 Open Access Article
 Open Access Article
Garrett L. Reinhard *a, 
Reed Dowlingac, 
Patrick Hewitt
*a, 
Reed Dowlingac, 
Patrick Hewitt ab, 
Vikas Varshney
ab, 
Vikas Varshney a and 
Davide L. Simone
a and 
Davide L. Simone *a
*a
aAir Force Research Laboratory, Polymer Branch (AFRL/RXNP), Wright-Patterson AFB, OH 45433-7750, USA. E-mail: davide.simone@us.af.mil; garrett.reinhard@us.af.mil
bAeroVironment, Inc., Dayton, OH 45432, USA
cUniversity of Dayton, Dayton, OH 45469, USA
First published on 20th June 2025
A convergent synthetic strategy to high-persistence length helicenes via a regioselective and scalable menthyloxycarbonato-[4]-helicene synthon accessed in 84% yield is reported. To demonstrate the utility of the [4]-helicene building block, bis(menthyloxycarbonato)-[11]-helicene diastereomers were prepared via palladium-mediated cross-couplings, followed by a Mallory-type photo-induced annulation reaction. Our synthetic strategy leverages Stille and Heck couplings to synthesize bis(aryl)ethene precursors that exhibit preferential formation of helical products via 6π-electrocyclization. [11]-Helicene product yields are enhanced from 7 to 42% by maintaining photoreaction temperatures above 40 °C limiting linear and cyclobutane byproduct formation while recrystallization yielded diastereo-enrichment. This work enables the end goal of embedding enantiopure hydroxy-terminated helicenes into polymer backbones to create chiroptical responsive strain sensors.
Our interests lie in exploring the viability of helicenes as sensors and their potential to racemize under the application of a mechanical force or temperature change while bonded within a polymer backbone. To that end, we wish to develop a versatile synthesis of end-terminal functional helicenes, whereupon inversion barriers can be tailored as a function of helicene length.14 Before embarking on developing a highly efficient synthesis of helicenes, we feel it important to review various synthetic methodologies that have been reported in the literature particularly toward the isolation of optically active (non-racemic) helicenes.
The synthesis of carbohelicenes follow five main mechanistic routes; the photocyclization, the Diels–Alder reaction, Friedel–Crafts-type reactions, metal catalyzed cyclizations, and radical cyclizations.1 The synthetic pathway focused upon in this work is the Mallory reaction, a photochemical–electrocyclization–elimination type reaction developed by Mallory et al. in which stilbenes are subjected to ultraviolet radiation forming dihydro-intermediates which subsequently undergo an (dihydro) elimination reaction in the presence of an oxidant, commonly a catalytic amount of elemental iodine.16 This synthesis was improved upon by Katz et al., in which a stoichiometric amount of iodine is used in the presence of propylene oxide in an air free environment.17 Of the synthetic mechanisms reviewed, the Katz photocyclization has been shown to have improved yields, limit the distribution of potential side products, all while needing much less intense conditions compared to the other common synthetic mechanisms.17 It is important to note that although metal-catalyzed cyclizations tend to provide an elegant way to selectively cyclize helicene precursors, the syntheses tend to require a large number of synthetic steps.18,19 To date, the longest helicene synthesized is a [16]-helicene that contains terminal triisopropylsilyl ether functionality, with an overall yield of less than 7%.20 The goal of this publication is to define a more efficient synthetic path to high persistence length helicenes derived from [4]-helicene building blocks.
The mechanism of the Mallory-type reaction is shown in Fig. 1, where a bis(aryl)ethene undergoes photo-induced ring closure to form either a [4]-helicene or its linear analogue. Bis(aryl)ethenes are conventionally referred via [x] + [y] terminology, such that x and y are the number of aromatic rings bridged by an alkene. As an example, Scheme 1 showcases a [2] + [1] precursor electrocyclizing to a [4]-helicene.21,22
During photocyclization, in addition to expected helical and linear products, there also exists the occurrence of an intermolecular [2 + 2] cycloaddition reaction, which yields cyclobutane based side products. The side products have a greater tendency of forming when the concentration of helicene precursor is sufficiently high to enable the bi-molecular photocylization to occur. It has been reported that the formation of these side products is minimized by limiting the concentration of bis(aryl)ethene species to 1 mM for each site undergoing photocyclization.23 This concentration limitation poses a challenge in scalable, high-yielding syntheses of helicenes, as chemists are then limited by the volume and efficiency of their photo-reaction vessel.
Developing a synthetic pathway to enable pure isolation of higher order enantiopure helicenes is highly desired for two reasons; (a) higher order helicenes possess enhanced chiroptical properties, and (b) it significantly minimizes their tendency to racemize due to a high inversion activation energy. Towards achieving this, we find that the design of the stilbene precursor has been shown to heavily influence the production of the helicene over their linear isomers.22 The precursors that take on the design of [4] + [1] or [4] + [1] + [4] give the highest yields of un-substituted [6]- and [11]-helicene, achieving 87% and 84% isolated yields respectively. Other precursor designs, such as a [1] + [2] + [1] result in a diminished yield of [6]-helical product at 60% yield, and in some other cases, such as the [3] + [2], the linear analogue is preferred.22
In this work, we synthesize [2] + [1] and [4] + [1] + [4] helicene precursors to allow us to efficiently isolate [4]-helicene and [11]-helicene diastereomers enables separation into enantiopure forms. The [4]-helicene that we are reporting is terminated with a bromine to enable the synthesis of other higher order helicenes, such as a [9]- via a [4] + [4], a [6]- via a [4] + [1] or the [11]-helicene via [4] + [1] + [4], etc.
(1R)-(+)-Menthyl chloroformate was selected as the chiral resolving agent to terminate the ends of the [11]-helicene and its precursors, enhancing solubility, enabling diastereomeric separation, and, upon removal, afford hydroxyl terminated helicenes that can be functionalized further to incorporate into polymer backbones. The first step to creating a [4] + [1] + [4] helicene precursor which is envisioned to provide improved long helicene yields, is through the efficient synthesis of a functional [4]-helicene building block, Scheme 1. Using commercially available 6-bromo-2-naphthol, (1R)-(+)-menthyl chloroformate was installed via a deprotonation of the hydroxyl group with triethylamine affording the (1R)-(+)-menthyl carbonate in high yielding, multi-decagram quantity batches.
Extension of the menthyl protected bromonapthyl building block to the [2] + [1] helicene precursor is achieved via subsequent palladium-catalyzed cross-coupling reactions, a Stille coupling followed by a chemoselective Heck coupling. The Stille couplings executed in this paper are modeled after syntheses found in the review by Farina.24 The Heck coupling conditions were modeled after improvements made by the Buchwald group.25
The synthesis of 2 made use of the Farina reaction conditions without major changes. Aqueous KF washes were avoided to reduce any risk of interaction between fluoride anions and the menthyl carbonate groups due to their susceptibility towards even mild nucleophiles. Instead, once the crude reaction mixtures were deposited onto a silica column, approximately 1–1.5 column volumes of pure hexanes was flushed through the crude material and stationary phase to remove most organotin impurities before switching to the main eluent. Further recrystallization after chromatography was required to remove residual organotin impurities in order to mitigate any potential side reactions in subsequent palladium-mediated cross couplings.
The Heck coupling of 2 with 1-bromo-4-iodobenzene was found to be chemo-selective at 50 °C, preferentially reacting with the iodo- instead of bromo-position owing to the greater ease of oxidative addition towards the weaker Ph–I bond. Reactions conducted below 50 °C yielded poor reaction progression, while increasing the temperature to 100 °C causes coupling at both halogenated positions and a total loss of chemo-selectivity.
As introduced previously, high concentrations of bis(aryl)ethenes in solution are to be avoided to reduce [2 + 2] cyclo-additions, so to improve scalability and mitigate these side products, a syringe pump (pseudo-high dilution) was utilized to add a stock solution of precursor 3 and iodine in THF at a rate in which reaction concentration was calculated to be at a constant 1 mM. Initial studies indicated via NMR that the cyclization of 3 to 4 fully converted within 2 hours, and so the rate of addition of precursor 3 was adjusted to match this conversion. The syringe pump method allowed for a substantial increase in scale, allowing for multi-gram quantities to be made in a single batch, as well as minimizing the potential of [2 + 2] side reactions. Without a syringe pump, using a 1 L photoreaction vessel limits scale of this reaction to yield 380 mg; with the syringe pump, an order of magnitude greater yields were achieved.
With compound 4 in hand, the synthesis of higher order helicenes is enabled, Scheme 2.
Following a similar Stille and this case a non-chemoselective Heck reaction sequence, 4 is converted to [4] + [1] + [4] helicene precursor 6 in a respectable yield. The photocyclization of 6 to the desired helicene product introduced several competing factors from a reaction optimization perspective. First, the photocyclization of 6 was designed at a 0.5 mM overall concentration, accounting for the two sites of conversion and to reduce the probability of [2 + 2] cyclo-additions. Temperature and time were two other variables that had a noticeable impact on the outcome of the reaction. Literature precedence indicates that lower temperatures favor the kinetic product (helicene) rather than the non-strained thermodynamic product (linear/pseudo-helical). During our initial attempts to photocyclize the [4] + [1] + [4], the photo-reaction vessel was chilled with an ethylene glycol solution to lower the reaction temperature to sub-ambient conditions with the ability to precisely control the temperature of the reaction solvent via an external thermocouple.
Equilibrating the reaction temperature at ∼14 °C, the photocyclization reaction yielded the helical product, with a final isolated yield of ∼3% after a total reaction time of 144 hours in the photoreactor. In order to further investigate the impact of reaction temperature on the nature of the photocyclization reaction, additional small scale exploratory vial reactions were performed at ambient temperature between 20 °C and 25 °C. The findings from these experiments indicated that at ambient temperatures faster reaction rates were observed, with full consumption of starting material in 53 hours, with no noticeable increase in the magnitude of impurity resonances as observed via 1H NMR. The vial scale reactions were scaled up to the 1 L photoreactor with the ethylene glycol chiller adjusted until the temperature of the reaction solvent was stabilized at 24 °C after the UV light source reached full power. As observed in the vial scale tests, the reaction appeared to proceed at an elevated rate compared to sub-ambient temperatures, however over-exposure to UV irradiation appeared to cause transformation of helicene product after the same 144 hours reaction period.
Thus, in order to balance the preferential formation of the kinetic helicene product while mitigating photo-driven product transformation arising from extended UV exposure, an intermediate reaction time of 91 hours was observed as the point of full conversion of the starting material at 24 °C. After a similar purification process to the previous large scale reaction, the final isolated yield of 7 was increased from ∼3% to ∼10%. Many unaccounted for aromatic resonances exist in the pre-work up material, as well as resonances in the 2–4 ppm region of the 1H NMR, which have been previously associated with protons in cyclobutane derivatives arising from undesirable [2 + 2] cycloadditions.26 This leads to our hypothesis that linear products [2 + 2] products, pseudo-helical, and various other products of the aforementioned species all compete with the formation of helical product.
Based on the observed trend of increasing temperature affording improved reaction yields, a final trial at the 1 L scale was performed at an increased reaction temperature of 44 °C using a hot water supply, measured to be 31 °C. The total residence time in the photoreactor in this trial was 66 hours with an isolated yield of 42%. A summary the relationship between isolated yield and temperature can be found in Fig. 2. The 1H NMR spectra of pre-worked up material was less complicated, with many aromatic resonances and resonances in the cyclobutane region previously observed in the sub-ambient and ambient temperature trials were no longer observed in the elevated temperature experiment. We postulate that the increase in temperature assists in the photoinduced trans to cis isomerization of 6 allowing for a greater probability of bond formation between the carbons adjacent to the double bond and subsequent oxidation of the dihydrophenanthrene-like intermediate to form the fused-ring backbone. In turn, this reduces the probability of intermolecular side reactions, such as the [2 + 2] cycloaddition, as there is less of a chance for reaction-inducing collisions before annulation occurs.
In our study, the isolated yields of 7 were noticeably lower than the similar [4] + [1] + [4] system shown by Martin and Baes in 1975, who reported an 84% isolated yield of [11]-carbohelicene upon photocyclization. It is evident that the photocyclization of 6 is improved by providing sufficient activation energy to increase the rate of trans to cis isomerization which is required before C–C bond formation can occur between the adjacent rings to form the helicene backbone. Avoiding the need to undergo trans to cis isomerization could indicate that formulating a synthetic route to the Z-bisarylethenes, as was done in Baes and Martin's work through a Wittig reaction, over the E-bisarylethenes typically produced via organometallic couplings, provides a predisposition to form helicene over other side products by reducing the energetic barrier required for annulation to the helical product. It should be noted that the Wittig route to helicene precursors is most useful for non-functional (carbohelicene) products as precursors containing more diverse chemical functionality (i.e. menthyl carbonate groups) would complicate desired Z-alkene formation.
Though direct studies on the effect of molecular orbital populations on the annulation products of bis(aryl)ethenes are not widely reported in the literature, there are still indications that manipulating the electronic structure of bis(aryl)ethenes could have significant impacts on the outcome of photocyclization of studied substrates. Matsuda et al. found that the introduction of electron-withdrawing groups to the C-5 and C-10 positions of [5]-helicene significantly suppressed the undesired over-annulation of the [5]-helicene by removing orbital degeneracy of lower UMOs (unoccupied molecular orbitals), stabilizing the C2-antisymmetric UMO, and making it such that the only transition is not suitable for conrotary cyclization, reducing the amount of over-annulated product in their cyano-substituted [5]-helicene.27 A similar affect may be occurring with the cyclization of 6, where the menthyl-carbonate groups could be stabilizing the symmetric UMO configuration such that the helical product is less favored for conrotary cyclization compared to other side products. Further computational modeling as well as a systematic study of substituent effects on the photocyclization product distribution could provide deeper insight into the specifics of the reaction and begin to establish a design framework for optimization of synthesizing specific structural isomers.
In alignment with our interest in obtaining enantiopure helicenes for future applications, diastereomeric separation of 7 has been achieved on an analytical scale LC-MS using a Waters XBridge BEH phenyl OBD column (2.1 × 50 mm with a 1.7 μm particle size) using 99.5![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) :
:![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) 0.5 MeCN
0.5 MeCN![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) :
:![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) toluene as mobile phase eluting at 0.050 mL min−1. The transfer of analytical scale conditions to preparative scale LC-MS using a number of different column compositions and mobile phases is currently being explored. Fortuitously, an interesting phenomenon was observed when recrystallizing a small sample (5 mg) of 7 using acetonitrile during small-scale purification development. Upon isolating the recrystallized solids, it was found that the crystals were diastereomerically enriched when characterized using 1H NMR and by CD spectroscopy shown in Fig. 3. This behavior could be used to assist in the synthesis of diastereomerically pure 7 at scale in the future by using diastereomerically pure crystals to seed growth at a large scale, potentially avoiding the use of preparative LC. The synthesis of this functionalized [11]-helicene enables the validation of spring-force models, the development of elastomeric materials for strain sensing materials, and postures the team to develop even longer order helicenes.
toluene as mobile phase eluting at 0.050 mL min−1. The transfer of analytical scale conditions to preparative scale LC-MS using a number of different column compositions and mobile phases is currently being explored. Fortuitously, an interesting phenomenon was observed when recrystallizing a small sample (5 mg) of 7 using acetonitrile during small-scale purification development. Upon isolating the recrystallized solids, it was found that the crystals were diastereomerically enriched when characterized using 1H NMR and by CD spectroscopy shown in Fig. 3. This behavior could be used to assist in the synthesis of diastereomerically pure 7 at scale in the future by using diastereomerically pure crystals to seed growth at a large scale, potentially avoiding the use of preparative LC. The synthesis of this functionalized [11]-helicene enables the validation of spring-force models, the development of elastomeric materials for strain sensing materials, and postures the team to develop even longer order helicenes.
|  | ||
| Fig. 3 CD and NMR (protons 4 and 4′) of diastereomerically-enriched 7. Full 1H NMR spectrum of 7 available in ESI.† | ||
All synthesized materials were characterized via 1H and 13C NMR spectroscopy as well as single-crystal X-ray diffractometry, except for 6 in which suitable crystals could not be grown. High-resolution mass spectroscopy for all synthesized compounds determined using a Waters Corporation Xevo G3 QToF with electrospray ionization in positive mode (ES+) with solutions of concentrations ∼1 ng mL−1. A Bruker Ascend™ 400 MHz NMR spectrometer was used to collect 1H and 13C NMR spectra for all materials. X-ray intensity data was collected on a Rigaku XtaLAB Synergy-S, PhotonJet-i with a CCD detector using Cu Kα radiation (λ = 1.5406 Å). Images were interpreted and integrated with CrysAlisPRO and structures were solved using Olex2 with the ShelXT solution program and further refinement using ShelXL.28 Non-hydrogen atoms were refined anisotropically.
CCDC 2428292–2428297 contain the supplementary crystallographic data for this paper. These data are provided free of charge by the Cambridge Crystallographic Data Centre.
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) 195 reflections measured (8.496° ≤ 2ϕ ≤ 153.586°), 7620 independent (Rint = 0.0328, Rsigma = 0.0388). Final R1 = 0.0407 [I > = 2σ(I)] and wR2 = 0.1102 [all data].
195 reflections measured (8.496° ≤ 2ϕ ≤ 153.586°), 7620 independent (Rint = 0.0328, Rsigma = 0.0388). Final R1 = 0.0407 [I > = 2σ(I)] and wR2 = 0.1102 [all data].Single crystals of 5 were grown via the bulk recrystallization in hexanes to afford colorless, needle-like crystals. Crystals of 6 could not be grown sufficiently for single crystal XRD. Crystalline solids of 7 were collected via the bulk recrystallization in acetonitrile.
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) 976 reflections measured (3.668° ≤ 2ϕ ≤ 154.756°), 10
976 reflections measured (3.668° ≤ 2ϕ ≤ 154.756°), 10![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) 128 independent (Rint = 0.0197, Rsigma = 0.0108). Final R1 = 0.0774 [I > = 2σ(I)] and wR2 = 0.2301 [all data].
128 independent (Rint = 0.0197, Rsigma = 0.0108). Final R1 = 0.0774 [I > = 2σ(I)] and wR2 = 0.2301 [all data].Crystallographic analysis of compounds 4 and 7 confirmed the anticipated synthetic outcomes, helicene over linear, as well bond lengths and torsion angles that are common in these helical structures. In the inner helix, the average C–C bond length of 4 at 1.45 Å versus all other bond lengths (outer helix, terminal aromatic rings), deviates from an ideal aromatic bond of 1.39 Å. The same can be said of 7 and is attributed to the non-planar, strained conformation of helicenes relative to planar linear structural isomers. The solved crystal structure and packing of 4 and 7 can be found in Schemes 1 and 2 respectively.
Crystallographic parameters of 1, 2, 3 and 5 can be found in the ESI† of this manuscript.
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) :
:![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) hexanes on 10 wt% K2CO3
hexanes on 10 wt% K2CO3![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) :
:![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) silica) performed to afford puffy colorless solid. Recrystallization from heptane afforded colorless crystalline solid (33.431 g, 77% yield). 1H NMR (δ, CDCl3, 400 MHz): 0.860 (d, 3H), 0.913 (d, 1H), 0.956 (d, 6H), 1.097 (m, 1H), 1.171 (q, 1H), 1.522 (m, 2H), 1.718 (m, 2H), 2.080 (m, 1H), 2.202 (m, 1H), 4.656 (m, 1H), 5.340 (d, 1H), 5.866 (d, 1H), 6.865 (dd, 1H), 7.301 (dd, 1H), 7.621 (d, 1H), 7.649 (dd, 1H), 7.757 (m, 2H), 7.815 (d, 1H). 13C NMR (δ, CDCl3, 101 MHz): 16.389, 20.770, 22.004, 23.392, 26.259, 31.457, 34.075, 40.676, 47.061, 79.635, 114.385, 117.907, 121.021, 124.028, 126.122, 127.952, 129.526, 131.542, 133.383, 135.021, 136.696, 148.919, 153.410. ES+ calcd for C23H28NaO3: m/z 375.1936 observed: m/z 375.1930.
silica) performed to afford puffy colorless solid. Recrystallization from heptane afforded colorless crystalline solid (33.431 g, 77% yield). 1H NMR (δ, CDCl3, 400 MHz): 0.860 (d, 3H), 0.913 (d, 1H), 0.956 (d, 6H), 1.097 (m, 1H), 1.171 (q, 1H), 1.522 (m, 2H), 1.718 (m, 2H), 2.080 (m, 1H), 2.202 (m, 1H), 4.656 (m, 1H), 5.340 (d, 1H), 5.866 (d, 1H), 6.865 (dd, 1H), 7.301 (dd, 1H), 7.621 (d, 1H), 7.649 (dd, 1H), 7.757 (m, 2H), 7.815 (d, 1H). 13C NMR (δ, CDCl3, 101 MHz): 16.389, 20.770, 22.004, 23.392, 26.259, 31.457, 34.075, 40.676, 47.061, 79.635, 114.385, 117.907, 121.021, 124.028, 126.122, 127.952, 129.526, 131.542, 133.383, 135.021, 136.696, 148.919, 153.410. ES+ calcd for C23H28NaO3: m/z 375.1936 observed: m/z 375.1930.
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) :
:![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) hexanes) to afford colorless solid. Solids recrystallized from hexanes to afford colorless crystalline solid (310 mg, 69% yield). 1H NMR (δ, CDCl3, 400 MHz): 0.820 (d, 3H), 0.902 (t, 3H), 1.052 (m, 1H), 1.134 (q, 1H), 1.468 (m, 2H), 1.659 (m, 2H), 2.047 (m, 1H), 2.171 (m, 1H), 4.631 (td, 1H), 5.332 (d, 1H), 5.874 (d, 1H), 6.919 (dd, 1H), 7.459 (dd, 1H), 7.695 (dd, 2H), 7.759 (m, 4H), 7.883 (d, 1H), 8.911 (s, 1H), 9.024 (d, 1H). 13C NMR (δ, 101 MHz, CDCl3): 14.16. 16.44, 20.81, 22.04, 23.43, 26.31, 31.50, 31.62, 34.10, 40.73, 47.11, 76.73, 77.04, 77.25, 77.36, 79.73, 79.77, 114.53, 118.93, 120.14, 123.28, 126.54, 126.84, 127.04, 127.11, 127.22, 127.38, 127.83, 128.27, 128.86, 129.29, 130.31, 131.12, 133.25, 134.31, 135.58, 137.32, 148.81, 153.49. ES+ calcd for C31H32NaO3: m/z 475.2249 observed: m/z 475.2243.
hexanes) to afford colorless solid. Solids recrystallized from hexanes to afford colorless crystalline solid (310 mg, 69% yield). 1H NMR (δ, CDCl3, 400 MHz): 0.820 (d, 3H), 0.902 (t, 3H), 1.052 (m, 1H), 1.134 (q, 1H), 1.468 (m, 2H), 1.659 (m, 2H), 2.047 (m, 1H), 2.171 (m, 1H), 4.631 (td, 1H), 5.332 (d, 1H), 5.874 (d, 1H), 6.919 (dd, 1H), 7.459 (dd, 1H), 7.695 (dd, 2H), 7.759 (m, 4H), 7.883 (d, 1H), 8.911 (s, 1H), 9.024 (d, 1H). 13C NMR (δ, 101 MHz, CDCl3): 14.16. 16.44, 20.81, 22.04, 23.43, 26.31, 31.50, 31.62, 34.10, 40.73, 47.11, 76.73, 77.04, 77.25, 77.36, 79.73, 79.77, 114.53, 118.93, 120.14, 123.28, 126.54, 126.84, 127.04, 127.11, 127.22, 127.38, 127.83, 128.27, 128.86, 129.29, 130.31, 131.12, 133.25, 134.31, 135.58, 137.32, 148.81, 153.49. ES+ calcd for C31H32NaO3: m/z 475.2249 observed: m/z 475.2243.
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) :
:![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) 95 dioxane
95 dioxane![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) :
:![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) heptane) to afford powdery yellow solid (154 mg, 42% yield). 1H NMR (δ, CDCl3, 400 MHz): 0.880 (q, 8H), 0.972 (dd, 14H), 1.138 (m, 5H), 1.549 (m, 4H), 1.733 (m, 4H), 2.044 (m, 2H), 2.178 (m, 2H), 4.625 (sep, 2H), 6.001 (dt, 2H), 6.681 (d, 2H), 7.114 (m, 4H), 7.282 (m, 14H), 7.394 (d, 2H). 13C NMR (δ, 101 MHz, CD2Cl2): 17.48, 17.52, 21.85, 23.10, 24.70, 24.72, 27.59, 27.64, 31.01, 32.78, 35.42, 42.02, 48.49, 48.54, 80.41, 118.06, 118.11, 118.15, 123.98, 125.62, 125.94, 126.57, 126.77, 126.79, 127.26, 127.30, 127.40, 127.49, 127.67, 127.69, 127.78, 127.83, 127.89, 128.42, 131.25, 131.86, 132.82, 133.01, 133.59, 148.87, 154.49, 154.55. ES+ calcd for C68H62NaO6: m/z 997.4444 observed: m/z 997.4443.
heptane) to afford powdery yellow solid (154 mg, 42% yield). 1H NMR (δ, CDCl3, 400 MHz): 0.880 (q, 8H), 0.972 (dd, 14H), 1.138 (m, 5H), 1.549 (m, 4H), 1.733 (m, 4H), 2.044 (m, 2H), 2.178 (m, 2H), 4.625 (sep, 2H), 6.001 (dt, 2H), 6.681 (d, 2H), 7.114 (m, 4H), 7.282 (m, 14H), 7.394 (d, 2H). 13C NMR (δ, 101 MHz, CD2Cl2): 17.48, 17.52, 21.85, 23.10, 24.70, 24.72, 27.59, 27.64, 31.01, 32.78, 35.42, 42.02, 48.49, 48.54, 80.41, 118.06, 118.11, 118.15, 123.98, 125.62, 125.94, 126.57, 126.77, 126.79, 127.26, 127.30, 127.40, 127.49, 127.67, 127.69, 127.78, 127.83, 127.89, 128.42, 131.25, 131.86, 132.82, 133.01, 133.59, 148.87, 154.49, 154.55. ES+ calcd for C68H62NaO6: m/z 997.4444 observed: m/z 997.4443.
| Footnote | 
| † Electronic supplementary information (ESI) available: 1H and 13C NMR of compounds 1–7 along with crystal structures and parameters of 1–3 and 5. CCDC 2428292–2428297. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5ra02685g | 
| This journal is © The Royal Society of Chemistry 2025 |