3390/molecules27175404
Submission received: 10 August 2022 / Revised: 21 August 2022 / Accepted: 22 August 2022 / Published: 24 August 2022
(This article belongs to the Special Issue Fluorine Chemistry 2.0)

Abstract

:
Herein, we present the application of fluorinated carbohydrate-derived building blocks for α-hydroxy β-fluoro/β-trifluoromethyl and unsaturated phosphonates synthesis. Pudovik and Horner–Wadsworth–Emmons reactions were applied to achieve this goal. The proposed pathway of the key reactions is supported by the experimental results, as well as quantum chemical calculations. The structure of the products was established by spectroscopic (1D, 2D NMR) and spectrometric (MS) techniques. Based on our data received, we claim that the progress of the Pudovik and HWE reactions is significantly influenced by the acidic protons present in the molecules as assessed by pKa values of the reagent.

Graphical Abstract

1. Introduction

The synthesis of biologically significant compounds is based mainly on the ability to introduce an appropriate group of atoms into a target molecule [1,2]. An interesting approach relevant to this case may be introduction of fluorine containing groups into organic molecules. The relatively small size and electron properties of fluorine gave it a versatile use as a bioisostere and hence it is commonly applied as a substitute for a hydrogen atom. In addition, it is also an excellent mimetic of a carbonyl group, for example in peptide chains [3,4,5,6]. Nature has not endowed us with numerous fluorinated organic derivatives [7,8]. For this reason, synthesis in this area is the only way to obtain such molecules. Due to its remarkable chemical and physical properties, fluorine possessing compounds play an important role in biologically active molecules. This is particularly important in the aspect of obtaining new drugs [9,10,11,12,13,14,15]. Phosphonic acid derivatives are also among the known and used drugs (Fosfomycin 1, Figure 1). These derivatives have found application as analogues of phosphates commonly occurring in living organisms and involved in important biological processes (component of cell membranes—Phosphatidylcholine 2, Figure 1) [16,17]. Replacement of a bridging oxygen atom (R-O-P) in a phosphate group by a methylene group (-CH2-) can have significant biological effects. This modification may affect greater stability because carbon-phosphorus bond cannot be hydrolyzed by the enzyme that cleaves O-phosphate esters. For this reason, these isosteric analogues can be better substrates due to their similarity in size and shape to natural phosphates [18].
In the literature, there are many examples of various properties and biological activities of phosphonic acid derivatives containing a hydroxyl group [19,20] and/or fluorine atom in α position to a phosphonate group [21,22,23,24,25,26]. These compounds are also isosteric analogues of natural products [27] and enzyme inhibitors (Figure 2) [19,28]. There are also examples of phosphonate derivatives having a fluorine atom in a β position to a phosphorus atom [24,29,30] and fluorinated or nonfluorinated unsaturated phosphonates [31,32,33,34,35,36,37,38].
The most reliable approaches widely used in the synthesis of both fluorine-containing and non-fluorinated phosphonate derivatives are the Pudovik reaction (a) and the Horner–Wadsworth–Emmons olefination (b) (Scheme 1). The most frequently used method of α-hydroxyphosphonates synthesis is addition of dialkyl phosphite to a carbonyl compound, named after its discoverer—the Pudovik reaction [39,40]. HWE is a very highly stereoselective method for preparation of not only α,β-unsaturated phosphonates, but also other compounds, especially various types of natural product derivatives [41,42,43,44,45].
An efficient approach towards the synthesis of fluoro-containing compounds is through modular assembly using fluorinated building blocks. In the recent study we investigated application of the Claisen rearrangement reaction for construction of fluorine-containing scaffolds [46]. Carbohydrates were used as starting compounds, which are a very valuable source of building blocks, as well as chiral auxiliaries [47,48,49,50,51,52]. Specific intent of our present research effort is using early obtained carbohydrate-derived building blocks to receive new α-hydroxy β-fluoro/β-trifluoromethyl and unsaturated phosphonates via Pudovik and Horner–Wadsworth–Emmons reactions. The modifiable functional groups of our newly synthesized compounds give access for their further transformations. They can be considered feasible tools for more complex molecule design.

2. Results and Discussion

In the present studies we report synthesis of fluorine containing phosphonates, 14–15, 20–21. 1,2;5,6-Di-O-isopropylidene-α-d-mannitol 9 was used as a starting material and successfully converted into 14–15, 20–21 via 12 and 18 (Scheme 2). Fluorinated ethers 11 and 12 were prepared using a strategy previously developed in our research group [46,48].
Allyl-vinyl ethers 12, as well as scaffolds bearing fluorine atoms resulting from Claisen rearrangement 18 were used as excellent precursors for the synthesis of fluorinated phosphonates (Scheme 2). The former 12, depending on the olefin employed, can be yielded as a single Z isomer or a Z/E mixture, and readily transformed to a carbonyl compound 13 by reacting with periodic acid. The aldehyde 13 after evaporation of solvent was pure enough be subjected to subsequent reactions. For this purpose, tetraethyl methylenebisphosphonate was converted into appropriate anion with NaH and proceeded with pure Z isomer of 13 to HWE protocol (Scheme 3). The reaction was carried out with 1.75 eq of both tetraethyl methylenebisphosphonate and NaH in THF, for 15 h at room temperature. Surprisingly, the reaction did not progress according to our assumption. On the basis of 1H, 13C NMR spectra of the obtained product 14, it can be concluded that in the initial stage of the process, under basic conditions, the substrate 13 isomerized to intermediate compound int-A, and afterwards the expected HWE reaction took place. In the following Scheme 3, a pathway of the discussed transformation is proposed. The observed 1,3-shift in the first step of the overall process is related to the presence of an acidic proton in the α position to the carbonyl group of fluorinated β,γ-unsaturated aldehyde 13, which can be abstracted under reaction conditions. In the literature, a few examples of this type of isomerization of carbonyl compounds can be found, some of them occurring even with very weak bases. [53,54,55] However, to the best of our knowledge, this is the first known example of such a 1,3-shift involving β,γ-unsaturated aldehydes. As a result of HWE reaction, between α,β-unsaturated aldehyde int-A and nucleophilic derivative of tetraethyl methylenebisphosphonate, an α,β-unsaturated phosphonate with a fluorinated allyl group 14 was synthesized. In spectra of purified product 14, two sets of signals in the ratio 1:0.2 were observed. It is caused by the presence of two diene conformers: “s-cis14′ and “s-trans14. Geometry of the double bond located directly to the phosphorus atom was determined based on the NMR analysis, and is consistent with the expectations of the HWE outcome where the E isomer should be formed as the main product. [56] A coupling constant value was crucial for the analysis- it was found to be 17 Hz, which confirmed the trans relationship of the protons across the double bond.
In order to study the reactivity and diastereoselectivity of the reaction with another nucleophilic agent, this time a mixture of Z/E isomers of aldehyde 13 was subjected to the Pudovik reaction (Scheme 4). This time the reaction was carried out under very mild conditions, without a solvent, and in the presence of stoichiometric amounts of diethyl phosphite, and triethylamine at room temperature. The reaction, as described previously, proceeded in an alkaline environment. In the first step, the double bond rearrangement was observed, and then followed by the actual nucleophilic attack. The Pudovik reaction was highly stereoselective, such a great selectivity can be explained by the Felkin-Ahn model presented in Figure 3.
Having results of both reactions, it can be concluded that the progress of those transformations is strongly dependent on the presence of an acidic proton in the α position of the carbonyl group in 13.
In the present work we show a second class of molecules that we functionalized- these are the previously obtained by Claisen rearrangement α-fluoro-α-trifluoromethyl-γ,δ-unsaturated 18a/18a′ (1:0.65 ratio) and α-trifluoromethyl-γ,δ-unsaturated 18b/18b′ (1:1 ratio) alcohols [46]. These compounds were oxidized with quantitative yields to the corresponding aldehydes 19a/19a′ and 19b/19b′ using Dess-Martin reagent. The reactions were carried out under mild conditions at room temperature. Complete disappearance of the substrates and selective formation of new products was observed after only 5 min. Aldehydes 19a/19a′ and 19b/19b′ were isolated by simple extraction, and were pure enough to be subjected to further transformations. Both fluorinated carbonyl compounds 19a/19a′, 19b/19b′ were subjected to HWE and Pudovik reactions under the same conditions as described above as well.
The HWE reaction carried out on α-fluoro-α-trifluoromethyl-γ,δ-unsaturated 19a/19a′ derivatives proceeded as expected and yielded E olefins 20a/20a′ as the main products (Scheme 5). In this case the coupling constant between the protons located at the newly produced double bond was also 17 Hz. Unfortunately, an analogous reaction with α-trifluoromethyl-γ,δ-unsaturated 19b/19b′ aldehydes provided only traces of the expected HWE products 20b/20b′. These compounds were characterized on the basis of 1H, 19F, 31P and HRMS spectra. 20b/20b′ occurred to be unstable during prolonged NMR experiments (13C, 2D NMR). Perhaps, the initial abstraction of acidic proton occurs, and then highly electrophilic and reactive structures are formed.
Aldehydes 19a/19a′ and 19b/19b′ were also subjected to the Pudovik reaction (Scheme 6). When the derivatives containing an α-fluoro-α-trifluoromethyl-γ,δ-unsaturated moiety 19a/19a′ were used as substrates, the corresponding hydroxy phosphonates 21a/21a′ were obtained with good yield. A very high diastereoselectivity of the process was observed this time. The reaction led to only one new stereogenic center on a carbon atom bonded to a phosphorus atom. The reaction with the second aldehydes 19b/19b′, resulted in a slight reduction in the yield of the products 21b/21b′. This may also be related to the presence of an acidic proton located adjacent to the -CF3 group.
In order to explain and investigate the unexpected reactions of aldehydes 13 and 19b/19b, we decided to perform quantum mechanical calculations to determine their pKa values.
DFT Calculations
The aim of the calculations was to obtain calculated value of pKa to estimate the acid strength of 2-((2,3,3,3-tetrafluoroprop-1-en-1-yl)oxy)but-3-enal 13 and α-trifluoromethyl-γ,δ-unsaturated aldehyde 19b in comparison to other organic acids whose pKa were measured in aqueous solution. Energy calculations were performed within DFT framework at M062X/6-31+G(d) level of theory [57,58]. This method and basis set was used by Zeng et al. [59] for similar calculations which resulted in calculated values of pKa. We also conducted frequency analysis [60] to access thermochemistry data needed, and to verify that the geometry optimization led to the potential energy minimum. Calculations were carried out in vacuo and aqueous solution with use of the Polarizable Continuum Model (PCM) [61]. All quantum mechanics calculations were performed with the GAUSSIAN 09 [62] available within Pl-Grid infrastructure. To calculate pKa we used thermodynamic cycle depicted in Figure 4 below.
Such a type of thermodynamic cycle was recommended in papers by Liptak et al. and by Zeng et al. [59,63]. We applied methodology presented in those articles in our work. To calculate pKa we used equation (Equation (1)) proposed by Zeng et al. [59] included below.
pK a = A Δ G aq RTln 10 + B
where:
Δ G aq solution   phase   free   energy
R gas   constant
T temperature
A   &   B linear   parameters
Parameters A and B were determined on the basis of known pKa values for five chemical compounds: formic acid, trifluoroacetic acid, dimethyl malonate, acetone and chloroacetone. These compounds were also optimized in vacuo and water solution, and on the basis of information gathered from frequency analyses we were able to calculate pKa values for each structure. Calculated and known experimental values of pKa of formic acid, trifluoroacetic acid, dimethyl malonate, acetone and chloroacetone are gathered in Table 1 below.
As shown in Table 1 values of calculated pKa differed slightly from experimentally measured values of pKa (values of empirical parameters A and B needed in Equation (1). were 0.3541 and −0.3387 for parameters A and B, respectively). The calculated values of pKa for 2-((2,3,3,3-tetrafluoroprop-1-en-1-yl)oxy)but-3-enal 13 and α-trifluoromethyl-γ,δ-unsaturated aldehyde 19b were 13.3 and 15.1, respectively. Obtained pKa values suggest that acidity of 13 is similar to acidity of dimethyl malonate (experimental pKa value 13.5), while acidity of 19b is greater than the acidity of chloroacetone (experimental pKa value 16.5). Examples of use of, weak bases such as K2CO3, Et3N, and NaH to abstract acidic protons from dimethyl malonate or chloroacetone can be found in the literature. [65,66,67,68] Thus, the obtained calculated results of pKa values of 13 and 19b correlate with experimental observations, in which under the reaction conditions the abstraction of acidic proton can be observed in the first step of the reaction.

3. Conclusions

In conclusion, starting from conveniently obtained via Claisen rearrangement fluorinated scaffolds 18 or allyl-vinyl ethers 12, new α-hydroxy β-fluoro/β-trifluoromethyl and unsaturated phosphonates have been achieved. The key reactions proceeded under mild conditions with good yields. The unexpected progress of the reactions has been correlated with quantum mechanical calculations and indicates connection with pKa values of the molecules 13 and 19b/19b′. To the best of our knowledge of the ever-growing interest in fluorinated phosphonates as enzyme inhibitors, the compounds obtained in this research can serve as material for future synthetic studies. These compounds have been also prepared as examples to illustrate the synthetic potential of fluorinated derivatives 12 and 18.

4. Materials and Methods

4.1. General Methods

1H NMR, 13C NMR, 19F NMR, 31P NMR and 2D-NMR spectra were performed on Bruker ASCEND 400 (400 MHz) and Bruker ASCEND 600 (600 MHz) spectrometers. Chemical shifts of 1H NMR were expressed in parts per million downfield from TMS as an internal standard (δ = 0) in CDCl3. Chemical shifts of 13C NMR were expressed in parts per million downfield and upfield from CDCl3 as an internal standard (δ = 77.0). Chemical shifts of 19F NMR were expressed in parts per million upfield from CFCl3 as an internal standard (δ = 0) in CDCl3. 31P NMR chemical shifts were expressed in parts per million in CDCl3. High-resolution mass spectra were recorded by electron spray (MS-ESI) technique using QToF Impact HD Bruker spectrometer. Reagent grade chemicals were used and solvents were dried by refluxing with sodium metal-benzophenone (THF) and distilled under an argon atmosphere. All moisture sensitive reactions were carried out under an argon atmosphere using oven-dried glassware. Reaction temperatures below 0 °C were obtained using a bath cooling (dry ice/iso-propanol). Thin-layer chromatography (TLC) was performed on Merck Kieselgel 60-F254 with EtOAc/hexane as developing systems. Visualization of the reactions products was achieved using UV light (254 nm) and a standard procedure (solution of phosphomolybdenic acid or solution of potassium permanganate). Merck Kieselgel 60 (230–400 mesh) was used for column chromatography (more detailed data in Supplementary Materials).

4.2. General Procedure of Periodic Acid Oxidation Synthesis of 13

A solution of 12/12′ (1 equiv.) and H5IO6 (2 equiv.) in EtOAc (5 mL) was stirred for 4.5 h. The mixture was filtered, and the filtrate was evaporated to give the crude product as a creamy oil. The crude product was used for the next step [69].

4.3. General procedure of Dess-Martin oxidation Synthesis of 19

The Dess-Martin periodinane (1.2 equiv.) was added to a stirred solution of 18 (1 equiv.) in 5 mL of CH2Cl2 at RT. After 5 min the solution was diluted with CH2Cl2 and a mixture of sat. NaHCO3 and aq. Na2S2O3 (1:1) 5 mL. After stirring for 5 min the phases were separated. Organic phases were dried (Na2SO4) and concentrated to give oils which were used without further purification [70].

4.4. General Procedure of HWE Reaction Synthesis of 14 and 20

Tetraethyl methylenediphosphonate (1.75 equiv.) was added dropwise to a mixture of sodium hydride (60% dispersion in oil, 1.75 equiv.) in 2 mL of dry THF at 0 °C under argon. The reaction mixture was stirred for 10 min at 0 °C and then stirred at room temperature for 20 min. The reaction mixture was cooled to 0 °C, a solution of aldehyde 13/13′ (1.75 equiv.) in 1 mL of dry THF was added, and the resulting solution was stirred at room temperature for 15 h. The reaction mixture was concentrated, and the residue was purified using column chromatography (n-hexane/ethyl acetate 90:10 50:40) [71].
Diethyl ((1E,3E)-3-(((E)-2,3,3,3-tetrafluoroprop-1-en-1-yl)oxy)penta-1,3-dien-1-yl)phosphonate (14/14′): Colorless oil (73 mg, 39 %, ratio 1:0.2): Major: 1H NMR (400 MHz, Chloroform-d) δ = 6.95 (dd, J = 21.9, 17.0 Hz, 1H, PCH=CH), 6.39 (dq, J = 18.5, 1.1 Hz, 1H, OCH=CFCF3), 5.81 (t, J = 17.3 Hz, 1H, PCH=CH), 5.66 (dq, J = 7.2, 1.7 Hz, 1H, =CHCH3), 4.13 (q, J = 7.1 Hz, 2H, OCH2CH3), 1.83 (d, J = 7.2 Hz, 3H, =CHCH3), 1.34 (t, J = 7.0 Hz, 3H, OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 151.33 (d, J = 25.0 Hz, CH=CO), 142.00 (d, J = 7.8 Hz, OCH=CFCF3), 134.34 (dq, J = 251.9, 39.4 Hz, OCH=CFCF3), 132.89–132.66 (m, PCH=CH), 123.54 (s, CH=CH3), 119.52 (dq, J = 269.1, 34.5 Hz, CF3), 114.10 (d, J = 191.8 Hz, PCH=CH), 62.18 (d, J = 5.6 Hz, OCH2CH3), 61.98 (d, J = 5.5 Hz, OCH2CH3), 16.28 (s, 2 × OCH2CH3). 11.52 (s, CH3). 31P NMR (162 MHz, Chloroform-d) δ = 18.33 (s). 19F NMR (376 MHz, Chloroform-d) δ = −71.47 (d, J = 14.7 Hz, CF3), −160.73 (dq, J = 18.2, 14.5 Hz, F). Minor: 1H NMR (400 MHz, Chloroform-d) δ = 6.92 (dd, J = 21.8, 17.1 Hz, 1H, PCH=CH), 6.53 (dq, J = 18.1, 1.1 Hz, 1H, OCH=CFCF3), 5.99 (t, J = 16.9 Hz, 1H, PCH=CH), 5.67 (dq, J = 7.2, 2.0 Hz, 1H, =CHCH3), 4.10 (q, J = 7.2 Hz, 2H, OCH2CH3), 2.07 (d, J = 8.3 Hz, 3H, =CHCH3), 1.26 (t, J = 7.1 Hz, 3H, OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 152.04 (d, J = 25.0 Hz, CH=CO), 140.65 (d, J = 7.8 Hz, OCH=CFCF3), 134.34 (dq, J = 251.9, 39.4 Hz, OCH=CFCF3), 132.39–132.24 (m, PCH=CH), 120.56 (s, CH=CH3), 119.52 (dq, J = 269.1, 34.5 Hz, CF3), 118.23 (d, J = 190.6 Hz, PCH=CH), 62.18 (d, J = 5.6 Hz, OCH2CH3), 61.98 (d, J = 5.5 Hz, OCH2CH3), 16.22 (s, 2 × OCH2CH3). 11.50 (s, CH3). 31P NMR (162 MHz, Chloroform-d) δ = 16.82 (s). 19F NMR (376 MHz, Chloroform-d) δ = −71.59 (d, J = 14.6 Hz, CF3), −159.14 (dq, J = 18.4, 14.6 Hz, F). HRMS (ESI) calcd. for C12H18F4O4P ([M + H]+): 333.0879, found: 333.0873.
Diethyl ((1E,5E)-6-((S)-2,2-dimethyl-1,3-dioxolan-4-yl)-3-fluoro-3-(trifluoromethyl)hexa-1,5-dien-1-yl)phosphonate (20a/20a′): Colorless oil (82 mg, 72%, ratio:1:0.65): Major: 1H NMR (400 MHz, Chloroform-d) δ = 6.62 (dddd, J = 23.0, 20.4, 17.1, 6.2 Hz, 1H, CH=CHP), 6.19 (t, J = 17.1, 1H, CH=CHP), 5.64 (dd, J = 15.6, 6.2 Hz, 1H, CH=CHCH2), 5.61 (dd, J = 15.5, 6.0 Hz, 1H, CH=CHCH2), 4.46 (q, J = 7.6 Hz, 1H, CHCH=CH), 4.16–4.03 (m, 4H, 2 × OCH2CH3), 4.06 (dd, J = 8.2, 6.1 Hz, 1H, OCHH), 3.56 (t, J = 8.1 Hz, 1H, OCHH), 2.89–2.74 (m, 1H, CHHCF), 2.71–2.51 (m, 1H, CHHCF), 1.40 (s, 3H, CH3), 1.36 (s, 3H, CH3), 1.35–1.32 (m, 6H, 2 × OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 141.44 (dd, J = 10.6, 7.8 Hz, CH=CHP), 134.77 (s, CH=CHCH2), 123.40 (dd, J = 185.7, 8.8 Hz, CH=CHP), 123.08 (d, J = 4.2 Hz, CH=CHCH2), 122.36 (dq, J = 284.3, 30.3 Hz, CF3), 109.45 (s, C(CH3)2), 93.76 (dq, J = 194.4, 31.2 Hz, CF), 76.24 (s, CHCH=CH), 69.27 (s, OCH2), 62.34 (d, J = 5.5 Hz, 2 × OCH2CH3), 35.16 (d, J = 22.3 Hz, CH2CF), 26.58 (s, CH3), 25.81 (s, CH3), 16.40–16.22 (m, 2 × OCH2CH3). 31P NMR (162 MHz, Chloroform-d) δ = 14.96 (s). 19F NMR (376 MHz, Chloroform-d) δ = −80.24 (d, J = 7.3 Hz, CF3), −177.00 (dq, J = 7.1, 2.2 Hz, F). Minor: 1H NMR (400 MHz, Chloroform-d) δ = 6.62 (dddd, J = 23.0, 20.4, 17.1, 6.2 Hz, 1H, CH=CHP), 6.19 (t, J = 17.1, 1H, CH=CHP), 5.64 (dd, J = 15.6, 6.2 Hz, 1H, CH=CHCH2), 5.61 (dd, J = 15.5, 6.0 Hz, 1H, CH=CHCH2), 4.48 (q, J = 7.9 Hz, 1H, CHCH=CH), 4.16–4.03 (m, 4H, 2 × OCH2CH3), 4.06 (dd, J = 8.2, 6.1 Hz, 1H, OCHH), 3.56 (t, J = 8.1 Hz, 1H, OCHH), 2.89–2.74 (m, 1H, CHHCF), 2.71–2.51 (m, 1H, CHHCF), 1.41 (s, 3H, CH3), 1.38 (s, 3H, CH3), 1.35–1.32 (m, 6H, 2 × OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 141.29 (dd, J = 11.0, 8.2 Hz, CH=CHP), 134.83 (s, CH=CHCH2), 128.48 (dd, J = 186.8, 9.0 Hz, CH=CHP), 122.96 (d, J = 4.0 Hz, CH=CHCH2), 122.36 (dq, J = 284.3, 30.3 Hz, CF3), 109.51 (s, C(CH3)2), 93.63 (dq, J = 194.4, 31.3 Hz, CF), 76.22 (s, CHCH=CH), 69.27 (s, OCH2), 62.29 (d, J = 5.7 Hz, OCH2CH3), 62.20 (d, J = 5.7 Hz, OCH2CH3), 35.15 (d, J = 19.8 Hz, CH2CF), 26.58 (s, CH3), 25.80 (s, CH3), 16.40–16.22 (m, 2 × OCH2CH3). 31P NMR (162 MHz, Chloroform-d) δ = 14.95 (s). 19F NMR (376 MHz, Chloroform-d) δ = −80.27 (d, J = 7.3 Hz, CF3), −176.85 (dq, J = 7.0, 2.1 Hz, F). HRMS (ESI) calcd. for C16H25F4O5PNa ([M + Na]+): 427.1273, found: 427.1270.
Diethyl ((1E,5E)-6-((S)-2,2-dimethyl-1,3-dioxolan-4-yl)-3-(trifluoromethyl)hexa-1,5-dien-1-yl)phosphonate (20b/20b′): Colorless oil (8 mg, 16 %, ratio 1:0.9): Major: 1H NMR (400 MHz, Chloroform-d) δ = 6.87 (dd, J = 29.4, 1.9 Hz, 1H, CH=CHP), 5.72 (t, J = 17.2, 1H, CH=CHP), 5.80–5.66 (m, 1H, CH=CHCH2), 5.48 (dd, J = 15.4, 7.4 Hz, 1H, CH=CHCH2), 4.48 (q, J = 7.0 Hz, 1H, CHCH=CH), 4.09 (q, J = 6.9 Hz, 4H, 2 × OCH2CH3), 4.08 (dd, J = 8.1, 6.2 Hz, 1H, OCHH), 3.55 (q, J = 8.2 Hz, 1H, CHCF3), 3.52 (t, J = 7.9 Hz, 1H, OCHH), 3.20–2.94 (m, 2H, CH2CHCF3), 1.41 (s, 3H, CH3), 1.38 (s, 3H, CH3), 1.33 (t, J = 7.0 Hz, 6H, 2 × OCH2CH3). 31P NMR (162 MHz, Chloroform-d) δ = 18.76–18.35 (m). 19F NMR (376 MHz, Chloroform-d) δ = −65.53 (s, 3F, CF3). Minor: 1H NMR (400 MHz, Chloroform-d) δ = 7.24 (dd, J = 22.3, 17.4 Hz, 1H, CH=CHP), 5.72 (t, J = 17.2, 1H, CH=CHP), 5.80–5.66 (m, 1H, CH=CHCH2), 5.62 (dd, J = 15.4, 7.4 Hz, 1H, CH=CHCH2), 4.45 (q, J = 7.1 Hz, 1H, CHCH=CH), 4.09 (q, J = 6.9 Hz, 4H, 2 × OCH2CH3), 4.07 (dd, J = 8.2, 7.2 Hz, 1H, OCHH), 3.55 (q, J = 8.2 Hz, 1H, CHCF3), 3.52 (t, J = 7.9 Hz, 1H, OCHH), 3.20–2.94 (m, 2H, CH2CHCF3), 1.40 (s, 3H, CH3), 1.37 (s, 3H, CH3), 1.33 (t, J = 7.0 Hz, 6H, 2 × OCH2CH3). 31P NMR (162 MHz, Chloroform-d) δ = 18.76–18.35 (m). 19F NMR (376 MHz, Chloroform-d) δ = −65.44 (s, 3F, CF3). HRMS (ESI) calcd. for C16H26F3O5PNa ([M + Na]+): 409.1368, found: 409.1371.

4.5. General Procedure of Pudovik Reaction Synthesis of 15 and 21

TEA (1.0 equiv.) was added under an argon atmosphere to a mixture of stirred solution of diethyl phosphite (1.0 equiv.) and an aldehyde 13 or 19 (1.0 equiv.). The reaction mixture was stirred at room temperature overnight, then was diluted with 20 mL of water and extracted with ethyl acetate (3 × 15 mL). The organic layers were washed with NaClsat., dried over MgSO4 or Na2SO4, filtrated and concentrated under reduced pressure. The residue was purified using column chromatography (chloroform/methanol 100:0 100:0.5) [26].
Diethyl ((2Z)-1-hydroxy-2-((2,3,3,3-tetrafluoroprop-1-en-1-yl)oxy)but-2-en-1-yl)phosphonate (15/15′): Colorless oil (46 mg, 49%, ratio 1:0.1): Major: 1H NMR (400 MHz, Chloroform-d) δ = 6.63 (dd, J = 19.0, 1.2 Hz, 1H, OCH=CFCF3), 5.50 (dq, J = 7.0, 3.6 Hz, 1H, =CHCH3), 4.75 (bs, OH), 4.46 (d, J = 12.5 Hz, 1H, PCHOH), 4.20 (q, J = 7.2 Hz, 2H, OCH2CH3), 1.72 (d, J = 7.1 Hz, 3H, =CHCH3), 1.34 (t, J = 7.1 Hz, 3H, OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 150.02 (d, J = 4.0 Hz, CH(OH)=CO), 134.00 (m, OCH=CFCF3), 133.53 (dq, J = 249.4, 39.2 Hz, OCH=CFCF3), 119.87 (dq, J = 268.8, 34.6 Hz, CF3), 114.09 (d, J = 9.4 Hz, =CHCH3), 68.98 (d, J = 161.5 Hz, CH(OH)=CO), 63.81 (d, J = 7.0 Hz, OCH2CH3), 63.23 (d, J = 7.3 Hz, OCH2CH3), 16.26 (d, J = 5.9 Hz, OCH2CH3), 16.29 (d, J = 5.9 Hz, OCH2CH3),10.43 (d, J = 1.5 Hz, CH3). 31P NMR (162 MHz, Chloroform-d) δ = 19.87 (s). 19F NMR (376 MHz, Chloroform-d) δ = −71.40 (d, J = 14.7 Hz, CF3), −163.22 (dq, J = 19.1, 15.0 Hz, F). Minor: 1H NMR (400 MHz, Chloroform-d) δ = 6.99 (d, J = 8.8 Hz, 1H, OCH=CFCF3), 5.50 (dq, J = 7.0, 3.6 Hz, 1H, =CHCH3), 4.75 (bs, OH), 4.46 (d, J = 12.5 Hz, 1H, PCHOH), 4.20 (q, J = 7.2 Hz, 2H, OCH2CH3), 1.71 (d, J = 6.9 Hz, 3H, =CHCH3), 1.35 (t, J = 6.6 Hz, 3H, OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 150.02 (d, J = 4.0 Hz, CH(OH)=CO), 134.00 (m, OCH=CFCF3), 133.53 (dq, J = 249.4, 39.2 Hz, OCH=CFCF3), 119.87 (dq, J = 268.8, 34.6 Hz, CF3), 114.09 (d, J = 9.4 Hz, =CHCH3), 68.91 (d, J = 161.4 Hz, CH(OH)=CO), 63.76 (d, J = 6.7 Hz, OCH2CH3), 63.19 (d, J = 7.4 Hz, OCH2CH3), 16.26 (d, J = 5.9 Hz, OCH2CH3), 16.29 (d, J = 5.9 Hz, OCH2CH3),10.35 (d, J = 1.6 Hz, CH3). 31P NMR (162 MHz, Chloroform-d) δ = 19.95 (s). 19F NMR (376 MHz, Chloroform-d) δ = −68.45 (d, J = 11.4 Hz, CF3), −180.98 (dq, J = 11.4, 8.9 Hz, F). HRMS (ESI) calcd. for C11H17F4O5PNa ([M + Na]+): 359.0647, found: 359.0648.
Diethyl ((E)-5-((S)-2,2-dimethyl-1,3-dioxolan-4-yl)-2-fluoro-1-hydroxy-2-(trifluoromethyl)pent-4-en-1-yl)phosphonate (21a/21a′): Colorless oil (68mg, 87 %, ratio 1:0.65): Major: 1H NMR (400 MHz, Chloroform-d) δ = 5.77–5.73 (m, 1H, CH=CHCH2), 5.61 (d, J = 6.8 Hz, 1H, CH=CHCH2), 4.67 (td, J = 7.3, 1H, OH), 4.44 (q, J = 7.1 Hz, 1H, CHCH=CH), 4.36–4.25 (m, 1H, CHP), 4.22–4.06 (m, 4H, 2 × OCH2CH3), 4.01 (dd, J = 8.2, 6.2 Hz, 1H, OCHH), 3.52 (dd, J = 7.9, 2.3 Hz, 1H, OCHH), 3.08–2.92 (m, 1H, CHHCF), 2.92–2.74 (m, 1H, CHHCF), 1.35 (s, 3H, CH3), 1.31 (s, 3H, CH3), 1.28 (t, J = 7.0 Hz, 6H, 2 × OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 133.24 (s, CH=CHCH2), 125.32 (d, J = 5.9 Hz, CH=CHCH2), 127.63–118.60 (m, J = 287.3 Hz, CF3), 109.35 (s, C(CH3)2), 95.75–92.17 (m, CF), 76.52 (s, CHCH=CH), 69.26 (s, OCH2), 67.64 (dd, J = 163.9, 29.3 Hz, CP), 63.86 (d, J = 4.2 Hz, 2 × OCH2CH3), 63.40 (d, J = 7.2 Hz, OCH2CH3), 32.63 (d, J = 20.4 Hz, CH2CF), 26.57 (s, CH3), 25.81 (s, CH3), 16.30 (“t”, J = 5.8 Hz, 2 × OCH2CH3). 31P NMR (162 MHz, Chloroform-d) δ = 18.41–18.20 (m). 19F NMR (376 MHz, Chloroform-d) δ = −76.49–−76.61 (m, CF3), −174.27 (ddh, J = 27.8, 13.4, 6.7 Hz, F). 19F{1/H} NMR (376 MHz, Chloroform-d) δ = −76.53 (dd, J = 6.5, 3.3 Hz, CF3), −174.26 (p, J = 6.6 Hz, F). Minor: 1H NMR (400 MHz, Chloroform-d) δ = 5.83–5.77 (m, 1H, CH=CHCH2), 5.58 (d, J = 6.8 Hz, 1H, CH=CHCH2), 4.67 (td, J = 7.3, 1H, OH), 4.44 (q, J = 7.1 Hz, 1H, CHCH=CH), 4.36–4.25 (m, 1H, CHP), 4.22–4.06 (m, 4H, 2 × OCH2CH3), 4.01 (dd, J = 8.2, 6.2 Hz, 1H, OCHH), 3.50 (dd, J = 7.9, 2.3 Hz, 1H, OCHH), 3.08–2.92 (m, 1H, CHHCF), 2.92–2.74 (m, 2H), 1.35 (s, 3H, CH3), 1.31 (s, 3H, CH3), 1.28 (t, J = 7.0 Hz, 6H, 2 × OCH2CH3). 13C NMR (101 MHz, Chloroform-d) δ = 133.18 (s, CH=CHCH2), 125.32 (d, J = 5.9 Hz, CH=CHCH2), 127.63–118.60 (m, J = 287.3 Hz, CF3), 109.33 (s, C(CH3)2), 95.75–92.17 (m, CF), 76.59 (s, CHCH=CH), 69.22 (s, OCH2), 67.61 (dd, J = 163.0, 28.9 Hz, CP), 63.78 (d, J = 4.4 Hz, OCH2CH3), 63.38 (d, J = 7.3 Hz, OCH2CH3), 32.59 (d, J = 20.9 Hz, CH2CF), 26.58 (s, CH3), 25.81 (s, CH3), 16.30 (t, J = 5.8 Hz, 2 × OCH2CH3). 31P NMR (162 MHz, Chloroform-d) δ = 18.41–18.20 (m). 19F NMR (376 MHz, Chloroform-d) δ = −76.41–−76.49 (m, CF3), −174.95 (dtt, J = 28.0, 13.4, 6.7 Hz, F). 19F{1/H} (376 MHz, Chloroform-d) δ = −76.45 (dd, J = 6.4, 3.3 Hz, CF3), −174.94 (p, J = 6.7 Hz, F). HRMS (ESI) calcd. for C15H25F4O6PNa ([M + Na]+): 431.1223, found: 431.1226.
Diethyl ((E)-5-((S)-2,2-dimethyl-1,3-dioxolan-4-yl)-1-hydroxy-2-(trifluoromethyl)pent-4-en-1-yl)phosphonate (21b/21b′): Colorless oil (54 mg, 69%, ratio 1:1): Major/Minor: 1H NMR (400 MHz, Chloroform-d) δ = 5.95–5.82 (m, 1H, CH=CHCH2), 5.65–5.55 (CH=CHCH2), 4.50 (q, J = 8.0, 7.5 Hz, 1H, CHCH=CH), 4.36 (dd, J = 13.1, 6.9 Hz, 1H, CHP), 4.28–4.16 (m, 4H, 4H, 2 × OCH2CH3), 4.09 (ddd, J = 7.8, 5.4, 1.4 Hz, 1H, OCHH), 3.58 (td, J = 7.9, 3.9 Hz, 1H, OCHH), 2.80–2.68 (m, 1H, CHHCHCF3), 2.68–2.57 (m, 1H, CHHCHCF3), 2.56–2.45 (m, 1H, CHCF3), 1.44 (s, 3H, CH3), 1.40 (s, 3H, CH3), 1.39–1.34 (m, 6H, 2 × OCH2CH3). 31P NMR (162 MHz, Chloroform-d) δ = 21.41 (s). Major: 13C NMR (101 MHz, Chloroform-d) δ = 131.59 (s, CH=CHCH2), 130.23 (s, CH=CHCH2), 129.79–128.38 (m, CF3), 109.22 (s, C(CH3)2), 76.81 (s, CHCH=CH), 69.35 (s, OCH2), 65.30 (dd, J = 167.5, 2.8 Hz, CP), 63.20 (d, J = 2.4 Hz, OCH2CH3), 63.13 (d, J = 2.4 Hz, OCH2CH3), 45.13–43.97 (m, CCF3), 26.64 (s, CH3), 26.13 (s, CH2CHCF3), 25.85 (s, CH3), 16.47 (s, OCH2CH3), 16.42 (s, OCH2CH3). 19F NMR (376 MHz, Chloroform-d) δ = −67.82 (d, J = 3.2 Hz, CF3). Minor: 13C NMR (101 MHz, Chloroform-d) δ = 131.34 (s, CH=CHCH2), 130.23 (s, CH=CHCH2), 129.79–128.38 (m, CF3), 109.21 (s, C(CH3)2), 76.81 (s, CHCH=CH), 69.30 (s, OCH2), 65.27 (dd, J = 167.5, 2.9 Hz, CP), 63.40 (d, J = 4.5 Hz, OCH2CH3), 63.33 (d, J = 4.5 Hz, OCH2CH3), 45.05–44.11 (m, CCF3), 26.63 (s, CH3), 25.85 (s, CH3), 25.82 (s, CH2CHCF3), 16.38 (s, OCH2CH3), 16.36 (s, OCH2CH3). 19F NMR (376 MHz, Chloroform-d) δ = −67.96 (d, J = 3.6 Hz, CF3). HRMS (ESI) calcd. for C15H26F3O6PNa ([M + Na]+): 413.1317, found: 413.1320.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27175404/s1, Copies of NMR spectra, DFT coordinates.

Author Contributions

Conceptualization, M.K. and M.B.-M.; methodology, M.K. and W.J.; validation, M.K. and M.B.-M.; formal analysis, M.K., M.B.-M., W.J. and M.H.; investigation, M.K. and M.B.-M.; resources, M.K., M.B.-M. and W.J.; data curation, M.K. and M.B.-M.; writing—original draft preparation, M.K., M.B.-M., W.J. and M.H.; writing—review and editing, M.K.; visualization, M.K. and M.B.-M.; supervision, M.K.; project administration, M.K. and M.B.-M.; funding acquisition, M.H., M.K. and M.B.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by Adam Mickiewicz University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available from the corresponding author.

Acknowledgments

M. Bilska-Markowska would like to thank the National Science Centre for partial financial support under grant MINIATURA 2 No 2018/02/X/ST5/03007. M. Kaźmierczak would like to thank the National Science Centre for partial financial support under grant MINIATURA 2 No 2018/02/X/ST5/02867. This research was supported in part by PL-Grid infrastructure.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grygorenko, O.O.; Volochnyuk, D.M.; Vashchenko, B.V. Emerging Building Blocks for Medicinal Chemistry: Recent Synthetic Advances. Eur. J. Org. Chem. 2021, 2021, 6478–6510. [Google Scholar] [CrossRef]
  2. Zimmer, L.E.; Sparr, C.; Gilmour, R. Fluorine Conformational Effects in Organocatalysis: An Emerging Strategy for Molecular Design. Angew. Chem. Int. Ed. 2011, 50, 11860–11871. [Google Scholar] [CrossRef]
  3. Meanwell, N.A. Fluorine and Fluorinated Motifs in the Design and Application of Bioisosteres for Drug Design. J. Med. Chem. 2018, 61, 5822–5880. [Google Scholar] [CrossRef]
  4. Jäckel, C.; Koksch, B. Fluorine in Peptide Design and Protein Engineering. Eur. J. Org. Chem. 2005, 2005, 4483–4503. [Google Scholar] [CrossRef]
  5. Buer, B.C.; Marsh, E.N. Fluorine: A new element in protein design. Protein Sci. 2012, 21, 453–462. [Google Scholar] [CrossRef] [Green Version]
  6. Sloand, J.N.; Miller, M.A.; Medina, S.H. Fluorinated peptide biomaterials. Pept. Sci. 2021, 113, e24184. [Google Scholar] [CrossRef]
  7. O’Hagan, D.; Schaffrath, C.; Cobb, S.L.; Hamilton, J.T.G.; Murphy, C.D. Biosynthesis of an organofluorine molecule. Nature 2002, 416, 279. [Google Scholar] [CrossRef]
  8. Hall, R.J. The distribution of organic fluorine in some toxic tropical plants. New Phytol. 1972, 71, 855–871. [Google Scholar] [CrossRef]
  9. Wang, J.; Sánchez-Roselló, M.; Aceña, J.L.; del Pozo, C.; Sorochinsky, A.E.; Fustero, S.; Soloshonok, V.A.; Liu, H. Fluorine in Pharmaceutical Industry: Fluorine-Containing Drugs Introduced to the Market in the Last Decade (2001–2011). Chem. Rev. 2014, 114, 2432–2506. [Google Scholar] [CrossRef]
  10. Johnson, B.M.; Shu, Y.-Z.; Zhuo, X.; Meanwell, N.A. Metabolic and Pharmaceutical Aspects of Fluorinated Compounds. J. Med. Chem. 2020, 63, 6315–6386. [Google Scholar] [CrossRef]
  11. Filler, R.; Saha, R. Fluorine in medicinal chemistry: A century of progress and a 60-year retrospective of selected highlights. Future Med. Chem. 2009, 1, 777–791. [Google Scholar] [CrossRef] [PubMed]
  12. Gillis, E.P.; Eastman, K.J.; Hill, M.D.; Donnelly, D.J.; Meanwell, N.A. Applications of Fluorine in Medicinal Chemistry. J. Med. Chem. 2015, 58, 8315–8359. [Google Scholar] [CrossRef] [PubMed]
  13. Vulpetti, A.; Dalvit, C. Fluorine local environment: From screening to drug design. Drug Discov. Today 2012, 17, 890–897. [Google Scholar] [CrossRef] [PubMed]
  14. Mei, H.; Remete, A.M.; Zou, Y.; Moriwaki, H.; Fustero, S.; Kiss, L.; Soloshonok, V.A.; Han, J. Fluorine-containing drugs approved by the FDA in 2019. Chin. Chem. Lett. 2020, 31, 2401–2413. [Google Scholar] [CrossRef]
  15. Han, J.; Remete, A.M.; Dobson, L.S.; Kiss, L.; Izawa, K.; Moriwaki, H.; Soloshonok, V.A.; O’Hagan, D. Next generation organofluorine containing blockbuster drugs. J. Fluor. Chem. 2020, 239, 109639. [Google Scholar]
  16. Kafarski, P.; Lejczak, B. Biological Activity of Aminophosphonic Acids. Phosphorus Sulfur Silicon Relat. Elem. 1991, 63, 193–215. [Google Scholar] [CrossRef]
  17. Engel, R. Phosphonates as analogues of natural phosphates. Chem. Rev. 1977, 77, 349–367. [Google Scholar] [CrossRef]
  18. Azéma, L.; Baron, R.; Ladame, S. Targeting Enzymes with Phosphonate-Based Inhibitors: Mimics of Tetrahedral Transition States and Stable Isosteric Analogues of Phosphates. Curr. Enzym. Inhib. 2006, 2, 61–72. [Google Scholar] [CrossRef]
  19. Kolodiazhnyi, O.I. Asymmetric synthesis of hydroxyphosphonates. Tetrahedron 2005, 16, 3295–3340. [Google Scholar] [CrossRef]
  20. Rádai, Z. α-Hydroxyphosphonates as versatile starting materials. Phosphorus Sulfur Silicon Relat. Elem. 2019, 194, 425–437. [Google Scholar] [CrossRef]
  21. Romanenko, V.D.; Kukhar, V.P. Fluorinated Phosphonates:  Synthesis and Biomedical Application. Chem. Rev. 2006, 106, 3868–3935. [Google Scholar] [CrossRef] [PubMed]
  22. Cytlak, T.; Kaźmierczak, M.; Skibińska, M.; Koroniak, H. Latest achievements in the preparation of fluorinated aminophosphonates and aminophosphonic acids. Phosphorus Sulfur Silicon Relat. Elem. 2017, 192, 602–620. [Google Scholar] [CrossRef]
  23. Kaźmierczak, M.; Kubicki, M.; Koroniak, H. Regioselective Fluorination of α-Hydroxy-β-aminophosphonates by Using PyFluor. Eur. J. Org. Chem. 2018, 2018, 3844–3852. [Google Scholar] [CrossRef]
  24. Kaźmierczak, M.; Dutkiewicz, G.; Cytlak, T. Application of α-Amino-β-fluorophosphonates in Construction of their Dipeptide Analogues. Synthesis 2020, 52, 2364–2372. [Google Scholar] [CrossRef]
  25. Kazmierczak, M.; Dutkiewicz, G.; Koroniak, H. Deoxyfluorinating reagents as tools for γ-amino-α-hydroxyphosphonates modification. Org. Biomol. Chem. 2022, 20, 5615–5623. [Google Scholar] [CrossRef]
  26. Kaźmierczak, M.; Koroniak, H. Efficient synthesis of dipeptide analogues of α-fluorinated β-aminophosphonates. Beilstein J. Org. Chem. 2020, 16, 756–762. [Google Scholar] [CrossRef]
  27. Elliott, T.S.; Slowey, A.; Ye, Y.; Conway, S.J. The use of phosphate bioisosteres in medicinal chemistry and chemical biology. MedChemComm 2012, 3, 735–751. [Google Scholar] [CrossRef]
  28. Turcheniuk, K.V.; Kukhar, V.P.; Röschenthaler, G.-V.; Aceña, J.L.; Soloshonok, V.A.; Sorochinsky, A.E. Recent advances in the synthesis of fluorinated aminophosphonates and aminophosphonic acids. RSC Adv. 2013, 3, 6693–6716. [Google Scholar] [CrossRef]
  29. Cui, P.; McCalmont, W.F.; Tomsig, J.L.; Lynch, K.R.; Macdonald, T.L. alpha- and beta-substituted phosphonate analogs of LPA as autotaxin inhibitors. Bioorg. Med. Chem. 2008, 16, 2212–2225. [Google Scholar] [CrossRef] [Green Version]
  30. Kaźmierczak, M.; Koroniak, M. DAST mediated preparation of β-fluoro-α-aminophosphonates. J. Fluor. Chem. 2012, 139, 23–27. [Google Scholar] [CrossRef]
  31. Harnden, M.R.; Parkin, A.; Parratt, M.J.; Perkins, R.M. Novel acyclonucleotides: Synthesis and antiviral activity of alkenylphosphonic acid derivatives of purines and a pyrimidine. J. Med. Chem. 1993, 36, 1343–1355. [Google Scholar] [CrossRef]
  32. Lazrek, H.B.; Rochdi, A.; Khaider, H.; Barascut, J.L.; Imbach, J.L.; Balzarini, J.; Witvrouw, M.; Pannecouque, C.; De Clercq, E. Synthesis of (Z) and (E) α-alkenyl phosphonic acid derivatives of purines and pyrimidines. Tetrahedron 1998, 54, 3807–3816. [Google Scholar] [CrossRef]
  33. Thomas, A.A.; Sharpless, K.B. The Catalytic Asymmetric Aminohydroxylation of Unsaturated Phosphonates. J. Org. Chem. 1999, 64, 8379–8385. [Google Scholar] [CrossRef] [PubMed]
  34. Hammond, G.B.; De Mendonca, D.J. An efficient synthesis of (Z)-γ-fluoroallylphosphonates using a base-promoted deconjugation of (E)-γ-fluorovinylphosphonates, and its utility as fluoroolefin-containing building block. J. Fluor. Chem. 2000, 102, 189–197. [Google Scholar] [CrossRef]
  35. Pajkert, R.; Koroniak, H. Simple synthesis of some unsaturated 1,1-difluoro-2-hydroxyethylphosphonates. J. Fluor. Chem. 2007, 128, 1260–1263. [Google Scholar] [CrossRef]
  36. Cherkupally, P.; Slazhnev, A.; Beier, P. Stereoselective Synthesis of (E)-alpha-Fluorovinylphosphonates from alpha, alpha-Difluorophosphonates. Synlett 2011, 2011, 331–334. [Google Scholar]
  37. Kaźmierczak, M.; Kubicki, M.; Koroniak, H. Preparation and characterization of α-fluorinated-γ-aminophosphonates. J. Fluor. Chem. 2014, 167, 128–134. [Google Scholar] [CrossRef]
  38. Panigrahi, K.; Applegate, G.A.; Malik, G.; Berkowitz, D.B. Combining a Clostridial Enzyme Exhibiting Unusual Active Site Plasticity with a Remarkably Facile Sigmatropic Rearrangement: Rapid, Stereocontrolled Entry into Densely Functionalized Fluorinated Phosphonates for Chemical Biology. J. Am. Chem. Soc. 2015, 137, 3600–3609. [Google Scholar] [CrossRef] [Green Version]
  39. Rádai, Z.; Keglevich, G. Synthesis and Reactions of α-Hydroxyphosphonates. Molecules 2018, 23, 1493. [Google Scholar] [CrossRef] [Green Version]
  40. Szalai, Z.; Keglevich, G. Tetraalkyl Hydroxymethylene-bisphosphonate and Dialkyl 1-Diphenylphosphinoyl-1-hydroxy-ethylphosphonate Derivatives by the Pudovik Reaction and Their Rearranged Products. Molecules 2021, 26, 7575. [Google Scholar] [CrossRef]
  41. Kobayashi, K.; Tanaka, K.; Kogen, H. Recent topics of the natural product synthesis by Horner–Wadsworth–Emmons reaction. Tetrahedron Lett. 2018, 59, 568–582. [Google Scholar] [CrossRef]
  42. Roman, D.; Sauer, M.; Beemelmanns, C. Applications of the Horner–Wadsworth–Emmons Olefination in Modern Natural Product Synthesis. Synthesis 2021, 53, 2713–2739. [Google Scholar]
  43. Nakao, M.; Tanaka, K.; Kitaike, S.; Sano, S. Synthesis of Fluorine-Containing Analogues of 1-Lysoglycerophospholipids via Horner–Wadsworth–Emmons Reaction. Synthesis 2017, 49, 3654–3661. [Google Scholar]
  44. Wójtowicz-Rajchel, H. Synthesis and applications of fluorinated nucleoside analogues. J. Fluor. Chem. 2012, 143, 11–48. [Google Scholar] [CrossRef]
  45. Edmondson, S.D.; Wei, L.; Xu, J.; Shang, J.; Xu, S.; Pang, J.; Chaudhary, A.; Dean, D.C.; He, H.; Leiting, B.; et al. Fluoroolefins as amide bond mimics in dipeptidyl peptidase IV inhibitors. Bioorg. Med. Chem. Lett. 2008, 18, 2409–2413. [Google Scholar] [CrossRef]
  46. Bilska-Markowska, M.; Patyk-Kaźmierczak, E.; Lusina, A. Synthesis of Fluorinated Amides Starting from Carbohydrates Based on the Claisen Rearrangement. Eur. J. Org. Chem. 2022, 2022, e202101378. [Google Scholar] [CrossRef]
  47. Boysen, M.M.K. Carbohydrates as Synthetic Tools in Organic Chemistry. Chem. Eur. J. 2007, 13, 8648–8659. [Google Scholar] [CrossRef]
  48. Marciniak, B.; Bilska-Markowska, M.; Grzeszczuk, M.; Rapp, M.; Koroniak, H. A convenient synthesis of fluorinated vinyl ethers derived from di-O-isopropylidenehexofuranose and O-isopropylidenepentofuranose. J. Fluor. Chem. 2014, 167, 143–151. [Google Scholar] [CrossRef]
  49. Bilska-Markowska, M.; Szwajca, A.; Marciniak, B. Design, properties and applications of fluorinated and fluoroalkylated N-containing monosaccharides and their analogues. J. Fluor. Chem. 2019, 227, 109364. [Google Scholar] [CrossRef]
  50. Uhrig, M.L.; Lantaño, B.; Postigo, A. Synthetic strategies for fluorination of carbohydrates. Org. Biomol. Chem. 2019, 17, 5173–5189. [Google Scholar] [CrossRef]
  51. Linclau, B.; Ardá, A.; Reichardt, N.-C.; Sollogoub, M.; Unione, L.; Vincent, S.P.; Jiménez-Barbero, J. Fluorinated carbohydrates as chemical probes for molecular recognition studies. Current status and perspectives. Chem. Soc. Rev. 2020, 49, 3863–3888. [Google Scholar] [CrossRef] [PubMed]
  52. Shivatare, S.S.; Wong, C.-H. Synthetic Carbohydrate Chemistry and Translational Medicine. J. Org. Chem. 2020, 85, 15780–15800. [Google Scholar] [CrossRef] [PubMed]
  53. Lee, A.S.-Y.; Lin, M.-C.; Wang, S.-H.; Lin, L.-S. A Mild Isomerization Reaction for β,γ-Unsaturated Ketone to α,β-Unsaturated Ketone. J. Chin. Chem. Soc. 2004, 51, 371–376. [Google Scholar] [CrossRef]
  54. Fustero, S.; Moscardó, J.; Sánchez-Roselló, M.; Flores, S.; Guerola, M.; Pozo, C.D. Organocatalytic enantioselective synthesis of quinolizidine alkaloids (+)-myrtine, (−)-lupinine, and (+)-epiepiquinamide. Tetrahedron 2011, 67, 7412–7417. [Google Scholar] [CrossRef]
  55. Li, Y.; Li, K.; Wu, Y.; Ma, Q.; Lei, X. Facile synthesis of fluorovinyl-containing lactams via ring-closing metathesis of N-substituted 2-fluoroallylamides. Tetrahedron 2016, 72, 4845–4853. [Google Scholar] [CrossRef]
  56. Wang, H. Horner-Wadsworth-Emmons Olefination. In Comprehensive Organic Name Reactions and Reagents; Wiley: Hoboken, NJ, USA, 2010; pp. 1484–1490. [Google Scholar]
  57. 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 functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  58. Petersson, G.A.; Bennett, A.; Tensfeldt, T.G.; Al-Laham, M.A.; Shirley, W.A.; Mantzaris, J. A complete basis set model chemistry. I. The total energies of closed-shell atoms and hydrides of the first-row elements. J. Chem. Phys. 1988, 89, 2193–2218. [Google Scholar] [CrossRef]
  59. Zeng, Q.; Jones, M.R.; Brooks, B.R. Absolute and relative pKa predictions via a DFT approach applied to the SAMPL6 blind challenge. J. Comput.-Aided Mol. Des. 2018, 32, 1179–1189. [Google Scholar] [CrossRef]
  60. Foresman, J.B.; Frisch, A.E. Exploring Chemistry with Electronic Structure Methods, 3rd ed.; Gaussian, Inc.: Wallingford, CT, USA, 2015. [Google Scholar]
  61. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  62. Frisch, M.J.T.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision A.1; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  63. Liptak, M.D.; Shields, G.C. Accurate pKa Calculations for Carboxylic Acids Using Complete Basis Set and Gaussian-n Models Combined with CPCM Continuum Solvation Methods. J. Am. Chem. Soc. 2001, 123, 7314–7319. [Google Scholar] [CrossRef]
  64. Bordwell pKa Table. Available online: https://organicchemistrydata.org/hansreich/resources/pka/#ka-water (accessed on 21 July 2022).
  65. Leung, P.S.-W.; Teng, Y.; Toy, P.H. Rasta Resin-PPh3 and its Use in Chromatography-Free Wittig Reactions. Synlett 2010, 2010, 1997–2001. [Google Scholar]
  66. Payne, J.E.; Bonnefous, C.; Hassig, C.A.; Symons, K.T.; Guo, X.; Nguyen, P.-M.; Annable, T.; Wash, P.L.; Hoffman, T.Z.; Rao, T.S.; et al. Identification of KD5170: A novel mercaptoketone-based histone deacetylase inhibitor. Bioorg. Med. Chem. Lett. 2008, 18, 6093–6096. [Google Scholar] [CrossRef]
  67. Song, H.; Wu, D.; Mazunin, D.; Liu, S.M.; Sato, Y.; Broguiere, N.; Zenobi-Wong, M.; Bode, J.W. Post-Assembly Photomasking of Potassium Acyltrifluoroborates (KATs) for Two-Photon 3D Patterning of PEG-Hydrogels. Helv. Chim. Acta 2020, 103, e2000172. [Google Scholar] [CrossRef]
  68. Ortuzar, N.; Karu, K.; Presa, D.; Morais, G.R.; Sheldrake, H.M.; Shnyder, S.D.; Barnieh, F.M.; Loadman, P.M.; Patterson, L.H.; Pors, K.; et al. Probing cytochrome P450 (CYP) bioactivation with chloromethylindoline bioprecursors derived from the duocarmycin family of compounds. Bioorg. Med. Chem. 2021, 40, 116167. [Google Scholar] [CrossRef]
  69. Xie, M.; Berges, D.A.; Robins, M.J. Efficient “Dehomologation” of Di-O-isopropylidenehexofuranose Derivatives To Give O-Isopropylidenepentofuranoses by Sequential Treatment with Periodic Acid in Ethyl Acetate and Sodium Borohydride. J. Org. Chem. 1996, 61, 5178–5179. [Google Scholar] [CrossRef]
  70. Davis, F.A.; Kasu, P.V.N.; Sundarababu, G.; Qi, H. Nonracemic α-Fluoro Aldehydes:  Asymmetric Synthesis of 4-Deoxy-4-fluoro-d-arabinopyranose. J. Org. Chem. 1997, 62, 7546–7547. [Google Scholar] [CrossRef]
  71. Allan, A.L.; Gladstone, P.L.; Price, M.L.P.; Hopkins, S.A.; Juarez, J.C.; Doñate, F.; Ternansky, R.J.; Shaw, D.E.; Ganem, B.; Li, Y.; et al. Synthesis and Evaluation of Multisubstrate Bicyclic Pyrimidine Nucleoside Inhibitors of Human Thymidine Phosphorylase. J. Med. Chem. 2006, 49, 7807–7815. [Google Scholar] [CrossRef]
Figure 1. Structures of Fosfomycin 1 and Phosphatidylcholine 2.
Figure 1. Structures of Fosfomycin 1 and Phosphatidylcholine 2.
Molecules 27 05404 g001
Figure 2. Examples of synthetic hydroxyphosphonates [19].
Figure 2. Examples of synthetic hydroxyphosphonates [19].
Molecules 27 05404 g002
Scheme 1. Synthetic routes to: (a) α-hydroxyphosphonates, (b) unsaturated phosphonates.
Scheme 1. Synthetic routes to: (a) α-hydroxyphosphonates, (b) unsaturated phosphonates.
Molecules 27 05404 sch001
Scheme 2. Synthetic strategy towards 14–15, 20–21.
Scheme 2. Synthetic strategy towards 14–15, 20–21.
Molecules 27 05404 sch002
Scheme 3. Synthesis of 14.
Scheme 3. Synthesis of 14.
Molecules 27 05404 sch003
Scheme 4. Synthesis of 15.
Scheme 4. Synthesis of 15.
Molecules 27 05404 sch004
Figure 3. Felkin-Ahn model.
Figure 3. Felkin-Ahn model.
Molecules 27 05404 g003
Scheme 5. Synthesis of 20.
Scheme 5. Synthesis of 20.
Molecules 27 05404 sch005
Scheme 6. Synthesis of 21.
Scheme 6. Synthesis of 21.
Molecules 27 05404 sch006
Figure 4. Thermodynamic cycle used for pKa calculation.
Figure 4. Thermodynamic cycle used for pKa calculation.
Molecules 27 05404 g004
Table 1. Values of known experimental pKa, calculated pKa and calculated pKa with included obtained for Equation (1) parameters A and B.
Table 1. Values of known experimental pKa, calculated pKa and calculated pKa with included obtained for Equation (1) parameters A and B.
Exp. pKa [64]Calc. pKa
Formic Acid3.83.7
Trifluoroacetic Acid0.20.8
Dimethyl Malonate13.511.9
Acetone20.019.7
Chloroacetone16.517.9
2-((2,3,3,3-tetrafluoroprop-1-en-1-yl)oxy)but-3-enal 13-13.3
α-trifluoromethyl-γ,δ-unsaturated aldehyde 19b-15.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bilska-Markowska, M.; Jankowski, W.; Hoffmann, M.; Kaźmierczak, M. Design and Synthesis of New α-hydroxy β-fluoro/β-trifluoromethyl and Unsaturated Phosphonates from Carbohydrate-Derived Building Blocks via Pudovik and Horner–Wadsworth–Emmons Reactions. Molecules 2022, 27, 5404. https://doi.org/10.3390/molecules27175404

AMA Style

Bilska-Markowska M, Jankowski W, Hoffmann M, Kaźmierczak M. Design and Synthesis of New α-hydroxy β-fluoro/β-trifluoromethyl and Unsaturated Phosphonates from Carbohydrate-Derived Building Blocks via Pudovik and Horner–Wadsworth–Emmons Reactions. Molecules. 2022; 27(17):5404. https://doi.org/10.3390/molecules27175404

Chicago/Turabian Style

Bilska-Markowska, Monika, Wojciech Jankowski, Marcin Hoffmann, and Marcin Kaźmierczak. 2022. "Design and Synthesis of New α-hydroxy β-fluoro/β-trifluoromethyl and Unsaturated Phosphonates from Carbohydrate-Derived Building Blocks via Pudovik and Horner–Wadsworth–Emmons Reactions" Molecules 27, no. 17: 5404. https://doi.org/10.3390/molecules27175404

Article Metrics

Back to TopTop