• No results found

University of Groningen Design of new aggregates for catalysis Tosi, Filippo

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Design of new aggregates for catalysis Tosi, Filippo"

Copied!
19
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Design of new aggregates for catalysis

Tosi, Filippo

DOI:

10.33612/diss.107814277

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Tosi, F. (2019). Design of new aggregates for catalysis. Rijksuniversiteit Groningen. https://doi.org/10.33612/diss.107814277

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

Chapter 2

Reorganization from Metastable to

Stable

Aggregation

States

of

BINOL-derived Amphiphiles in Water

ABSTRACT: The synthesis and self-assembly behavior of newly designed

BINOL-based amphiphiles is presented. With minor structural modifications, the aggregation of these amphiphiles could be successfully tuned to form different types of assemblies in water, ranging from vesicles to cubic structures. Simple sonication induced rearrangement of different metastable aggregates into thermodynamically stable self-assembled nanotubes as observed by Cryo-TEM.

This chapter has been published as: Filippo Tosi, Marc C. A. Stuart,Hans Smit, Jiawen Chen, Ben L. Feringa, Langmuir 2019, 35, 11821−11828.

(3)

26

2

2.1 Introduction

The design and exploitation of novel amphiphiles is of paramount importance for various applications,[1–6] among which catalysis in water is an emerging field.[7–12] The use of surfactants to promote reactivity in water has proven to be particularly interesting,[10,13,14] and is reminiscent of the strategy employed by Nature where catalytic functions are embedded in compartmentalized systems. For example, in cellular membranes the lipid bilayer encapsulates organic substrates and generates a favorable environment for reactions to take place.[15] The most frequently used approach up to now has been to develop surfactants which self-assemble into micelles.[16–21] Due to their dynamic character, micellar reactors can enhance reaction rates and enantiomeric excesses under mild conditions with respect to most common organic solvents.[22–25]

The use of vesicular reactors[26–30] has been extensively explored in recent years. Compared to micelles, in selected examples vesicles have exhibited an even greater influence over reaction rate in water.[31,32] This suggests that mesoscopic differences in the morphology of the aggregates as well as variability in the self-assembly of soft materials can play an important role in catalysis. Reactivity and catalysis in more complex aggregates such as nanotubes and cubic structures have also recently received attention, showing fascinating opportunities to control transformations in water.[33–35] Considering that the nanostructure morphology has a major effect on reactivity, a key issue in the field of soft materials catalysis is the ability to easily access different morphologies of self-assembled nanosystems. Ideally, a relevant change in aggregation should be performed without the necessity to redesign the amphiphile or perform major structural modifications.

The large variety of structures that can be obtained through the self-assembly of amphiphiles in water has been described by Israelachvili.[36] The difference in aggregate formation is based on the balance between hydrophilic and hydrophobic character of the surfactant, resulting in the formation of micelles, vesicles and inverted structures.[37–39] It has been reported that additional supramolecular interactions (such as π-π stacking or hydrogen bonding) can influence these systems. These non-covalent interactions can lead to more complex morphologies such as nanotubes,[40] sheets,[41] ribbons[42] and helicates.[43] In the challenging task to easily access multiple distinct morphologies of soft materials, simple modifications are a more practical and therefore preferable way to operate. In order to screen variations of the aggregate morphologies and achieve control over reactivity in self-assembled confined space systems, we needed to design a system in which minimal modifications in the last steps of the synthesis could result in a variety of structures, without performing extensive and time-consuming synthetic modifications.

To address this challenge, we designed amphiphilic systems based on 1,1-bi-2-naphthol (BINOL),[44] featuring a unique atropoisomer biaryl core. These novel artificial amphiphiles were synthesized starting from enantiopure BINOL, which is commercially available at low cost in both enantiomeric forms. In addition, it has

(4)

27

2

numerous applications in asymmetric catalysis and there are plentiful synthetic structural modifications to the BINOL scaffold which are well documented in the literature.[45] The self-assembly of BINOL-based amphiphiles has been previously investigated in the solid state,[46] on silica support from the drying of a CH2Cl2 solution,[47] or BINOL has been included as a sodium salt in worm-like micelles.[48] The synthesis of chiral dendritic BINOL derivatives and their application in asymmetric hydrogenation in water has been achieved in a previous report,[49] although a material study was not performed. Moreover, amphiphilic polymers with BINOL moieties have been used for enantioselective recognition of amino acids in water,[50] and a recent example exploited a BINOL-based amphiphile for specific discrimination of arginine by gelation in water.[51] These reports are particularly relevant for the study of BINOL amphiphiles, however the formation and tuning of different aggregates in water has not yet been explored, to our knowledge.

The structural design of BINOL-based amphiphiles presented here is shown in Figure 2.1. Dodecyl chains are chosen as the hydrophobic tails of these amphiphiles since they are known to efficiently induce self-assembly, driving amphiphiles to form a stable bilayer structure.[52] Hydrophilic polyethylene glycol (PEG) chains were attached to allow for solubility in water, while preventing multiple layer stacking.[53– 56]

Figure 2.1: Design of amphiphilic BINOLs 1-4.

As the dodecyl chain is crucial for the self-assembly process, we decided to tune the hydrophilic component of the BINOL amphiphile. We envisioned that through modification of the PEG chains the packing parameter of the amphiphile would be influenced,[36] therefore allowing the morphology of the self-assembled nanostructure to be regulated. An advantage of this design lies in the possibility to modify the terminal moiety of these novel surfactants through late-stage functionalization, allowing for the straightforward synthesis of a small family of compounds, which then translates to the formation of a range of self-assembled morphologies. Therefore, ethylene glycol chains with different lengths and terminal groups were attached to the BINOL core to obtain the four derivatives 1-4, which were investigated in the present study (Figure 2.1).

Compounds 1 and 2 are amphiphilic BINOLs that contain triethylene glycol chains, terminating with OH and OMe groups, respectively. Compounds 3 and 4 represent amphiphiles with tetraethylene glycol units ending with OH and OMe, respectively. By employing this simple approach, we provide an effective platform to

(5)

28

2

easily access different aggregates characterized by the same scaffold, generating a potentially useful homochiral space in a soft material (the BINOL unit).

2.2 Synthesis

The synthesis of the desired amphiphilic BINOL derivatives (Scheme 2.1) started with di-bromination of enantiopure (S)-BINOL 5 at the 6,6’-positions.[57] The phenolic moieties of the obtained di-bromo compound 6 were subsequently protected with methoxymethyl (MOM) groups, which have the added benefit of being convenient

ortho-directing groups for further functionalization. Introduction of two dodecyl

chains to compound 7 could be achieved by several approaches, among which a Kumada cross-coupling reaction[58] allows for easy operation and gram scale synthesis of 8 in good yield (89%). The obtained di-alkylated compound 8 was then treated with n-BuLi and reacted with freshly distilled B(OMe)3, to provide the di-borylated intermediate, which was used without further purification. After addition of H2O2, the mixture was heated at reflux in THF for 1 h to generate the di-hydroxyl compound 9 in 60% isolated yield over two steps. Four different OTs-substituted PEG chains (10a-d), prepared according to reported procedures,[59–62] were reacted with 9 in the presence of base to afford the corresponding BINOL derivatives 11a-d. After deprotection of the MOM groups under acidic conditions and flash column chromatography, the designed BINOL amphiphiles were obtained, and fully characterized via 1H NMR, 13C NMR and HRMS.

(6)

29

2

2.3 Self-Assembly Investigation

With four BINOL amphiphiles in hand, the effect of different hydrophilic substituents on the self-assembly of the amphiphiles in water was studied by Cryo Transmission Electron Microscopy (Cryo-TEM). Compound 2, which contains the shortest PEG chains and is OMe terminated, is the least hydrophilic among the synthesized amphiphiles. As expected, this compound showed the poorest water solubility, and no aggregation was observed. In contrast, the other amphiphiles, with a longer PEG chain and/or OH terminality, present enhanced solubility in aqueous environment which allows for observation of well-defined structures. Amphiphilic BINOL 1, which features a triethylene glycol chain terminated with a OH moiety, was found to assemble into well-defined and tightly packed cubic structures (Figure 2.2a).[63] This self-assembled structure is characterized by highly curved bilayers tightly organized in a bi-continuous phase, presenting a porous system which is evident from the convolutions of the soft material reported in Figure 2.2a. On the other hand, BINOL 3 was found to self-assemble into vesicles. The different aggregated structure, i.e. vesicles, can be attributed to the increased hydrophilicity and larger headgroup surface area (smaller packing parameter) of 3 by extending one extra ethylene glycol unit compared to 1 (Figure 2.2b).[36,38]

Figure 2.2: Cryo-TEM image of self-assembly structures in water of (a) BINOL 1 (cubic structure) and (b) BINOL 3 (vesicles).

As a consequence of the difference in the glycol chain terminal group with respect to compound 3 (OMe instead of OH), we expected to observe a difference in self-assembly of compound 4. Most interestingly, Cryo-TEM studies of 4 revealed the formation of well-defined nanotubes (Figure 2.3), which differ from the usual

(7)

30

2

structures whose outcome is predicted by the packing parameter described by Israelachvili (i.e. micelles, vesicles, planar bilayers and inverted structures).[36]

Figure 2.3: Cryo-TEM image of nanotubes upon self-assembly of BINOL 4 in water.

The nanotubes visible in Figure 3 are several micrometers in length and present uniform diameters of 20 nm. The width of the wall of the tubes was found to be around 4 nm, which, based on the dimensions of 4, suggests the formation of a bilayer.[36] The simplicity of preparing such well-defined structures is worth noting. The addition of water to a dry thin film of 4 immediately resulted in the nanotube formation, without the need of additional operations. Since BINOLs are weak acids, deprotonation at higher pH can potentially lead to a change in packing parameter due to charge repulsion. No significant change of the structures could be observed by hydrating compound 4 in a pH range from 5 to 9, suggesting that the obtained nanotubes are stable within the abovementioned pH range (Figure 2.4).

(8)

31

2

In addition, when a dry thin film of 4 was directly hydrated at pH 10, coexistence of nanotubes and vesicles was observed (Figure 2.5a). The mixture was subsequently neutralized to pH 7, in situ, and only tubular structures were observed (Figure 2.5b). Bringing the solution again to pH 10, the tubular structures remained intact.

Figure 2.5: (a) Self-assembly of 4 at pH = 10; coexistence of vesicles and naotubes; (b) Self-assembly of 4 at pH = 7, after neutralizing from pH = 10.

The above experiment suggests that the nanotube structure is the thermodynamically more stable assembly of the system. The fact that the aggregates do not revert to a vesicular assembly upon making the solution more basic is also a hint that deprotonation is more difficult once the nanotubes are formed.[64]

2.4 Time-Dependent Self-Assembly

Further investigating amphiphiles 1 and 3, it was curiously observed that the aggregation of these compounds changed over time, by simply allowing the samples to stand for few hours after preparation. In all samples of the abovementioned amphiphiles the formation of nanotubular structures was clearly observed (Figure 2.6). Submitting the samples to ultrasound stimulation accelerated this process. In the case of compound 1, the first aggregate observed was indeed a cubic structure (Figure 2.2a), which underwent a rearrangement resulting in the formation of nanotubes. With the help of simple sonication, we could accelerate the process of reorganization. Using Cryo-TEM, we could effectively follow the intermediate step showing coexistence of nanotubes and cubic aggregates (Figure 2.6a), while further sonication of the sample resulted in the formation of nanotubes exclusively (Figure 2.6b). Similarly, compound 3 which first aggregated into vesicles (Figure 2.2b), showed the presence of nanotubes and vesicles simultaneously in an intermediate stage (Figure 2.6c), and could be further pushed to the formation of exclusively nanotubes by the use of sonication (Figure 2.6d).

(9)

32

2

Figure 2.6: (a) Self-assembly of BINOL 1, coexistence of cubic structure and nanotubes; (b) self-assembly of BINOL 1, only nanotubes after sonication; (c) self-assembly of BINOL 3, coexistence of vesicles and nanotubes; (d) self-assembly of BINOL 3, only nanotubes after sonication.

In some specific cases we can observe via Cryo-TEM the growth of nanotubes from the single vesicles themselves (Figure 2.7).

Figure 2.7: Detail of the self-assembled structures of compound 3 after 2 h standing at room temperature. Arrows are pointing to the elongated vesicles.

Upon submitting the formed nanotubes of these samples to environment changes (higher pH, further sonication, higher temperature), no changes in the morphologies of such aggregates were observed, indicating that even in this case the nanotubular structures formed are the thermodynamically most stable assembly of the system. As

(10)

33

2

noted for compound 4, the nanotubular aggregates show incredible stability with respect to external stimuli.

Scheme 2.2: Schematic representation of the self-assembly of amphiphilic BINOLs 1, 3 and 4.

Evidently, these novel BINOL based amphiphiles show various morphologies in the early stage of the self-assembly process (Scheme 2.2), presenting distinct differences in structure as a function of the PEG chain length and terminating substituents. With the simple use of ultrasound stimulation, these metastable aggregates are driven to the formation of nanotubes, probably as a result of the stacking of the BINOL aromatic core of the amphiphile. We anticipate that the torsion angle of the binaphthyl core adapts in the soft material as a result of ultrasound stimulation, causing the surfactant to assume a different morphology and stimulating the conversion from kinetic to thermodynamic self-assemblies (Figure 2.8a). Alternatively, sonication could help bring the BINOL amphiphiles closer together favoring their packing (Figure 2.8b). Due to the stability of the nanotubes under a range of stimuli, and the consistency in the packing of the different amphiphiles, we suggest that forces other than hydrophobic interactions might be involved in the packing (e.g. π-π stacking), although we are not able to provide a theoretical or experimental proof of this, at the moment.

Figure 2.8: (a) Schematic representation of the hypothetical change in torsion angle of the binaphthyl core of BINOL amphiphiles; (b) Schematic representation of the steric reorganization of the packing of BINOL amphiphiles.

(11)

34

2

2.5 Conclusions

In conclusion, by simple modifications we could easily tune the self-assembly behavior of novel artificial BINOL-based amphiphiles. By changing length and terminal moieties of the PEG chain, we were able to control the balance between hydrophilic and hydrophobic character of the surfactants. An additional PEG unit and/or the introduction of OH groups enhances the solubility of the amphiphile in water, hence controlling the self-assembled morphologies and greatly influencing the difference in structures observed, even though the differences in substitution are small. Distinct self-assembled aggregates have been observed, ranging from cubic structures to vesicles. These kinetically formed aggregates are driven to a thermodynamically more stable state by simple ultrasound stimulation, resulting in the formation of nanotubes regardless of the substitution of the chain length of the amphiphiles. This study demonstrates that by rational design important organic ligands such as BINOL derivatives can be assembled into tightly packed, well-organized supramolecular structures in water. Their aggregation morphologies can be addressed during different stages of aggregation i.e. the formation of metastable or thermodynamically stable aggregates. The obtained structures provide a first step towards the effective control of reactivity in self-assembled nanosystems by influencing their morphology. Our work contributes to generate a platform for further applications for instance in asymmetric catalysis in self-assembled bilayers and dynamic responsive soft materials.

2.6 Contributions

The project was carried out under the supervision of dr. J. Chen and prof. dr. B. L. Feringa. The synthesis of compound 1 was carried out H. Smit under the supervision of dr. J. Chen. The synthesis and optimization of all derivatives was carried out by F. Tosi. Cryo-TEM studies were performed by M. C. A. Stuart.

2.7 Experimental Section

Methods. Sample Preparation: 1 mL of a 2 mM solution of the compound (1-4) in technical grade CHCl3 was placed in a 4 mL vial. The solvent (CHCl3) was slowly evaporated in a nitrogen flow and a thin film of the amphiphilic compound was formed, which was subsequently hydrated with double distilled water (1 mL).

pH Dependence Measurement: The pH of a water solution was adjusted with 100

mM aq. HCl or aq. NaOH and measured by a pH meter in a 500 µl solution.

Cryo-TEM. The samples were prepared by depositing a few µL of amphiphile solution on holey carbon coated grids (Quantifoil 3.5/1, Quantifoil Micro Tools, Jena, Germany). After blotting the excess liquid, the grids were vitrified in liquid ethane (Vitrobot, FEI, Eindhoven, The Netherlands) and transferred to a FEI Tecnai T20 cryo-electron microscope equipped with a Gatan model 626 cryo-stage operating at 200

(12)

35

2

keV. Micrographs were recorded under low-dose conditions with a slow-scan CCD camera. The bilayer thickness was measured on slightly defocused Cryo-electron microscopy images to obtain maximal phase contrast.

2.8 Synthetic Procedures

Compound 6[57] and 7[58] were synthesized according to known literature procedures, and analytical data were found in accordance with those reported. Compounds 10a-d were prepared according to a known literature procedure, and characterization data were found according to literature.[59–62]

Grignard reagent synthesis: Magnesium turnings (519 mg,

21.3 mmol) were suspended in THF (25 mL) under a nitrogen atmosphere and the mixture was heated at reflux. The suspension was treated with a crystal of iodine and 1-bromododecane (0.23 mL, 0.97 mmol). When the mixture started boiling, the rest of 1-bromododecane (4.42 mL, 18.4 mmol) was added dropwise at a rate sufficient to maintain gentle reflux. The mixture was left stirring at reflux for 1 h and then allowed to cool down to rt.

Kumada cross-coupling: Compound 7 (1.0 g, 1.9 mmol) and PdCl2(dppf) (73 mg, 0.1 mmol) were dissolved in dry THF (40 mL) under a nitrogen atmosphere and the mixture was cooled to 0 °C. Freshly prepared n-dodecyl-MgBr was added dropwise at 0 °C and the mixture was heated at reflux for 16 h. The mixture was cooled down to rt, quenched with sat. aq. NH4Cl (60 mL), extracted with EtOAc, washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, Pentane/EtOAc: 9:1) gave 8 (1.22 g, 89%) as a pale yellow solid; (Rf: 0.6 Pentane/EtOAc: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.87 (d, J = 9.0 Hz, 2H), 7.63 (s, 2H), 7.53 (d, J = 9.0 Hz, 2H), 7.08 (s, 4H), 5.05 (d, J = 6.7 Hz, 2H), 4.95 (d, J = 6.7 Hz, 2H), 3.14 (s, 6H), 2.70 (t, 4H, J = 7.7 Hz), 1.66 (m, 4H), 1.25 (m, 36H), 0.88 (t, J = 6.4 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 152.4, 138.9, 132.7, 130.4, 129.0, 128.2, 126.5, 125.8, 121.9, 117.8, 95.8, 56.1, 36.2, 32.2, 31.6, 30.0, 30.0, 29.9, 29.9 (2C), 29.7, 29.7, 23.0, 14.4; HRMS (ESI-ion trap) m/z: [M+Na]+ Calcd for C48H70O4Na 733.5166, found 733.5155.

Compound 8 (100 mg, 0.14 mmol) was dissolved in dry THF (8 mL) and cooled down to –78 °C. A solution of t-BuLi (0.43 mL, 0.56 mmol, 1.3 M in Pentane) was added to the reaction mixture, which was left stirring at rt for 1 h and subsequently cooled down to –78 °C. Freshly distilled trimethyl borate (0.14 mL, 0.49 mmol) was added to the mixture which was allowed to warm up to rt and left stirring for 16 h. Aq. H2O2 (30% w/w, 0.08 mL) was added to the mixture which was heated at reflux for 1 h, and the progress of the reaction was followed by

(13)

36

2

TLC analysis (Pentane/EtOAc: 9:1) until complete disappearance of the starting material. The reaction mixture was cooled down to rt, carefully quenched with sat. aq. Na2S2O3 (10 mL) extracted with EtOAc, washed with brine, dried over MgSO4 and concentrated in vacuo. Purified by flash column chromatography (SiO2, Pentane/EtOAc: 9:1 → 8:2) gave 9 (62 mg, 60%)* as a colorless oil; (Rf: 0.1 Pentane/EtOAc: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.53 (s, 2H), 7.43 (s, 2H), 7.35 (s, 2H), 6.97 (s, 2H), 4.71 (d, J = 6.3 Hz, 2H), 4.64 (d, J = 6.3 Hz, 2H), 3.38 (s, 6H), 2.69 (t, J = 7.8 Hz, 4H), 1.66 (m, 4H), 1.31 (m, 36H), 0.88 (t, J = 6.3 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 148.3, 144.5, 140.7, 132.6, 126.9, 126.1, 125.9, 125.7, 125.4, 111.6, 100.0, 57.7, 36.3, 32.2, 31.6, 30.0, 30.0, 29.9 (2C), 29.9, 29.8, 29.7, 23.0, 14.4.

Compound 9 (80 mg, 0.11 mmol), potassium carbonate (152 mg, 1.1 mmol) and compound 10a (70 mg, 0.23 mmol) were dissolved in DMF (30 mL) and the reaction mixture was heated at 85 °C for 16 h. After cooling to room temperature, DMF was removed by rotary evaporation. The solid residue was dissolved in CH2Cl2 (40 mL), washed with MilliQ water (5 mL), washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH 95:5) gave 11a (66.5 mg, 60%) as a colorless oil; (Rf: 0.3 CH2Cl2/MeOH: 8:2); 1H NMR (400 MHz, CDCl3) δ 7.49 (s, 2H), 7.22 (s, 2H), 7.06 (d, J = 8.6 Hz, 2H), 6.96 (d, J = 8.6, 2H), 4.98 (d, J = 5.5 Hz, 2H), 4.87 (d, J = 5.5 Hz, 2H), 4.32 (t, J = 4.9 Hz, 4H), 3.95 (m, 4H), 3.74 (m, 4H), 3.69 (m, 8H), 3.58 (t, J = 4.6 Hz, 4H), 2.67 (t, J = 7.7 Hz, 4H), 2.53 (s, 6H), 1.64 (m, 4H), 1.33 – 1.25 (m, 36H), 0.87 (t, J = 6.2 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 151.0, 143.6, 139.9, 131.3, 127.5, 126.8, 126.2, 125.7, 124.9, 107.9, 98.1, 72.6, 70.9, 70.5, 69.6, 67.8, 61.7, 55.8, 35.9, 31.9, 31.3, 29.7, 29.6, 29.6 (2C), 29.5, 29.4, 29.3, 22.7, 14.1.

Compound 9 (62 mg, 0.083 mmol), potassium carbonate (115 mg, 0.831 mmol) and compound 10b (54 mg, 0.17 mmol) were dissolved in DMF (20 mL) and the mixture was heated at 85 °C for 16 h. After cooling to room temperature, DMF was removed by rotary evaporation. The solid residue was dissolved in CH2Cl2 (40 mL), washed with MilliQ water (5 mL), washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH 99:1 → 95:5) gave 11b (51.7 mg, 60%) as a colorless oil (Rf: 0.1 CH2Cl2/MeOH: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.49 (s, 2H), 7.22 (s, 2H), 7.06 (d, J = 8.5 Hz, 2H), 6.96 (d, J = 8.5 Hz, 2H), 4.97 (d, J = 5.6 Hz, 2H), 4.84 (d, J = 5.6 Hz, 2H), 4.32 (t, J = 5.0 Hz, 4H), 3.95 (t, J = 4.2 Hz, 4H), 3.74 (m, 4H),

* Due to instability, compound 9 was synthesized and used immediately in the following step without being stored.

(14)

37

2

3.71 – 3.61 (m, 8H), 3.54 (m, 4H), 3.37 (s, 6H), 2.68 (t, J = 7.7 Hz, 4H), 2.55 (s, 6H), 1.65 (m, 4H), 1.25 (m, 36H), 0.87 (t, J = 6.7 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 151.4, 144.1, 140.1, 131.6, 127.8, 127.2, 126.5, 125.9, 125.2, 108.2, 98.4, 72.2, 71.2, 71.0, 70.9, 70.0, 68.1, 59.3, 56.2, 36.3, 32.2, 31.6, 30.0 (2C), 30.0, 29.9, 29.9, 29.7, 29.7, 23.0, 14.4.

Compound 9 (109 mg, 0.146 mmol), potassium carbonate (201 mg, 1.41 mmol) and compound 10c (102 mg, 0.292 mmol) were dissolved in DMF (48 mL) and the mixture was heated at 85 °C for 16 h. After cooling to room temperature, DMF was removed by rotary evaporation. The solid residue was dissolved in CH2Cl2 (60 mL), washed with MilliQ water (10 mL), washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH 95:5 → 8:2) gave 11c (88 mg, 55%) as a colorless oil; (Rf: 0.1 CH2Cl2/MeOH: 8:2); 1H NMR (400 MHz, CDCl3) δ 7.50 (s, 2H), 7.22 (s, 2H), 7.07 (d, J = 8.7 Hz, 2H), 6.99 (d, J = 8.7, 2H), 4.96 (d, J = 5.4 Hz, 2H), 4.84 (d, J = 5.4 Hz, 2H), 4.32 (m, 4H), 3.96 (m, 4H), 3.76 – 3.66 (m, 20H), 3.56 (m, 2H), 3.39 (s, 6H), 2.70 (m, 4H), 2.57 (s, 6H), 1.67 (m, 4H), 1.33 – 1.28 (m, 36H), 0.87 (t, J = 6.2 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 151.4, 144.0, 140.1, 131.6, 127.8, 127.1, 126.5, 125.9, 125.2, 108.2, 98.4, 72.8, 71.1, 71.0, 70.9, 70.6, 69.9, 68.1, 62.0, 56.2, 36.3, 32.2, 31.6, 30.0, 29.9, 29.9, 29.8 (2C), 29.7, 29.6, 23.0, 14.4.

Compound 9 (80 mg, 0.11 mmol), potassium carbonate (152 mg, 1.1 mmol) and compound 10d (84 mg, 0.23 mmol) were dissolved in dry DMF (30 mL) and the mixture was heated at 85 °C for 16 h. After cooling to room temperature, DMF was removed by rotary evaporation. The solid residue was dissolved in CH2Cl2 (40 mL), washed with MilliQ water (5 mL), washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH: 99:1 → 95:5) gave 11d (80.3 mg, 65%) as a colorless oil; (Rf: 0.1 CH2Cl2/MeOH: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.49 (s, 2H), 7.22 (s, 2H), 7.05 (d, J = 8.7 Hz, 2H), 6.97 (d, J = 8.7 Hz, 2H), 4.97 (d, J = 5.4 Hz, 2H), 4.45 (d, J = 5.4 Hz, 2H), 4.31 (m, 4H), 3.95 (m, 4H), 3.73 (m, 4H), 3.67 – 3.63 (m, 16H), 3.54 (m, 4H), 3.37 (s, 6H), 2.68 (m, 4H), 2.55 (s, 6H), 1.64 (m, 4H), 1.31 – 1.25 (m, 36H), 0.88 (t, J = 6.2 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 151.4, 144.0, 140.1, 131.6, 127.7, 127.1, 126.5, 125.9, 125.2, 108.1, 98.4, 72.2, 71.1, 70.9, 70.9, 70.9, 70.8, 69.9, 68.0, 59.3, 56.2, 36.2, 32.2, 31.6, 30.0 (2C), 29.9, 29.9, 29.8, 29.7, 29.6, 23.0, 14.4; HRMS (ESI-ion trap) m/z: [M+H]+ Calcd for C66H107O14 1123.7661, found 1123.7637.

Compound 11a (65 mg, 0.065 mmol) was dissolved in 1,4-dioxane (10 mL). The reaction mixture was treated with aq. HCl (37%, 0.32 mL) and left stirring at rt for

(15)

38

2

16 h. The mixture was poured into water (2 mL) and extracted with CH2Cl2 (3x10 mL). The combined organic layers were washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH: 9:1) gave 1 (47.8 mg, 80%) as a colorless oil; (Rf: 0.2 CH2Cl2/MeOH: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.49 (s, 2H), 7.21 (s, 2H), 7.04 (d, J = 8.5 Hz, 2H), 6.96 (d, J = 8.5 Hz, 2H), 4.35 (m, 4H), 3.92 (m, 4H), 3.71 (m, 8H), 3.58 (m, 8H), 2.67 (t, J = 7.7 Hz, 4H), 1.64 (m, 4H), 1.33 – 1.25 (m, 36H), 0.88 (d, J = 6.6 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 147.3, 144.1, 138.4, 129.2, 128.2, 126.2, 125.6, 125.2, 116.4, 108.0, 72.8, 71.1, 70.5, 69.7, 68.5, 62.0, 36.2, 32.2, 31.8, 30.0 (2C), 29.9 (2C), 29.9, 29.8, 29.6, 23.0, 14.41; HRMS (ESI-ion trap) m/z: [M+H]+ Calcd for C56H87O10 919.6293, found 919.6297.

Compound 11b (25 mg, 0.024 mmol) was dissolved in dry 1,4-dioxane (1 mL). The reaction mixture was treated with aq. HCl (37%, 0.08 mL) and left stirring at rt for 16 h. The mixture was poured into water (2 mL) and extracted with CH2Cl2 (3x10 mL). The combined organic layers were washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH: 97:3) gave 2 (14 mg, 68%) as a colorless oil; (Rf: 0.5 CH2Cl2/MeOH: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.51 (s, 2H), 7.24 (s, 2H), 7.05 (d, J = 8.6 Hz, 2H), 6.98 (d, J = 8.6 Hz, 2H), 4.38 (m, 4H), 3.95 (t, J = 4.7 Hz, 4H), 3.75 (m, 4H), 3.67 (m, 4H), 3.61 (m, 4H), 3.49 (m, 4H), 3.31 (s, 6H), 2.67 (t, J = 7.7 Hz, 4H), 1.64 (m, 4H), 1.31 – 1.25 (m, 36H), 0.87 (d, J = 6.5 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 146.9, 143.8, 138.6, 129.3, 128.0, 126.4, 125.8, 125.1, 115.5, 108.3, 72.2, 71.1, 70.9, 70.9, 69.8, 69.1, 59.3, 36.2, 32.2, 31.7, 30.0, 29.9, 29.9 (2C), 29.9, 29.7, 29.7, 23.0, 14.4; HRMS (ESI-ion trap) m/z: [M+H]+ Calcd for C58H91O10 947.6607, found 947.6637, [M+Na]+ Calcd for C58H90O10Na1 969.6426, found 969.6455.

Compound 11c (40 mg, 0.036 mmol) was dissolved in dry 1,4-dioxane (5 mL). The reaction mixture was treated with aq. HCl (37%, 0.16 mL) and left stirring at rt for 16 h. The mixture was poured into water (2 mL) and extracted with CH2Cl2 (3x10 mL). The combined organic layers were washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH: 9:1) gave 3 (11.6 mg, 32%) as a colorless oil; (Rf: 0.3 CH2Cl2/MeOH: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.49 (s, 2H), 7.20 (s, 2H), 7.04 (d, J = 8.6 Hz, 2H), 6.95 (d, J = 8.6 Hz, 2H), 4.37 (m, 4H), 3.96 (m, 4H), 3.74 (m, 4H), 3.68 (m, 4H), 3.60 (m, 12H), 3.48 (m, 4H), 2.66 (t, J = 7.8 Hz, 4H), 1.63 (m, 4H), 1.33 – 1.21 (m, 36H), 0.88 (d, J = 7.0 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 147.4, 144.1, 138.3, 129.2, 128.1, 126.1, 125.6, 125.2, 116.1, 107.6, 72.7, 70.9, 70.8, 70.7, 70.4, 69.8, 68.4, 61.7, 36.2, 32.2, 31.8, 30.0 (2C), 29.9 (2C), 29.9, 29.8, 29.7, 23.0, 14.4; HRMS

(16)

(ESI-39

2

ion trap) m/z: [M+H]+ Calcd for C66H95O12 1007.6818, found 1007.6823, [M+Na]+ Calcd for C66H94O12Na1 1029.6650, found 1029.6638.

Compound 11d (35 mg, 0.031 mmol) was dissolved in dry 1,4-dioxane (5 mL). The reaction mixture was treated with aq. HCl (37%, 0.16 mL) and left stirring at rt for 16 h. The mixture was poured into water (2 mL) and extracted with CH2Cl2 (3x10 mL). The combined organic layers were washed with brine, dried over MgSO4 and concentrated in vacuo. Purification by flash column chromatography (SiO2, CH2Cl2/MeOH: 95:5 → 9:1) gave 4 (24 mg, 74%) as a colorless oil; (Rf: 0.5 CH2Cl2/MeOH: 9:1); 1H NMR (400 MHz, CDCl3) δ 7.51 (d, J = 1.8 Hz, 2H), 7.24 (s, 2H), 7.05 (d, J = 8.5 Hz, 2H), 6.98 (dd, J = 8.5 Hz, 1.8 Hz, 2H), 6.30 (br, 2H), 4.38 (m, 4H), 3.96 (t, J = 4.6 Hz, 4H), 3.75 – 3.54 (m, 24H), 3.31 (s, 6H), 2.67 (t, J = 7.8 Hz, 4H), 1.62 (m, 4H), 1.33 – 1.25 (m, 36H), 0.87 (d, J = 6.7 Hz, 6H); 13C NMR (101 MHz, CDCl3) δ 147.2, 144.0, 138.3, 129.1, 128.0, 126.1, 125.6, 125.1, 115.9, 107.5, 72.0, 70.9, 70.7 (2C), 70.7 (2C), 69.7, 68.4, 59.1, 36.2, 32.2, 31.8, 30.0, 29.9, 29.8, 29.8, 29.7, 29.7, 29.6, 23.0, 14.4; HRMS (ESI-ion trap) m/z: [M+H]+ Calcd for C62H99O12 1035.7131, found 1035.7104.

2.9 References

[1] S. Polarz, M. Kunkel, A. Donner, M. Schlötter, Chem. Eur. J. 2018, 24, 18842–18856. [2] T. Dam, J. B. F. N. Engberts, J. Karthäuser, S. Karaborni, N. M. Van Os, Colloids Surfaces A

Physicochem. Eng. Asp. 1996, 118, 41–49.

[3] Y. Rui, S. Wang, P. S. Low, D. H. Thompson, J. Am. Chem. Soc. 1998, 120, 11213–11218. [4] J. P. Behr, Acc. Chem. Res. 1993, 26, 274–278.

[5] A. D. Miller, Angew. Chem. Int. Ed. 2002, 37, 1768–1785.

[6] R. C. MacDonald, G. W. Ashley, M. M. Shida, V. A. Rakhmanova, Y. S. Tarahovsky, D. P. Pantazatos, M. T. Kennedy, E. V. Pozharski, K. A. Baker, R. D. Jones, et al., Biophys. J.

1999, 77, 2612–2629.

[7] J. B. F. N. Engberts, Pure Appl. Chem. 2007, 64, 1653–1660. [8] F. M. Menger, C. E. Portnoy, J. Am. Chem. Soc. 1967, 89, 4698–4703.

[9] T. Dwars, E. Paetzold, G. Oehme, Angew. Chem. Int. Ed. 2005, 44, 7174–7199.

[10] D. M. Vriezema, M. C. Aragonès, J. A. A. W. Elemans, J. J. L. M. Cornelissen, A. E. Rowan, R. J. M. Nolte, Chem. Rev. 2005, 105, 1445–1489.

[11] J. H. Fendler, E. J. Fendler, Catalysis in Micelles and Macromolecular Systems, Academic Press, London, 1975.

[12] F. M. Menger, J. U. Rhee, H. K. Rhee, J. Org. Chem. 1975, 40, 3803–3805. [13] B. H. Lipshutz, Curr. Opin. Green Sustain. Chem. 2018, 11, 1–8.

[14] S. Serrano-Luginbühl, K. Ruiz-Mirazo, R. Ostaszewski, F. Gallou, P. Walde, Nat. Rev.

Chem. 2018, 2, 306–327.

[15] F. Fernandez-Trillo, L. M. Grover, A. Stephenson-Brown, P. Harrison, P. M. Mendes,

Angew. Chem. Int. Ed. 2017, 56, 3142–3160.

[16] K. Fuji, T. Morimoto, K. Tsutsumi, K. Kakiuchi, Angew. Chem. Int. Ed. 2003, 42, 2409– 2411.

[17] P. A. Grieco, K. Yoshida, P. Garner, J. Org. Chem. 1983, 48, 3137–3139.

[18] B. H. Lipshutz, S. Ghorai, G. T. Aguinaldo, Adv. Synth. Catal. 2008, 350, 953–956. [19] T. Rispens, J. B. F. N. Engberts, J. Org. Chem. 2002, 67, 7369–7377.

(17)

40

2

[20] S. Kobayashi, T. Wakabayashi, S. Nagayama, H. Oyamada, Tetrahedron Lett. 1997, 38, 4559–4562.

[21] K. Manabe, Y. Mori, T. Wakabayashi, S. Nagayama, S. Kobayashi, J. Am. Chem. Soc. 2000,

122, 7202–7207.

[22] S. Otto, J. B. F. N. Engberts, J. C. T. Kwak, J. Am. Chem. Soc. 1998, 120, 9517–9525. [23] R. Selke, J. Holz, A. Riepe, A. Börner, Chem. Eur. J. 1998, 4, 769–771.

[24] S. Handa, J. D. Smith, M. S. Hageman, M. Gonzalez, B. H. Lipshutz, ACS Catal. 2016, 6, 8179–8183.

[25] G. Oehme, E. Paetzold, R. Selke, J. Mol. Catal. 1992, 71, L1–L5.

[26] J. van Esch, M. F. Roks, R. J. M. Nolte, J. Am. Chem. Soc. 1986, 108, 6093–6094. [27] J. T. Groves, R. Neumann, J. Am. Chem. Soc. 1987, 109, 5045–5047.

[28] D. Paprocki, D. Koszelewski, P. Walde, R. Ostaszewski, RSC Adv. 2015, 5, 102828– 102835.

[29] A. Madej, D. Paprocki, D. Koszelewski, A. Żądło-Dobrowolska, A. Brzozowska, P. Walde, R. Ostaszewski, RSC Adv. 2017, 7, 33344–33354.

[30] M. Hirose, T. Ishigami, K. Suga, H. Umakoshi, Langmuir 2015, 31, 12968–12974. [31] J. H. Fendler, W. L. Hinze, J. Am. Chem. Soc. 1981, 103, 5439–5447.

[32] J. Pérez-Juste, F. Hollfelder, A. J. Kirby, J. B. F. N. Engberts, Org. Lett. 2000, 2, 127–130. [33] K. S. Lee, J. R. Parquette, Chem. Commun. 2015, 51, 15653–15656.

[34] M. Duss, L. Salvati Manni, L. Moser, S. Handschin, R. Mezzenga, H. J. Jessen, E. M. Landau,

ACS Appl. Mater. Interfaces 2018, 10, 5114–5124.

[35] T. Zhou, J. J. Vallooran, S. Assenza, A. Szekrenyi, P. Clapés, R. Mezzenga, ACS Catal. 2018,

8, 5810–5815.

[36] J. N. Israelachvili, Intermolecular and Surface Forces, Academic Press, London, 2011. [37] Y. Chevalier, T. Zemb, Reports Prog. Phys. 1990, 53, 279–371.

[38] D. E. Discher, A. Eisenberg, Science 2002, 297, 967–973.

[39] J. M. Seddon, R. H. Templer, P. Trans, R. S. Lond, Philos. Trans. R. Soc. A 1993, 344, 377– 401.

[40] T. G. Barclay, K. Constantopoulos, J. Matisons, Chem. Rev. 2014, 114, 10217–10291. [41] L. C. Palmer, S. I. Stupp, Acc. Chem. Res. 2008, 41, 1674–1684.

[42] Y. V Zastavker, N. Asherie, A. Lomakin, J. Pande, J. M. Donovan, J. M. Schnur, G. B. Benedek, Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 7883–7887.

[43] Y. Zhang, R. Tan, G. Zhao, X. Luo, D. Yin, Catal. Sci. Technol. 2016, 6, 488–496. [44] J. M. Brunel, Chem. Rev. 2005, 105, 4233–4233.

[45] D. Parmar, E. Sugiono, S. Raja, M. Rueping, Chem. Rev. 2014, 114, 9047–9153.

[46] J. Lu, S.-Z. Kang, S.-L. Xu, Q.-D. Zeng, C. Wang, L.-J. Wan, C.-L. Bai, Chem. Commun. 2003, 1498.

[47] X. Xiao, Y. Wu, M. Sun, J. Zhou, Z. Bo, L. Li, C. Chan, J. Polym. Sci. Part A Polym. Chem.

2008, 46, 574–584.

[48] T. H. Ito, A. G. Salles, J. P. Priebe, P. C. M. L. Miranda, N. H. Morgon, D. Danino, G. Mancini, E. Sabadini, Langmuir 2016, 32, 8461–8466.

[49] Q.-H. Fan, R. Zhang, G.-J. Deng, X.-M. Chen, J. Chem. Res. 2009, 2002, 318–320. [50] S. Nian, L. Pu, Chem. Eur. J. 2017, 23, 18066–18073.

[51] L. Jia, J. Yin, X. Guo, G. Cao, X. Tian, B. Zhu, L. Pu, Soft Matter 2017, 13, 5453–5462. [52] R. Nagarajan, Langmuir 2002, 18, 31–38.

[53] A. C. Coleman, J. M. Beierle, M. C. A. Stuart, B. Maciá, G. Caroli, J. T. Mika, D. J. van Dijken, J. Chen, W. R. Browne, B. L. Feringa, Nat. Nanotechnol. 2011, 6, 547–552.

[54] D. J. van Dijken, J. Chen, M. C. A. Stuart, L. Hou, B. L. Feringa, J. Am. Chem. Soc. 2016, 138, 660–669.

[55] F. K. C. Leung, T. Van Den Enk, T. Kajitani, J. Chen, M. C. A. Stuart, J. Kuipers, T. Fukushima, B. L. Feringa, J. Am. Chem. Soc. 2018, 140, 17724–17733.

[56] J. Chen, F. K. C. Leung, M. C. A. Stuart, T. Kajitani, T. Fukushima, E. Van Der Giessen, B. L. Feringa, Nat. Chem. 2018, 10, 132–138.

[57] J. E. Beaver, B. C. Peacor, J. V. Bain, L. I. James, M. L. Waters, Org. Biomol. Chem. 2015, 13, 3220–3226.

(18)

41

2

[58] N. Lv, M. Xie, W. Gu, H. Ruan, S. Qiu, C. Zhou, Z. Cui, Org. Lett. 2013, 15, 2382–2385.

[59] K. Niikura, N. Iyo, T. Higuchi, T. Nishio, H. Jinnai, N. Fujitani, K. Ijiro, J. Am. Chem. Soc.

2012, 134, 7632–7635.

[60] H. Pan, M. E. Grow, O. Wilson, M. C. Daniel, Tetrahedron 2013, 69, 2799–2806.

[61] E. Mayot, P. Lemière, C. Gérardin-Charbonnier, European J. Org. Chem. 2008, 2008, 2232–2239.

[62] C. Wendeln, S. Rinnen, C. Schulz, T. Kaufmann, H. F. Arlinghaus, B. J. Ravoo, Chem. Eur. J.

2012, 18, 5880–5888.

[63] D. P. Siegel, Biophys. J. 1986, 49, 1171–1183.

(19)

Referenties

GERELATEERDE DOCUMENTEN

Here we present a salen-based amphiphile which, using complexation with distinct transition metal ions, allows to control effectively the self-assembly morphology in

The thermo-responsive phase transition observed with micro-DSC was unequivocably confirmed by Cryo-TEM measurements of the heated samples, which revealed the formation of vesicles

been reported to be catalyzed by Co and Fe salen complexes, which provided us with a further possibility to find application for metallo-amphiphiles L1-Co and L1-Fe. The

Gratifyingly, identical results were achieved with both electron-poor and electron-rich substrates (5a-8a). Unwanted side reactions were suppressed with near perfect

Employing our recently discovered highly active palladium nanoparticle based catalyst (Chapter 6), we were able to couple an alkenyllithium reagent with high (Z/E)

Het hoofdstuk beschrijft de verschillende aggregaten die met amfifielen in het water kunnen worden verkregen en hun gebruik voor chemische reacties in het water.. Een selectie

I had to write you few times before I got an interview with you for a PhD position and my level of excitement (and anxiety) in that moment was beyond the roof. I was

With the aim of directing self-assembly of amphiphiles in water to different stages, transition metals were used as an effective tool to access different geometries, ranging from