Next Article in Journal
trans-Dihydroxo[5,10,15,20-tetrakis(3-pyridinium)porphyrinato]tin(IV) Nitrate
Previous Article in Journal
Synthesis of a New Bichalcone via Suzuki–Miyaura Coupling Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

Diisopropyl (E)-(1-hydroxy-3-phenylallyl)phosphonate

EaStCHEM School of Chemistry, University of St Andrews, North Haugh, St Andrews, Fife KY16 9ST, UK
*
Author to whom correspondence should be addressed.
Molbank 2025, 2025(2), M2013; https://doi.org/10.3390/M2013 (registering DOI)
Submission received: 12 May 2025 / Revised: 21 May 2025 / Accepted: 23 May 2025 / Published: 27 May 2025
(This article belongs to the Section Structure Determination)

Abstract

:
We report the synthesis and the molecular structure as determined by single-crystal X-ray diffraction of diisopropyl (E)-(1-hydroxy-3-phenylallyl)phosphonate. The compound was fully characterised by 1H, 13C, and 31P NMR spectroscopy, IR spectroscopy, and Mass Spectrometry.

1. Introduction

Traditionally, α-ketophosphonates can be easily prepared from acyl chlorides and trialkyl phosphites [1]. These can be converted to α-hydroxyphosphonates which are famous for their bioactive properties and versatility as starting materials [2,3]. Developed in the 1950s, the Pudovik reaction [4] involves the base-promoted nucleophilic addition of the P−H bond to unsaturated systems such as C=N, C=C, C≡C, or C=O bonds [5,6,7,8]. Compound 1 can be prepared by various methods. Racemic 1 can be prepared by reacting trans-cinnamaldeyde with diisopropyl phosphonate in the presence of potassium or caesium fluoride affording the racemic α-hydroxyphosphonate 1 [9,10,11,12]. Using chiral catalysts, diisopropyl cinnamoylphosphonate can be reduced to the α-hydroxyphosphonate 1 affording non-racemic products. The (S)-enantiomer of 1 has also been isolated in 75% yield with a 56% ee (enantiomeric excess) through the oxazaborolidine-catalysed enantioselective reduction of the diisoproply cinnamoylphosphonate [13]. This was improved several years later through the ruthenium(II) catalysed enantioselective transfer hydrogenation of α-ketophosphonates affording (R)-1 in a 78% yield with a 97% ee [14]. Other methods for preparing α-hydroxphosphonates include palladium catalysed asymmetric hydrogenation of the parent α-ketophosphonates [15] or copper-catalysed coupling of alcohols with diaryl/alkyl phosphonates [16].

2. Results and Discussion

2.1. Synthesis and Spectroscopy

The ‘Green’ synthesis of α-hydroxphosphonates was reported recently by utilising catalytic triethylamine to facilitate the Pudovik reaction between aldehydes and dialkyl phosphonates [17]. A similar method was reported by Arseniyadis and Kaïm et al. with the modification of dilution with chloroform afterwards to facilitate crystallisation [18]. These methods were adapted to prepare racemic 1 from trans-cinnamaldehyde and diisopropyl phosphonate (with stoichiometric triethylamine) in a 55% yield after recrystallisation (Scheme 1).
The High Resolution Mass Spectrometry (HRMS) data show two major peaks in the spectrum (Figure S15). The m/z peak is at 619.2555 amu which is indicative of 2M+Na, and the other peak at 321.1225 amu is indicative of M+Na (where M = 1). The peaks appear as the sodium adducts due to the analyte used in the electrospray ionisation (ESI+).
The 31P{1H} NMR spectrum shows a sharp singlet at δP 20.2 ppm (Figure S9). The 1H NMR spectrum displays a set of multiplets in the aromatic region corresponding to the phenyl group (δH 7.43−7.22 ppm, H-7-9) (Figure 1 and Figures S1−S5). The signal for the hydroxyl group is apparent using d6-DMSO as the NMR solvent at δH 5.86 ppm. The signal is a doublet of doublets due to coupling with the phosphorus (3JHP 13.1 Hz) and the hydrogen atom on the α-carbon, H-3 with 3JHH 5.9 Hz. The two alkene hydrogen atoms appear as two separate doublets of doublets of doublets at δH 6.69 and 6.28 ppm, H-5 and H-4, respectively. The largest J-coupling of 15.9 Hz is common to both signals as expected for an (E)-alkene [19]. The other couplings from H-5 (δH 6.69 ppm) are 3JHP 6.0 Hz and 3JHH 4.8 Hz to H-3, and from H-4 (δH 6.28 ppm) are 4JHP 5.0 Hz and 4JHH 1.8 Hz to H-3.
The signal from the hydrogen atom on the α-carbon (H-3) appears as a doublet of doublet of doublet of doublets (dddd) at δH 4.49 ppm, which collapses to a doublet of doublet of doublets (ddd) in the 1H{31P} NMR spectrum, although due to similar 3JHH coupling between the neighbouring hydrogen atoms it appears as a pseudo-doublet of triplets. Due to the stereocentre, the CH groups (H-2 and H-B) of the two isopropoxy groups are inequivalent (diastereotopic) causing them to appear at very slightly different chemical shifts. The 1H NMR spectrum shows a complex multiplet at δH 4.66−4.54 ppm, however the 1H{31P} spectrum shows two closely overlapping heptets (3JHH 6.1 Hz) at δH 4.61 and 4.60 ppm (see Figure S6).
The 13C DEPTQ NMR spectrum (see Figures S7 and S8) shows the aromatic signals (C-6−C-9) and the alkene signals (C-4 and C-5) ranging δC 136.9−127.7 ppm with C-4, C-5, and C-6 appearing as doublets owing to J-coupling with 31P with coupling constants of 3.4, 12.9 and 2.9 Hz, respectively. It is well established that the magnitude of 2JCP coupling is often much lower than 3JCP coupling when the phosphorus lone pair is no longer available [20]. In the same vein, C-3 shows a very large 1JCP of 165.5 Hz, which is typical of tetravalent phosphorus atoms. Similarly to H-2 and H-B having different chemical shifts, C-2 and C-B also have slightly different chemical shifts and appear as two distinct doublets at δC 70.8 and 70.6 ppm, with 2JCP 7.2 and 7.0 Hz, respectively.

2.2. Structural Study

Crystals of 1 that were of suitable quality for single crystal X-ray diffraction were grown from the vapour diffusion of petroleum ether into a solution of 1 in dichloromethane. The structure of racemic 1 was confirmed from these crystals which crystallised in a centrosymmetric space group (P21/c) (Figure 2).
The phosphorus adopts a distorted tetrahedral geometry with angles ranging 101.02(7)° to 115.50(7)° (Table 1). The bonds are all typical length, P1−O1 is 1.472(1) Å, clearly showing the P=O character, with the other P−O bonds being 1.566(1) and 1.574(1) Å [21]. The styryl moiety is disordered over two orientations, which are ~180° rotated around C7–C8 with a small rock around P1–C7 (Figure S16).
The hydroxyl group, in both disordered positions, forms a hydrogen bond to O1 of adjacent molecules (H···O 1.74(2) and 1.78(2) Å, O−H···O 171(6) and 173(6)°, for H7A and B, respectively), giving C(5) chains propagating along [0 0 1] (Figure 3). These chains further pack together into sheets across [0 0 1] through a mixture of weaker aromatic C−H···O interactions (H···O 2.7409(12)–3.3278(12) Å) and edge to face C−H···π interactions (H···centroid 3.192(4) and 2.709(5) Å, for each disordered part, respectively) between adjacent styrene groups. These C-H···π interactions are at or beyond the conventional van der Waals limit, but such interactions have been suggested to be effective at distances beyond this value [22].

3. Materials and Methods

All synthetic manipulations were performed in air. Glassware was dried in an oven (ca. 110 °C) prior to use. Diisopropyl phosphonate was purchased from ThermoFisher and used as received. All other chemicals were used as provided from the laboratory inventory without further purification. The IR spectrum was recorded on a Perkin Elmer Spectrum Two instrument with a Deuterated Triglycine Sulphate (DTGS) detector and diamond Attenuated Total Reflectance (ATR) attachment (Bruker, Billerica, MA, USA). The High Resolution Mass Spectrometry (HMRS) data were acquired from the University of St Andrews Mass Spectrometry Service. All NMR spectra were recorded using a Bruker Avance III 400 MHz spectrometer at 20 °C (Bruker, Billerica, MA, USA). The 13C spectrum was recorded using the DEPTQ-135 pulse sequence with broadband proton decoupling. Tetramethylsilane was used as external standard for 1H and 13C NMR (δHC 0.00 ppm), and 85% aqueous H3PO4 was used as the external standard for 31P NMR (δP 0.00 ppm). Where possible, the residual solvent signal was used as a secondary reference (d6-DMSO, δH 2.50, δC 39.52 ppm). Spectra were analysed using the MestReNova software package (Santiago de Compostela, Spain) (version 14). The NMR numbering scheme for 1 is provided in Figure 1.

3.1. Synthesis and Characterisation of Diisopropyl (E)-(1-hydroxy-3-phenylallyl)phosphonate (1)

Diisopropyl phosphonate (1.66 g, 1.6 mL, 10 mmol) and trans-cinnamaldehyde (1.32 g, 1.3 mL, 10 mmol) were mixed. To this, triethylamine (2.22 g, 3.1 mL, 22 mmol) was added. The mixture was heated to 75 °C with continuous stirring for 10 h. The solution was cooled to ambient conditions upon which a solid precipitated from the mixture. This was collected via filtration and washed with ice-cold toluene (5 mL) and dried in vacuo to afford 1 (1.65 g, 55%). This was recrystallised from petroleum ether 40/60 affording analytically pure 1 as pale-yellow crystals.
1H NMR (400.3 MHz, d6-DMSO) δH 7.43−7.39 (2H, m, H-7), 7.37−7.32 (2H, m, H-8), 7.28−7.22 (1H, m, H-9), 6.69 (1H, ddd, 3JHH 15.9, 4JHP 5.0, 4JHH 1.8 Hz, H-5), 6.28 (1H, ddd, 3JHH 15.9, 3JHH 6.0, 3JHP 4.8 Hz, H-4) 5.86 (1H, dd, 3JHP 13.1, 3JHH 5.9 Hz, OH), 4.66−4.54 (2 × 1H, m, H-2 & H-B), 4.49 (1H, dddd~ddt, 2JHP 15.8, 3JHH 5.9, 4JHH 1.7 Hz, H-3), 1.28−1.18 (12H, m, H-1 & H-A). 1H{31P} NMR (400.3 MHz, d6-DMSO) δH 7.43−7.39 (m, H-7), 7.37−7.32 (m, H-8), 7.28−7.22 (m, H-9), 6.69 (dd, 3JHH 15.9, 4JHH 1.8 Hz, H-5), 6.28 (dd, 3JHH 15.9, 3JHH 6.0 Hz, H-4) 5.86 (d, 3JHH 6.0 Hz, OH), 4.61 (hept, 3JHH 6.2 Hz, H1/H1′), 4.60 (hept, 3JHH 6.1 Hz, H1/H1′), 4.49 (ddd ~dt, 3JHH 6.0, 4JHH 1.7 Hz, H-3), 1.28−1.18 (m, H-2 & H2′). 13C DEPTQ (100.6 MHz, d6-DMSO) δC 136.9 (d, 4JCP 2.9 Hz, qC-6), 131.0 (d, 3JCP 12.9 Hz, C-5), 129.2 (s, H-8), 128.1 (s, H-9), 126.9 (d, 2JCP 3.4 Hz, C-4), 126.7 (s, C-7), 70.8 (d, 2JCP 7.2 Hz, C-1/C-1′), 70.6 (d, 2JCP 7.0 Hz, C-2/C-B), 69.1 (d, 1JCP 165.5 Hz, C-3), 24.4 (br-s, C-1/C-A), 24.2 (d, 3JCP 5.0 Hz, C-1/C-A). 31P{1H} NMR (162.0 MHz, d6-DMSO) 20.2 (s). HRMS (ESI+) m/z (%) Calcd. for C15H23O4PNa 321.1232; found 321.1225 [M+Na]+ (45); Cacld. For C30H46O8P2Na 619.2566; found 619.2555 [2M+Na]+ (100). Infrared (IR) νmax (ATR/cm−1) 3285m (νO−H), 2981m (νC−H), 1680w (νC=C), 1449m (νC=C), 1377m (νP−C), 1220s (νP=O), 984vs (νP−O), 761s, 694s, 564s.

3.2. X-Ray Crystallography

X-ray diffraction data for compound 1 were collected at 100 K using a Rigaku MM-007HF High Brilliance RA generator/confocal optics with XtaLAB P200 diffractometer [Cu Kα radiation (λ = 1.54187 Å)]. Data were collected (using a calculated strategy) and processed (including correction for Lorentz, polarisation, and absorption) using CrysAlisPro [23]. The structure was solved by dual-space methods (SHELXT [24]) and refined by full-matrix least-squares against F2 (SHELXL-2019/3 [25]). Non-hydrogen atoms were refined anisotropically, and hydrogen atoms were refined using a riding model except for the hydrogen atoms on O7A and O7B which were located from the difference Fourier map and refined isotropically subject to a distance restraint and with Ueq riding on their parent atom. The styryl moiety is disordered (occupancies 51.6:48.4) over two orientations, which are ~180° rotated around C7–C8 with a small rock around P1–C7. The two orientations were modelled with various geometric and thermal restraints on both parts due to their high level of overlap. All calculations were performed using the Olex2 interface [26]. Selected crystallographic data: C15H23O4P, M = 298.30, monoclinic, a = 10.20384(14), b = 8.16198(8), c = 20.1399(2) Å, β = 102.0831(11)°, U = 1640.16(3) Å3, T = 100 K, space group P21/c (no. 14), Z = 4, 30669 reflections measured, 3386 unique (Rint = 0.0631), which were used in all calculations. The final R1 [I > 2σ(I)] was 0.0471 and wR2 (all data) was 0.1418. CCDC 2443736 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/structures.

Supplementary Materials

Figures S1−S13: NMR spectra of 1 (1H, 1H{31P}, 13C DEPTQ, 31P{1H}, HH COSY, HC HSQC, HC HMBC, HP HMBC) Figure S14: IR spectrum of 1; Figure S15: HRMS spectrum of 1; Figure S16: View of the structure of 1 showing the disorder in the styryl moiety.

Author Contributions

All the required synthetic steps and analysis of NMR, IR, and HRMS data were carried out by A.I.C. The X-ray data were obtained and solved by A.P.M. and D.B.C. The study was designed by B.A.C. The manuscript was written by B.A.C. with all other authors contributing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

CCDC 2443736 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/structures.

Acknowledgments

The authors express gratitude to the University of St Andrews School of Chemistry for the use of their laboratory facilities and provision of materials.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations were used:
NMR Nuclear Magnetic Resonance
HRMS High Resolution Mass Spectrometry
DMSO dimethyl sulfoxide
ee enantiomeric excess
vsVery strong
mMedium
wWeak

References

  1. Benech, J.M.; Coindet, M.; El Manouni, D.; Leroux, Y. Synthesis of New α-Ketophosphonates. Phosphours Sulfur Silicon 1997, 123, 377–383. [Google Scholar] [CrossRef]
  2. Rádai, Z. α-Hydroxyphosphonates as versatile starting materials. Phosphorus Sulfur Silicon Relat. Elem. 2019, 194, 425–437. [Google Scholar] [CrossRef]
  3. Rádai, Z.; Keglevich, G. Synthesis and Reactions of α-Hydroxyphosphonates. Molecules 2018, 23, 1493. [Google Scholar] [CrossRef]
  4. Pudovik, A.N. Addition of dialkyl phosphites to unsaturated compounds. A new method of synthesis of β-keto phosphonic and unsaturated α-hydroxyphosphonic esters. Doklady Akad. Nauk. SSSR 1950, 73, 449. [Google Scholar]
  5. Abell, J.P.; Yamamoto, H. Catalytic Enantioselective Pudovik Reaction of Aldehydes and Aldimines with Tethered Bis(8-quinolinato) (TBOx) Aluminum Complex. J. Am. Chem. Soc. 2008, 130, 10521–10523. [Google Scholar] [CrossRef]
  6. Fener, B.E.; Schuler, P.; Ueberschaar, N.; Bellstedt, P.; Görls, H.; Krieck, S.; Westerhausen, M. Scope and Limitations of the s-Block Metal-Mediated Pudovik Reaction. Chem. Eur. J. 2020, 26, 7235–7243. [Google Scholar] [CrossRef] [PubMed]
  7. Yang, J.; Qian, D.-W.; Yang, S.-D. Lewis acid-catalysed Pudoivk reaction−phospha-Brook rearrangement sequence to access phosphoric esters. Beilsteon J. Org. Chem. 2022, 18, 1188–1194. [Google Scholar] [CrossRef]
  8. Yokomatsu, T.; Yamagishi, T.; Shibuya, S. Enantioselective synthesis of α-hydroxyphosphonates through asymmetric Pudovik reactions with chiral lanthanoid and titanium alkoxides. J. Chem. Soc. Perkin Trans. 1997, 1, 1527–1534. [Google Scholar] [CrossRef]
  9. Texier-Boullet, F.; Foucaud, A. A Convenient Synthesis of Dialkyl 1-Hydroxyalkanephosphonates using Potassium or Caesium Fluoride without Solvent. Synthesis 1982, 2, 165–166. [Google Scholar] [CrossRef]
  10. Smith, S.R.; Leckie, S.M.; Holmes, R.; Douglas, J.; Fallan, C.; Shapland, P.; Pryde, D.; Slawin, A.M.Z.; Smith, A.D. α-Ketophosphonates as Ester Surrogates: Isothiourea-Catalysed Asymmetric Diester and Lactone Synthesis. Org. Lett. 2014, 16, 2506–2509. [Google Scholar] [CrossRef]
  11. Öhler, E.; Kanzler, S. Synthesis of Phosphonic Acids Related to The Antibiotic Fosmidomycin From Allylic α- and γ- Hydroxyphosphonates. Phosphorus Sulfur Silicon Relat. Elem. 1996, 112, 71–74. [Google Scholar] [CrossRef]
  12. Leckie, S.M.; Fallan, C.; Taylor, J.E.; Brown, T.B.; Pryde, D.; Lébl, T.; Slawin, A.M.Z.; Smith, A.D. Enantiselective NHC-Catalysed Formal [4+2] Cycloaddition of Alkylarylketenes with β,γ-Unsaturated α-Ketophosphonates. Synlett 2013, 24, 1243–1249. [Google Scholar] [CrossRef]
  13. Meier, C.; Laux, W.H.G.; Bats, J.W. Asymmetric Synthesis of Chiral, Nonracemic Dialkyl α-Hydroxyarylmethyl- and α-, β- and γ-Hydroxyalkylphosphonates from Keto Phosphonates. Liebigs Ann. Recl. 1995, 11, 1936–1979. [Google Scholar] [CrossRef]
  14. Plouard, P.; Elmerich, U.; Hariri, M.; Loiseau, S.; Clarion, L.; Pirat, J.L.; Echeverria, P.-G.; Ayad, T.; Virieux, D. Ru(II)-Catalysed Enantioselective Transfer Hydrogenation of α-Ketophosphonates: Straightforward Access to Valuable Chiral α-Hydroxy Phosphonates. J. Org. Chem. 2023, 88, 16661–16665. [Google Scholar] [CrossRef]
  15. Goulioukina, N.S.; Bondaernko, G.N.; Gavrilov, K.N.; Beletskaya, I.P. Asymmetric Hydrogenation of α-Keto Phosphonates with Chiral Palladium Catalysts. Eur. J. Org. Chem. 2009, 2009, 510–515. [Google Scholar] [CrossRef]
  16. Zhao, Z.; Xue, W.; Gao, Y.; Tang, G.; Zhao, Y. Copper-Catalysed Synthesis of α-Hyroxy Phosphonates from H-Phosphonates and Alcohols or Ethers. Chem. Asian J. 2013, 8, 713–716. [Google Scholar] [CrossRef]
  17. Keglevich, G.; Rádai, Z.; Kiss, N.Z. To date the greenest method for the preparation of α-hydroxyphosphonates from substituted bezaldehydes and dialkyl phosphites. Green. Process. Synth. 2017, 6, 197–201. [Google Scholar] [CrossRef]
  18. Kerim, M.D.; Katsina, T.; Cattoen, M.; Fincias, N.; Arseniyadis, S.; Kaïm, L.E. O-Allylated Pudovik and Passerini Adducts as Versatile Scaffolds for Product Diversification. J. Org. Chem. 2020, 85, 12514–12525. [Google Scholar] [CrossRef]
  19. McKay, A.P.; Cordes, D.B.; Smellie, I.A.; Chalmers, B.A. (E)-3-Mesityl-1-(2,3,4,5-tetramethylphenyl)prop-2-en-1-one. Molbank 2025, 2025, M1952. [Google Scholar] [CrossRef]
  20. Kühl, O. Phosphoros-31 NMR Spectroscopy: A Concise Introduction for the Synthetic Organic and Organometallic Chemist; Springer: Berlin/Heidelberg, Germany, 2008; pp. 7–22. [Google Scholar]
  21. Eslami, F.; Porayoubi, M.; Sabbaghi, F.; Dušek, M.; Baniyaghoob, S.; Skořepová, E. Database Survey of Single-and-Half Phosphorus–Oxygen Bonds in Salts with the C2PO2 Segment: Crystal Structure of [NH2C5H4NH][(C6H5)2P(O)(O)]·2H2O. Crystallogr. Rep. 2022, 67, 218–223. [Google Scholar] [CrossRef]
  22. Umezawa, Y.; Tsuboyama, S.; Honda, K.; Uzawa, J.; Nishio, M. CH/π Interaction in the Crystal Structure of Organic Compounds. A Database Study. Bull. Chem. Soc. Jpn. 1998, 71, 1207–1213. [Google Scholar] [CrossRef]
  23. CrysAlisPro, version 1.171.42.94a and 43.144a; Rigaku Oxford Diffraction, Rigaku Corporation: Tokyo, Japan, 2023.
  24. Sheldrick, G.M. SHELXT—Integrated space-group and crystal structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  25. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  26. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Scheme 1. The synthesis of 1.
Scheme 1. The synthesis of 1.
Molbank 2025 m2013 sch001
Figure 1. Partial 1H NMR spectrum (top) and 1H{31P} NMR spectrum (bottom) of 1 shown from 7.30−4.30 ppm with numbering. Acquired at 400.3 MHz in d6-DMSO at 298 K. The numbering scheme is shown. As the isopropoxy groups are diastereotopic, they have been given different labels.
Figure 1. Partial 1H NMR spectrum (top) and 1H{31P} NMR spectrum (bottom) of 1 shown from 7.30−4.30 ppm with numbering. Acquired at 400.3 MHz in d6-DMSO at 298 K. The numbering scheme is shown. As the isopropoxy groups are diastereotopic, they have been given different labels.
Molbank 2025 m2013 g001
Figure 2. The molecular structure of 1. The anisotropic displacement ellipsoids of non-hydrogen atoms are set at the 50% probability level. Minor component of disorder omitted for clarity.
Figure 2. The molecular structure of 1. The anisotropic displacement ellipsoids of non-hydrogen atoms are set at the 50% probability level. Minor component of disorder omitted for clarity.
Molbank 2025 m2013 g002
Figure 3. The hydrogen bonding chains in 1 shown running along [0 0 1]. The isopropoxy groups are shown as wireframe, and minor component of disorder and carbon-bound hydrogen atoms omitted for clarity.
Figure 3. The hydrogen bonding chains in 1 shown running along [0 0 1]. The isopropoxy groups are shown as wireframe, and minor component of disorder and carbon-bound hydrogen atoms omitted for clarity.
Molbank 2025 m2013 g003
Table 1. Selected bond lengths (Å) and angles (°) for 1.
Table 1. Selected bond lengths (Å) and angles (°) for 1.
P1−O11.472(1)C8A−C9A1.318(5)
P1−O21.566(1)C7−O7A1.381(6)
P1−O31.574(1)O1···H7A1.75(6)
P1−C71.822(2)
O1−P1−O2115.50(7)O2−P1−O3103.34(6)
O1−P1−O3113.98(7)C7−P1−O2101.02(7)
O1−P1−C7114.51(8)O1···H7A−O7A172(5)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Coker, A.I.; McKay, A.P.; Cordes, D.B.; Chalmers, B.A. Diisopropyl (E)-(1-hydroxy-3-phenylallyl)phosphonate. Molbank 2025, 2025, M2013. https://doi.org/10.3390/M2013

AMA Style

Coker AI, McKay AP, Cordes DB, Chalmers BA. Diisopropyl (E)-(1-hydroxy-3-phenylallyl)phosphonate. Molbank. 2025; 2025(2):M2013. https://doi.org/10.3390/M2013

Chicago/Turabian Style

Coker, Andy I., Aidan P. McKay, David B. Cordes, and Brian A. Chalmers. 2025. "Diisopropyl (E)-(1-hydroxy-3-phenylallyl)phosphonate" Molbank 2025, no. 2: M2013. https://doi.org/10.3390/M2013

APA Style

Coker, A. I., McKay, A. P., Cordes, D. B., & Chalmers, B. A. (2025). Diisopropyl (E)-(1-hydroxy-3-phenylallyl)phosphonate. Molbank, 2025(2), M2013. https://doi.org/10.3390/M2013

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop