Next Article in Journal
Biohydrogen Production: Strategies to Improve Process Efficiency through Microbial Routes
Next Article in Special Issue
Reconciling Experiment and Theory in the Use of Aryl-Extended Calix[4]pyrrole Receptors for the Experimental Quantification of Chloride–π Interactions in Solution
Previous Article in Journal
MicroRNAs Regulate Bone Development and Regeneration
Previous Article in Special Issue
Surface Modification of ZnO Nanorods with Hamilton Receptors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Revisiting the Formation and Tunable Dissociation of a [2]Pseudorotaxane Formed by Slippage Approach

1
Department of Chemistry and Institute of Creativity, The Hong Kong Baptist University, Kowloon Tong, Kowloon, Hong Kong, China
2
Institute of Molecular Functional Materials, University Grants Committee, Hong Kong, China
3
Department of Chemistry, The Chinese University of Hong Kong, Shatin, NT, Hong Kong, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2015, 16(4), 8254-8265; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms16048254
Submission received: 26 March 2015 / Revised: 9 April 2015 / Accepted: 10 April 2015 / Published: 13 April 2015
(This article belongs to the Special Issue Supramolecular Interactions)

Abstract

:
A new [2]pseudorotaxane DB24C81-H·PF6 with dibenzo[24]crown-8 (DB24C8) crown ether-dibenzylammonium (1-H·PF6) binding which was formed by slippage approach at different solvents and temperature, had been isolated and characterized by NMR spectroscopy and mass spectrometry. The [2]pseudorotaxane DB24C81-H·PF6 was stable at room temperature. The dissociation rate of [2]pseudorotaxane DB24C81-H·PF6 could be tuned by using different stimuli such as triethylamine (TEA)/diisopropylethylamine (DIPEA) and dimethyl sulfoxide (DMSO). In particular, the dissociation of [2]pseudorotaxane DB24C81-H·PF6 by an excess of TEA/DIPEA base mixture possessed a long and sustained, complete dissociation over 60 days. Other stimuli by DMSO possessed a relatively fast dissociation over 24 h.

Graphical Abstract

1. Introduction

Rotaxanes contain a linear dumbbell-shaped component bearing bulky end-groups or stoppers around which one or more macrocycles are trapped. On the other hand, pseudorotaxanes are temporally encircled around an unstoppered thread through noncovalent interactions from which they are readily susceptible to dissociation without breaking a covalent bond [1,2,3,4]. Rotaxanes and pseudorotaxanes have been studied extensively for the ability of the interlocked ring to be switched on demand by external stimuli such as pH [5,6], electrochemical reagents [7], heat [8,9], moisture [10,11], salt [12], light [13], etc. Coupled with their ability to be customized and optimized for nanoscale functions, these interlocked molecules are excellent candidates as movable elements in the construction of nanovalves [14] based on a porous, solid-phase support, for controlled substrate release. Many examples have been demonstrated about their relatively fast substrate releases within 24 h by tuning the pH values in the solution [12]. Currently, no pseudorotaxane or rotaxane building blocks are available and suitable for the construction of vehicles for sustained substrate release over a week.
The construction of rotaxane-like assemblies has recently relied upon thermodynamic, templated reactions with enhanced efficiencies [15,16,17,18,19,20,21,22,23,24]. The term “slippage” has been coined [25,26,27,28] for pseudorotaxane synthesis that employs thermodynamic threading of macrocycle to a “dumbbell”. This strategy utilizes (1) the size complementarily between the macrocycle and the “dumbbell’s” stoppers; and (2) the stabilizing noncovalent bonding interactions between the macrocycle and the “dumbbell’s” rod. In this strategy, the macrocycle and the dumbbell have been separately synthesized, prior to heating them together in solution so that the free energy of activation for the thermodynamic threading (slippage) of macrocycle to dumbbell can be overcome. The presence of a template on the dumbbell’s rod renders the pseudorotaxane structure more stable so that the free energy of activation for its dissociation becomes insurmountable when the solution has been cooled to ambient temperature. There are several disadvantages of the slippage approach in terms of reaction time and stability. Generally, it requires a long reaction time with a large slippage rate (kon) for pseudorotaxane formation from usually a few days up to 90 days [27]. However, the dissociation of pseudorotaxane will sometimes have an uncontrollable, small slippage rate (koff). Therefore, there is a need to study a better reaction condition for pseudorotaxane formation by slippage and to tune the pseudorotaxane dissociation rate (koff << kon) from days to hours.

2. Results and Discussion

Herein, we employ a slippage approach in one pot (Figure 1) to yield a thermodynamically stable [2]pseudorotaxane [29,30] by mixing an ammonium thread (dumbbell) 1-H·PF6 with dibenzo[24]crown-8 (DB24C8). Benzo-crown ether DB24C8 is capable to recognize with secondary ammonium ions by virtue of [N+–H···O] and [N+–C–H···O] hydrogen bonds, electrostatic interactions, and augmented with some aromatic π-π interactions [31,32,33,34,35,36]. The 3,5-dimethoxyaryl moiety at one end of the dumbbell 1-H·PF6, is a relatively bulky stopper that can effectively block the passageway of DB24C8 at high reaction temperature. On the other hand, the cyclohexyl ring of the dumbbell-like thread, in contrast, allow the crown ether to thread through to the dumbbell’s rod at elevated reaction temperature with the molecular flipping of between their chair-boat-chair forms [31,32,33,34,35,36,37].
Figure 1. Molecular structures of the crown ether DB24C8 and the ammonium thread “dumbbell” 1-H·PF6. The 3,5-dimethoxyaryl group of the thread acts as an effective stopper to prevent the extrusion of macrocycle while the cyclohexyl group of the thread allows the slipping of macrocycle upon heating in selected solvents. The thermodynamically stable [2]pseudorotaxane DB24C81-H·PF6 was isolated and subjected to dissociation studies with excess amine base (Et3N and DIPEA) mixture and dimethyl sulfoxide (DMSO).
Figure 1. Molecular structures of the crown ether DB24C8 and the ammonium thread “dumbbell” 1-H·PF6. The 3,5-dimethoxyaryl group of the thread acts as an effective stopper to prevent the extrusion of macrocycle while the cyclohexyl group of the thread allows the slipping of macrocycle upon heating in selected solvents. The thermodynamically stable [2]pseudorotaxane DB24C81-H·PF6 was isolated and subjected to dissociation studies with excess amine base (Et3N and DIPEA) mixture and dimethyl sulfoxide (DMSO).
Ijms 16 08254 g001
The reaction time of the pseudorotaxane synthesis was fixed in 4 days with reasonable yields (~50%). The percentage yields (Table 1) of individual [2]pseudorotaxane after a reaction time of 4 days, have been determined by 1H NMR spectroscopy according to their “bound” and “free” signal intensities [36,37,38]. From the results, the reaction yields and rates in synthesizing the [2]pseudorotaxanes are sensitive to different solvents and temperatures. In particular, the percentage yield of the [2]pseudorotaxane DB24C81-H·PF6 was found to be the highest (46%) after a reaction temperature at 70 °C with MeCN compared to using PhMe (70 °C) and CH2Cl2 (40 °C). This is partially because, in polar solvents, the [2]pseudorotaxane might undergo intra and/or intermolecular aggregation of the hydrophobic alkyl chains and this will overcome the extrusion effect. The [2]pseudorotaxane DB24C81-H·PF6 could be isolated by flash chromatography on silica gel (CH2Cl2/THF = 7/1) and characterized by NMR spectroscopy and mass spectrometry.
Table 1. Percentage yields of the [2]pseudorotaxanes after slippage reaction for 4 days. Analyzed by 1H NMR spectroscopy (400 MHz, 298 K).
Table 1. Percentage yields of the [2]pseudorotaxanes after slippage reaction for 4 days. Analyzed by 1H NMR spectroscopy (400 MHz, 298 K).
Solvent (Reaction Temperature in Sealed Tube)DB24C8⊃1-H·PF6
MeCN (70 °C)46%
PhMe (70 °C)30%
CH2Cl2 (40 °C)10%
1H NMR spectroscopy was employed to evaluate the structural features of [2]pseudorotaxanes. By way of an example, the isolated and pure [2]pseudorotaxane DB24C81-H·PF6 reveals (Figure 2) proton chemical shifts of the –CH2NH2+CH2– moieties at δ = 3.14 and 3.32 ppm (bound), comparing to the –CH2NH2+CH2– moieties of its thread 1-H·PF6 at δ = 2.81 and 3.06 ppm (free).
Figure 2. Partial 1H NMR spectra (400 MHz, CD3CN, 298 K) of crown ether DB24C8, isolated pure [2]pseudorotaxane DB24C81-H·PF6, and ammonium thread 1-H·PF6.
Figure 2. Partial 1H NMR spectra (400 MHz, CD3CN, 298 K) of crown ether DB24C8, isolated pure [2]pseudorotaxane DB24C81-H·PF6, and ammonium thread 1-H·PF6.
Ijms 16 08254 g002
High-resolution electrospray ionization mass spectrometry (ESI-MS) has been employed (Figure 3) to further characterize the [2]pseudorotaxane. The molecular ion peak at m/z 740 which is the most abundant peak in the spectrum, is corresponded to the [M–PF6]+ ion of the [2]pseudorotaxane.
Figure 3. Electrospray ionization mass spectrometry (ESI-MS) of the [2]pseudorotaxane DB24C81-H·PF6, showing the molecular ion signal [M–PF6]+ at m/z 740.
Figure 3. Electrospray ionization mass spectrometry (ESI-MS) of the [2]pseudorotaxane DB24C81-H·PF6, showing the molecular ion signal [M–PF6]+ at m/z 740.
Ijms 16 08254 g003
Furthermore, the dissociation and stabilities of [2]pseudorotaxane towards organic amine bases and a hydrogen bond disrupting solvent were evaluated. In particular, pure [2]pseudorotaxane DB24C81-H·PF6 was dissolved in CD3CN and was treated with an excess of triethylamine (TEA)/diisopropylethylamine (DIPEA) mixture [5,6,12]. The ammonium ion of the thread 1-H·PF6 could be successfully deprotonated by the bases wherein the DB24C8 loss its binding affinity towards the deprotonated, amine thread 1. Since the template effect is lost, the [2]pseudorotaxane is no longer stable whereas extrusion of macrocycle occurs with the molecular flipping of the cyclohexyl ring. This extrusion behavior was monitored (Figure 4) by observing a significant decrease of characteristic signal at δ = 6.16 ppm (bound ArH of 1-H·PF6) as well as an increase of signal at δ = 6.38 ppm (free ArH of 1) from their 1H NMR spectra over time. Interestingly, the [2]pseudorotaxane requires almost 60 days to dissociate completely into two separate components at ambient temperature. The half-life of dissociation was determined to be approximately 2 days (Figure 6).
Figure 4. Partial 1H NMR spectra (400 MHz, CD3CN, 298 K) showing the extrusion of DB24C8 macrocycle from the isolated [2]pseudorotaxane DB24C81-H·PF6 with excess triethylamine (TEA)/diisopropylethylamine (DIPEA) mixture in CD3CN over time.
Figure 4. Partial 1H NMR spectra (400 MHz, CD3CN, 298 K) showing the extrusion of DB24C8 macrocycle from the isolated [2]pseudorotaxane DB24C81-H·PF6 with excess triethylamine (TEA)/diisopropylethylamine (DIPEA) mixture in CD3CN over time.
Ijms 16 08254 g004
On the other hand, by dissolving the pure [2]pseudorotaxane DB24C81-H·PF6 in a hydrogen bond disfavored solvent—dimethyl sulfoxide (DMSO) [1,27], extrusion of DB24C8 from the thread 1-H·PF6 occurred. Since the template effect is lost, the [2]pseudorotaxane is no longer stable whereas extrusion of macrocycle occurs with the molecular flipping of the cyclohexyl ring. This extrusion behavior was monitored (Figure 5) by observing the decreases of characteristic signals at δ = 3.81, and 4.14 ppm (bound –OCH2CH2O– of DB24C8) as well as the increases of signals at δ = 3.74, and 4.07 ppm (free –OCH2CH2O– of DB24C8) from their 1H NMR spectra (in DMSO-d6) over time. The [2]pseudorotaxane required only 24 h for a complete dissociation. The half-life of dissociation was determined to be approximately 4 h (Figure 6).
Figure 5. Stacked 1H NMR spectra (400 MHz, CD3SOCD3, 298 K) showing the extrusion of DB24C8 macrocycle from the isolated [2]pseudorotaxane DB24C81-H·PF6 in DMSO-d6 over time.
Figure 5. Stacked 1H NMR spectra (400 MHz, CD3SOCD3, 298 K) showing the extrusion of DB24C8 macrocycle from the isolated [2]pseudorotaxane DB24C81-H·PF6 in DMSO-d6 over time.
Ijms 16 08254 g005
Figure 6. Dissociation data of the [2]pseudorotaxane DB24C81-H·PF6 in the presence of base and DMSO, characterized by 1H NMR spectroscopy. A plot of the concentration (%) of the [2]pseudorotaxane DB24C81-H·PF6 versus time (day).
Figure 6. Dissociation data of the [2]pseudorotaxane DB24C81-H·PF6 in the presence of base and DMSO, characterized by 1H NMR spectroscopy. A plot of the concentration (%) of the [2]pseudorotaxane DB24C81-H·PF6 versus time (day).
Ijms 16 08254 g006

3. Experimental Section

General Information. 1H NMR (400 MHz) and 13C NMR (100 MHz) spectra were recorded at room temperature in CDCl3 unless otherwise stated. Each solvent residual signal was used as the internal standard. Chemical shifts were reported as parts per million (ppm) in δ scale and coupling constants (J) were reported in hertz. Mass spectra were obtained on a double focusing sector mass spectrometer with electrospray ionization (ESI) technique. The reported molecular mass (m/z) values, unless otherwise specified, were mono-isotopic mass. All reactions were carried out under N2. All reactions were monitored by thin layer chromatography (TLC) performed on pre-coated silica gel 60 F254 plates, and compounds were visualized with a spray of 5% (w/v) dodecamolybdophosphoric acid in ethanol and subsequent heating. Flash chromatography was carried out on columns of silica gel (230–400 mesh). Tetrahydrofuran (THF) was freshly distilled prior to use from sodium/benzophenone ketyl under N2. CH2Cl2 was freshly distilled from CaH2. Cyclohexanemethylamine (3) and dibenzo[24]crown-8 (DB24C8) were commercially available from Sigma-Aldrich (St. Louis, MO, USA) while the starting compound (2) was synthesized according to the literature procedures [39]. The synthetic scheme of new compounds is shown in Scheme 1.
Amide 4. 1-[3-(Dimethylamino)propyl]-3-ethylcarbodiimide methiodide (EDCI, 2.53 g, 8.52 mmol) was added to a stirred solution which contained 3-(3,5-dimethoxyphenyl)-propionic acid (2, 1.43 g, 6.81 mmol), cyclohexanemethylamine (3, 1.00 mL, 7.69 mmol) and 1-hydroxybenzotriazole (HOBt) (1.14 g, 8.45 mmol) in CH2Cl2 (25 mL). The reaction mixture was stirred for 18 h at 25 °C and concentrated under reduced pressure. The residue was purified by flash chromatography on silica gel (hexane/EtOAc = 4/1) to afford the amide (4) (1.50 g, 72%) as a white solid. M.p.: 72.6–74.3 °C. Rf: 0.58 (hexane/EtOAc = 4/1). 1H NMR: δ 0.64–0.81 (2 H, m, CH2), 0.92–1.14 (3 H, m), 1.20–1.34 (1 H, m), 1.43–1.62 (5 H, m), 2.38 (2 H, t, J = 7.6, CH2CH2), 2.77 (2 H, t, J = 7.6, CH2C=O), 2.92 (2 H, t, J = 6.4, CHCH2), 3.60 (6 H, s, CH3), 6.16 (1 H, s, ArH), 6.23 (2 H, s, ArH), 6.61 (1 H, t, J = 5.2, NH). 13C NMR: δ 25.5, 26.1, 30.5, 31.8, 37.57, 37.63, 45.4, 54.7, 97.6, 106.0, 143.1, 160.5, 172.1. ESI-MS: m/z 328 ([M + Na]+, 100%). HRESI-MS: calcd m/z for C18H27NO3Na: 328.1883, found: 328.1883 ([M + Na]+, 100%).
Scheme 1. Synthesis of the ammonium thread 1-H·PF6.
Scheme 1. Synthesis of the ammonium thread 1-H·PF6.
Ijms 16 08254 g007
Amine 1. Lithium aluminum hydride (LAH) (0.77 g, 20.3 mmol) was added slowly to a stirred solution of amide (4) (1.43 g, 4.68 mmol) in THF (80 mL) at 0 °C. The solution was then heated to reflux for 24 h. The solution was cooled down to room temperature and then poured into an ice-water mixture. The resultant mixture was then extracted with CH2Cl2 (50 mL × 3) and the combined extracts were washed with brine, dried (MgSO4), filtered and evaporated in vacuo to give the crude product which was purified by flash chromatography on silica gel (hexane/EtOAc/Et3N = 240/120/1) to afford the amine (1) (1.25 g, 92%) as a colorless liquid. Rf: 0.35 (hexane/EtOAc/Et3N = 240/120/1). 1H NMR: δ 0.73–0.88 (2 H, m, CH2), 0.99–1.23 (3 H, m), 1.28–1.42 (1 H, m), 1.50–1.83 (8 H, m), 2.34 (2 H, d, J = 6.8, CHCH2NH), 2.45–2.58 (4 H, m, ArCH2CH2 and ArCH2CH2), 3.65 (6 H, s, CH3), 6.20 (1 H, d, J = 2.0, ArH), 6.26 (2 H, d, J = 2.0, ArH). 13C NMR: δ 25.8, 26.4, 31.1, 31.2, 33.7, 37.6, 49.3, 54.8, 56.4, 97.4, 106.1, 144.2, 160.5. ESI-MS: m/z 292 ([M + H]+, 100%). HRESI-MS: calcd m/z for C18H30NO2: 292.2271, found: 292.2276 ([M + H]+, 100%).
Ammonium salt 1-H·PF6. To a stirred solution of amine (1) (1.25 g, 4.29 mmol) in a solvent mixture of CH3CN/CH2Cl2 (3:1) (20 mL), excess hexafluorophosphoric acid (HPF6) (1.20 mL, 8.73 mmol) was added dropwise at 0 °C. The reaction mixture was stirred for 3 h and then water (12 mL) was added to the mixture. The mixture was concentrated under reduced pressure and then extracted with CH2Cl2 (50 mL × 3). The combined extracts were washed with brine, dried (MgSO4), filtered and evaporated in vacuo to afford the ammonium salt 1-H·PF6 (1.50 g, 80%) as a white solid. M.p.: 173.6–174.8 °C. 1H NMR (CD3CN): δ 0.87–1.07 (2 H, m, CH2), 1.08–1.34 (3 H, m), 1.57–1.82 (6 H, m), 1.93–2.12 (2 H, m, CH2), 2.61 (2 H, t, J = 7.6, ArCH2), 2.75–2.90 (2 H, m, CHCH2NH2), 2.93–3.10 (2 H, m, CH2CH2NH2), 3.76 (6 H, s, CH3), 6.34 (1 H, s, ArH), 6.41 (2 H, s, ArH), 6.61 (2 H, br s, NH2). 13C NMR (CD3CN): δ 26.5, 26.8, 28.1, 30.9, 33.4, 35.8, 49.4, 55.2, 56.0, 99.1, 107.5, 144.1, 162.2. ESI-MS: m/z 292 ([M–PF6]+, 100%). HRESI-MS: calcd m/z for C18H30NO2: 292.2271, found: 292.2265 ([M–PF6]+, 100%).
Typical Synthesis of [2]Pseudorotaxane. A solution of 1-H·PF6 (17.5 mg, 0.04 mmol) and DB24C8 (35.9 mg, 0.08 mmol) in a typical solvent (1.5 mL) was heated at 40 or 70 °C in a sealed tube for 4 days and then concentrated in vacuo. For an alternative method, a solution of 1-H·PF6 (120 mg, 0.27 mmol) and DB24C8 (40 mg, 0.089 mmol) in MeCN (3 mL) was heated at 70 °C in a sealed tube for 4 days and then concentrated in vacuo. The resulting residue was purified by flash chromatography on silica gel (CH2Cl2/THF = 7/1) to afford the [2]pseudorotaxane DB24C81-H·PF6 (34–36 mg, 46%) as a white solid. M.p. was not determined due to thermal instability. Rf: 0.65 (CH2Cl2/THF = 7/1). 1H NMR (CD3CN): δ 0.56–0.72 (2 H, m, CH2), 0.78–0.93 (3 H, m), 1.40–1.55 (6 H, m, CH2), 1.79–1.91 (2 H, m, CH2), 2.41 (2 H, t, J = 7.2, ArCH2), 3.08–3.18 (2 H, m, CHCH2NH2), 3.27–3.38 (2 H, m, CH2CH2NH2), 3.58–3.70 (8 H, m, CH2CH2OCH2), 3.72 (6 H, s, CH3), 3.78–3.89 (8 H, m, ArOCH2CH2), 4.08–4.23 (8 H, m, ArOCH2CH2), 6.16 (2 H, d, J = 2.0, ArH), 6.31 (2 H, d, J = 2.0, ArH), 6.55–6.84 (2 H, br s, NH2), 6.90–7.02 (8 H, m, catechol ArH). 13C NMR (CD3CN): δ 26.1, 26.5, 28.7, 30.9, 33.4, 36.3, 49.2, 55.4, 55.9, 69.0, 71.3, 71.9, 98.7, 107.4, 113.6, 122.4, 144.0, 148.5, 162.0. ESI-MS: m/z 740 ([M–PF6]+, 100%). HRESI-MS: calcd m/z for C42H62NO10: 740.4368, found: 740.4359 ([M–PF6]+, 100%).
Typical [2]Pseudorotaxane Dissociation. For dissociation with bases, [2]pseudorotaxane DB24C81-H·PF6 (5 mg) was dissolved in CD3CN (0.7 mL) in a NMR tube followed by the addition of 3 drops of Et3N (TEA) and 3 drops of iPr2EtN (DIPEA) to the solution. 1H NMR spectra were recorded over time at 298 K. For dissociation with solvent, [2]pseudorotaxane DB24C81-H·PF6 (5 mg) was dissolved in CD3SOCD3 (0.7 mL) in a NMR tube. 1H NMR spectra were recorded over time at 298 K.

4. Conclusions

In summary, a [2]pseudorotaxane DB24C81-H·PF6 which was synthesized and isolated from a slippage approach, was found to be stable at room temperature. The [2]pseudorotaxane DB24C81-H·PF6 was characterized by NMR spectroscopy and mass spectrometry. The dissociation rate of [2]pseudorotaxane DB24C81-H·PF6 could be tuned by amine bases and DMSO solvent at room temperature from which the supramolecular interactions between the crown ether DB24C8 and the ammonium can be tuned. When the crown ether DB24C8 was lost its noncovalent interactions with the ammonium/amine, it could be detached from the dumbbell 1-H·PF6 via the molecular flipping of between the chair-boat-chair forms of the cyclohexyl ring. In particular, the dissociation of [2]pseudorotaxane DB24C81-H·PF6 with amine bases possessed a long and sustained, complete dissociation over 60 days. The use of DMSO possessed relatively fast and complete dissociation (24 h). The selectively tunable dissociation rates by potentially varying the ratios of stimuli for macrocycle extrusion from the novel [2]pseudorotaxane provide avenues for constructing novel functional nanovalves that sustained release substrates for a much longer period of time.

Acknowledgments

We acknowledge the financial support by the General Research Fund (201213) by the Research Grants Council of Hong Kong. This work is also partially supported by a grant from the University Grants Committee of HKSAR (Area of Excellence Scheme AoE/P-03/08).

Author Contributions

Conceived and designed the experiments: Ken Cham-Fai Leung, Kwun-Ngai Lau, Wing-Yan Wong. Performed the experiments: Kwun-Ngai Lau, Wing-Yan Wong. Analyzed the data: Ken Cham-Fai Leung, Kwun-Ngai Lau, Wing-Yan Wong. Contributed reagents/materials/analysis tools: Ken Cham-Fai Leung. Contributed to the writing of the manuscript: Ken Cham-Fai Leung, and Wing-Yan Wong.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ashton, P.R.; Baxter, I.; Fyfe, M.C.T.; Raymo, F.M.; Spencer, N.; Stoddart, J.F.; White, A.J.P.; Williams, D.J. Rotaxane or Pseudorotaxane? That Is The Question! J. Am. Chem. Soc. 1998, 120, 2297–2307. [Google Scholar]
  2. Affeld, A.; Hübner, G.M.; Seel, C.; Schalley, C.A. Rotaxane or Pseudorotaxane? Effects of Small Structural Variation on the Deslipping Kinetics of Rotaxanes with Stopper Groups of Intermediate Size. Eur. J. Org. Chem 2001, 2001, 2877–2890. [Google Scholar]
  3. Aricó, F.; Badjic, J.D.; Cantrill, S.J.; Flood, A.H.; Leung, K.C.-F.; Liu, Y.; Stoddart, J.F. Templated Synthesis of Interlocked Molecules. Top. Curr. Chem. 2005, 249, 203–259. [Google Scholar]
  4. Hirose, K.; Nakamura, Y.; Tobe, Y. Remarkable Effects of Chirality on Deslipping Reactions of Diasteromeric Rotaxanes and Relevant Mechanism Involving Pre-Equilibrium. Org. Lett. 2009, 11, 145–147. [Google Scholar] [CrossRef] [PubMed]
  5. Leung, K.C.-F.; Chak, C.-P.; Lo, C.-M.; Wong, W.-Y.; Xuan, S.; Cheng, C.H.K. pH-Controllable Supramolecular Systems. Chem. Asian J. 2009, 4, 364–381. [Google Scholar] [CrossRef] [PubMed]
  6. Wong, W.-Y.; Lee, S.-F.; Chan, H.-S.; Mak, T.C.M.; Wong, C.-H.; Huang, L.-S.; Stoddart, J.F.; Leung, K.C.-F. Recognition between V- and Dumbbell-shaped Molecules. RSC Adv. 2013, 3, 26382–26390. [Google Scholar] [CrossRef]
  7. Nijhuis, C.A.; Ravoo, B.J.; Huskens, J.; Reinhoudt, D.N. Redox-active Supramolecular Systems. Coord. Chem. Rev. 2007, 251, 1761–1780. [Google Scholar] [CrossRef]
  8. Liu, Y.; Flood, A. H.; Stoddart, J.F. Thermally and Electrochemically Controllable Self-Complexing Molecular Switches. J. Am. Chem. Soc. 2004, 126, 9150–9151. [Google Scholar] [CrossRef] [PubMed]
  9. Fernando, I.R.; Bairu, S.G.; Guda, R.; Mezei, G. Single-Color Pseudorotaxane-based Temperature Sensing. New J. Chem. 2010, 34, 2097–2100. [Google Scholar] [CrossRef]
  10. Leung, K.C.-F.; Wong, W.-Y.; Aricó, F.; Haussmann, P.C.; Stoddart, J.F. The Stability of Imine-containing Dynamic [2]Rotaxane to Hydrolysis. Org. Biomol. Chem. 2010, 8, 83–89. [Google Scholar] [CrossRef] [PubMed]
  11. Wong, W.-Y.; Leung, K.C.-F.; Stoddart, J.F. Self-assembly, Stability Quantification, Controlled Molecular Switching, and Sensing Properties of an Anthracene-containing Dynamic [2]Rotaxane. Org. Biomol. Chem. 2010, 8, 2332–2343. [Google Scholar] [CrossRef] [PubMed]
  12. Leung, K.C.-F.; Nguyen, T.D.; Stoddart, J.F.; Zink, J.I. Supramolecular Nanovalves Controlled by Proton Abstraction and Competitive Binding. Chem. Mater. 2006, 18, 5919–5928. [Google Scholar] [CrossRef]
  13. Saha, S.; Leung, K.C.-F.; Nguyen, T.D.; Stoddart, J.F.; Zink, J.I. Nanovalves. Adv. Funct. Mater. 2007, 17, 685–693. [Google Scholar] [CrossRef]
  14. Cotí, K.K.; Belowich, M.E.; Liong, M.; Ambrogio, M.W.; Lau, Y.A.; Khatib, H.A.; Zink, J.I.; Khashab, N.M.; Stoddart, J.F. Mechanised Nanoparticles for Drug Delivery. Nanoscale 2009, 1, 16–39. [Google Scholar] [CrossRef] [PubMed]
  15. Rowan, S.J.; Cantrill, S.J.; Cousins, G.R.L.; Sanders, J.K.M.; Stoddart, J.F. Dynamic Covalent Chemistry. Angew. Chem. Int. Ed. 2002, 41, 898–952. [Google Scholar] [CrossRef]
  16. Grigoras, M.; Catanescu, C.O. Imine Oligomers and Polymers. J. Macromol. Sci. Polym. Rev. 2004, C44, 131–173. [Google Scholar] [CrossRef]
  17. Meyer, C.D.; Joiner, C.S.; Stoddart, J.F. Template-directed Synthesis Employing Reversible Imine Bond Formation. Chem. Soc. Rev. 2007, 36, 1705–1723. [Google Scholar] [CrossRef] [PubMed]
  18. Haussmann, P.C.; Stoddart, J.F. Synthesizing Interlocked Molecules Dynamically. Chem. Rec. 2009, 9, 136–154. [Google Scholar] [CrossRef] [PubMed]
  19. Herrmann, A. Dynamic Mixtures and Combinatorial Libraries: Imines as Probes for Molecular Evolution at the Interface between Chemistry and Biology. Org. Biomol. Chem. 2009, 7, 3195–3204. [Google Scholar] [CrossRef] [PubMed]
  20. Maeda, T.; Otsuka, H.; Takahara, A. Dynamic Covalent Polymers: Reorganizable Polymers with Dynamic Covalent Bonds. Prog. Polym. Sci. 2009, 34, 581–604. [Google Scholar] [CrossRef]
  21. Klivansky, L.M.; Koshkakaryan, G.; Cao, D.; Liu, Y. Linear pi-acceptor-templated Dynamic Clipping to Macrobicycles and [2]Rotaxanes. Angew. Chem. Int. Ed. 2009, 48, 4185–4189. [Google Scholar] [CrossRef]
  22. Yin, J.; Dasgupta, S.; Wu, J. Template-directed Synthesis of Rotaxanes by Clipping: The Effect of the Dialkylammonium Recognition Sites. Org. Lett. 2010, 12, 1712–1715. [Google Scholar] [CrossRef] [PubMed]
  23. Leung, K.C.-F.; Lau, K.-N. Self-assembly and Thermodynamic Synthesis of Rotaxane Dendrimers and Related Structures. Polym. Chem. 2010, 1, 988–1000. [Google Scholar] [CrossRef]
  24. Ho, W.K.-W.; Lee, S.-F.; Wong, C.-H.; Zhu, X.-M.; Kwan, C.-S.; Chak, C.-P.; Mendes, P.M.; Cheng, C.H.K.; Leung, K.C.-F. Type III-B Rotaxane Dendrimers. Chem. Commun. 2013, 49, 10781–10783. [Google Scholar] [CrossRef]
  25. Harrison, I.T. The Effect of Ring Size on Threading Reactions of Macrocycles. J. Chem. Soc. Chem. Commun. 1972, 231, 231–232. [Google Scholar] [CrossRef]
  26. Raymo, F.M.; Stoddart, J.F. Slippage–A Simple and Efficient Way to Self-assemble [n]Rotaxanes. Pure Appl. Chem. 1997, 69, 1987–1997. [Google Scholar] [CrossRef]
  27. Elizarov, A.M.; Chang, T.; Chiu, S.-H.; Stoddart, J.F. Self-assembly of Dendrimers by Slippage. Org. Lett. 2002, 4, 3565–3568. [Google Scholar] [CrossRef] [PubMed]
  28. Hsueh, S.-Y.; Lai, C.-C.; Liu, Y.-H.; Wang, Y.; Peng, S.-M.; Chiu, S.-H. Protecting a Squaraine Near-IR Dye Through Its Incorporation in a Slippage-derived [2]Rotaxane. Org. Lett. 2007, 9, 4523–4526. [Google Scholar]
  29. Schill, G. Catenanes, Rotaxanes, and Knots; Academic Press: New York, NY, USA, 1971; p. 3. [Google Scholar]
  30. McConnell, A.J.; Beer, P.D. Kinetic Studies Exploring the Role of Anion Templation in the Slippage Formation of Rotaxane-Like Structures. Chem. Eur. J. 2011, 17, 2724–2733. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, M.; Xu, D.; Yan, X.; Chen, J.; Dong, S.; Zheng, B.; Huang, F. Self-Healing Supramolecular Gels Formed by Crown Ether Based Host-Guest Interactions. Angew. Chem. Int. Ed. 2012, 51, 7011–7015. [Google Scholar] [CrossRef]
  32. Wang, F.; Han, C.; He, C.; Zhou, Q.; Zhang, J.; Wang, C.; Li, N.; Huang, F. Self-sorting Organization of Two Heteroditopic Monomers to Supramolecular Alternating Copolymers. J. Am. Chem. Soc. 2008, 130, 11254–11255. [Google Scholar] [CrossRef] [PubMed]
  33. Ji, X.; Yao, Y.; Li, J.; Yan, X.; Huang, F. A Supramolecular Cross-linked Conjugated Polymer Network for Multiple Fluorescent Sensing. J. Am. Chem. Soc. 2013, 135, 74–77. [Google Scholar] [CrossRef] [PubMed]
  34. Dong, S.; Zheng, B.; Xu, D.; Yan, X.; Zhang, M.; Huang, F. A Crown Ether Appended Super Gelator with Multiple Stimulus Responsiveness. Adv. Mater. 2012, 24, 3191–3195. [Google Scholar] [CrossRef] [PubMed]
  35. Dong, S.; Luo, Y.; Yan, X.; Zheng, B.; Ding, X.; Yu, Y.; Ma, Z.; Zhao, Q.; Huang, F. A Dual-Responsive Supramolecular Polymer Gel Formed by Crown Ether Based Molecular Recognition. Angew. Chem. Int. Ed. 2011, 50, 1905–1909. [Google Scholar] [CrossRef]
  36. Cantrill, S.J.; Fulton, D.A.; Heiss, A.M.; Pease, A.R.; Stoddart, J.F.; White, A.J.P.; Williams, D.J. The Influence of Macrocyclic Polyether Constitution Upon Ammonium Ion/Crown Ether Recognition Processes. Chem. Eur. J. 2000, 6, 2274–2287. [Google Scholar] [CrossRef] [PubMed]
  37. Bolla, M.A.; Tiburcio, J.; Loeb, S.J. Characterization of a Slippage Stopper for the 1,2-Bis(pyridinium)ethane-[24]crown-8 ether [2]Pseudorotaxane Motif. Tetrahedron 2008, 64, 8423–8428. [Google Scholar] [CrossRef]
  38. Linnartz, P.; Bitter, S.; Schalley, C.A. Deslipping of Ester Rotaxanes: A Cooperative Interplay of Hydrogen Bonding with Rotational Barriers. Eur. J. Org. Chem. 2003, 24, 4819–4829. [Google Scholar] [CrossRef]
  39. Bauta, W.E.; Lovett, D.P.; Cantrell, W.R.; Burke, B.D. Formal Synthesis of Angiogenesis Inhibitor NM-3. J. Org. Chem. 2003, 68, 5967–5973. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Leung, K.C.-F.; Lau, K.-N.; Wong, W.-Y. Revisiting the Formation and Tunable Dissociation of a [2]Pseudorotaxane Formed by Slippage Approach. Int. J. Mol. Sci. 2015, 16, 8254-8265. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms16048254

AMA Style

Leung KC-F, Lau K-N, Wong W-Y. Revisiting the Formation and Tunable Dissociation of a [2]Pseudorotaxane Formed by Slippage Approach. International Journal of Molecular Sciences. 2015; 16(4):8254-8265. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms16048254

Chicago/Turabian Style

Leung, Ken Cham-Fai, Kwun-Ngai Lau, and Wing-Yan Wong. 2015. "Revisiting the Formation and Tunable Dissociation of a [2]Pseudorotaxane Formed by Slippage Approach" International Journal of Molecular Sciences 16, no. 4: 8254-8265. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms16048254

Article Metrics

Back to TopTop