Co-reporter:Terry L. Price Jr., Hanlie R. Wessels, Carla Slebodnick, and Harry W. Gibson
The Journal of Organic Chemistry August 4, 2017 Volume 82(Issue 15) pp:8117-8117
Publication Date(Web):July 17, 2017
DOI:10.1021/acs.joc.7b01389
Pyridinium bis(trifluoromethylsulfonyl)imide (PyTFSI)-templated syntheses of 2,6-pyridyl cryptands of cis(4,4′)-dibenzo-30-crown-10 (3a), the p-bromobenzyloxy derivative 3b, bis(m-phenylene)-32-crown-10 (5), cis(4,4′)-dibenzo-27S-crown-9 (7), cis(4,4′)-dibenzo-27L-crown-9 (9), and cis(4,4′)-dibenzo-24-crown-8 (11) are reported. Here we provide a fast (12 h), high-yielding (89%, 74%, 80%, and 62% for 3a, 3b, 5, and 9, respectively) templation method without the use of a syringe pump. The yields for 7 (19%) and 11 (26%) were lower than with the previous pseudo-high-dilution method, indicating ineffective templation in these cases. Coupled with our previously developed templated syntheses of dibenzo crown ethers, this protocol makes powerful cryptand hosts readily available in gram quantities in good yields from methyl 4(or 3)-hydroxy-3(or 4)-benzyloxybenzoate.
Co-reporter:Adam M.-P. Pederson, Terry L Price Jr., Carla Slebodnick, Daniel V. Schoonover, and Harry W. Gibson
The Journal of Organic Chemistry August 18, 2017 Volume 82(Issue 16) pp:8489-8489
Publication Date(Web):July 13, 2017
DOI:10.1021/acs.joc.7b01242
The two isomers 6 and 9 of cis(4,4′-)-dicarbomethoxydibenzo-27-crown-9 with tri- and tetra-(ethyleneoxy) linkages transposed were synthesized regiospecifically in high yields (94 and 92%, respectively) by the Wang–Pederson–Wessels (WPW) protocol and were converted via the corresponding diols 7 and 10 to the corresponding pyridyl cryptands 3 and 4 by reaction with pyridine-2,6-dicarbonyl chloride. As expected from Corey–Pauling–Koltun (CPK) models, the cryptand with the tri(ethyleneoxy) arm para to the ester linkages, “short-armed” cryptand 3, did display a higher binding constant (Ka = 2.4 × 105 M–1) with paraquats than the analogous dibenzo-30-crown-10-based cryptand previously studied; however, the effect was only twofold. Its binding to diquat was reduced by an order of magnitude (1.5 × 105 M–1), as expected on the basis of its narrower cavity. Also as expected, the cryptand with the tetra(ethyleneoxy) arm para to the ester linkages, “long-armed” cryptand 4, possessed diminished binding with both paraquats and diquat relative to the 30-crown-based analogue; in these systems, 2:1 H:G complexes were also detected by mass spectrometry. A crystal structure is reported for 3·DQ(PF6)2.
Co-reporter:Daniel V. Schoonover, Harry W. Gibson
Tetrahedron Letters 2017 Volume 58(Issue 3) pp:242-244
Publication Date(Web):18 January 2017
DOI:10.1016/j.tetlet.2016.12.014
•Tosylates are routinely used electrophiles in organic synthesis.•The synthesis of tosylates usually employs an excess of tosyl chloride.•Tosyl chloride is often a difficult impurity to remove from tosylates.•Addition of filter paper and sonication after tosylation removes all unreacted tosyl chloride.•This facile method is efficient and economical.Excess tosyl chloride used in the tosylation of alcohols is quickly and easily removed by reacting it with cellulosic materials, e.g., filter paper, and filtering.
Co-reporter:Zhenbin Niu, Terry L. Price Jr., Carla Slebodnick, Harry W. Gibson
Tetrahedron Letters 2016 Volume 57(Issue 1) pp:60-63
Publication Date(Web):6 January 2016
DOI:10.1016/j.tetlet.2015.11.061
The readily prepared dipicolinate, diquinaldate, and dinaphthyridate ester derivatives 1c–1e of bis(m-phenylene)-32-crown-10 (BMP32C10) diol 1b demonstrated improved association constants with diquat (3) compared with BMP32C10 itself (1a). Importantly, the association constant of dinaphthyridyl BMP32C10 (1e) with diquat is even higher than some BMP32C10-based cryptands, which are synthetically difficult to obtain. An unusual pseudocryptand-type alternating poly[2]pseudorotaxane was formed between 1e and diquat in the solid state.
Co-reporter:U Hyeok Choi;Anuj Mittal;Terry L. Price Jr.;Ralph H. Colby
Macromolecular Chemistry and Physics 2016 Volume 217( Issue 11) pp:1270-1281
Publication Date(Web):
DOI:10.1002/macp.201500424
Co-reporter:Hanlie R. Wessels, Harry W. Gibson
Tetrahedron 2016 Volume 72(Issue 3) pp:396-399
Publication Date(Web):21 January 2016
DOI:10.1016/j.tet.2015.11.055
Dibenzo-30-crown-10 (DB30C10, 1c), dibenzo-24-crown-8 (DB24C8, 1a), 4-carbomethoxydibenzo-24crown-8 (1b) and a new crown ether, 4-carbomethoxydibenzo-30-crown-10 (1d), were synthesized by a simple, high yielding, three-step method. Potassium ions from the highly soluble KPF6 were employed to template the cyclization step, which resulted in very high isolated yields (80–90%).
Co-reporter:Minjae Lee
Journal of Polymer Science Part A: Polymer Chemistry 2016 Volume 54( Issue 11) pp:1647-1658
Publication Date(Web):
DOI:10.1002/pola.28022
ABSTRACT
Rotaxane-type hyperbranched polymers are synthesized for the first time from A2B type semi-rotaxane monomers formed in situ via complexation of bis(m-phenylene)-32-crown-10 dimethanol (1) and two paraquat ω-n-alkylenecarboxylic acid derivatives with tris(p-t-butylphenyl)methylphenylalkylene stoppers (8 and 9). Rotaxane and taco complexes exist in solutions of the hyperbranched polyesters in CD3CN/CDCl3 as confirmed by NMR spectroscopy, but the taco complexes, which derive from non-rotaxanated paraquat units, disappear in DMSO-d6. NMR spectroscopy indicates the portion of rotaxanes strongly interlocked by the environment (inner rotaxanes) is larger in HP1•9, which has longer alkylene spacers, perhaps indicating a higher degree of polymerization. The molecular size increases upon formation of the hyperbranched polymers are confirmed by dynamic light scattering and by viscometry. As with covalent hyperbranched polymers a number of potential applications exist; the unique mechanically linked character and the presence of uncomplexed host and guest moieties foreshadow the use of such systems for their responses to external stimuli with the added benefit of providing molecular recognition sites useful as delivery vehicles. Use of other host-guest motifs to form the semirotaxane A2B monomers is possible and complementary systems with higher binding constants will enable efficient syntheses of high molecular weight, mechanically linked hyperbranched polymers. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016, 54, 1647–1658
Co-reporter:Minjae Lee;U Hyeok Choi;Ralph H. Colby
Macromolecular Chemistry and Physics 2015 Volume 216( Issue 3) pp:344-349
Publication Date(Web):
DOI:10.1002/macp.201400426
Co-reporter:Susmita Bhattacharjee, Caiguo Gong, Jason W. Jones, Harry W. Gibson
Polymer 2015 Volume 81() pp:99-110
Publication Date(Web):16 December 2015
DOI:10.1016/j.polymer.2015.10.069
•Crown ether monoalcohol and a 4,4′-bipyridinium diacid chloride formed a pseudorotaxane in situ at low temperature.•Addition of pyridine and increasing temperature led to a mechanically linked hyperbranched polyester.•Two components: diester (two macrocycles), monoester (one). Precipitation from DMSO removed any non-polymeric moieties.•NMR: macrocycles complex focal point bipyridinium units (acetone); no complexes (DMSO), only mechanically linked moieties.•MS confirmed the presence of hyperbranched structures with high molar masses, up to 16.3 kDa.The interaction of the diacid chloride (2) of N,N’-bis(5″-carboxypentyl)-4,4′bipyridnium bis(hexafluorophosphate) (1) with m-phenylene-5′-hydroxymethyl-1′,3′-phenylene-32-crown-10 (3) at −70 °C led to the in situ formation of a pseudorotaxane 4 as an AB2 monomer, which was allowed to polymerize by raising the temperature and adding a catalytic amount of pyridine. The resulting hyperbranched polymer was held together via mechanical bonds formed by threading of the crown ether units by the bipyridinium (paraquat) moieties. The polymer was initially precipitated into aqueous NH4PF6 to remove any unreacted, water soluble diacid 1 to form 5b and then precipitated from DMSO into aqueous NH4PF6 to remove any pyridinium monoester 8 complexed with crown ether units in the polymer, leading to 6. The hyprerbranched polyester was characterized by 1H NMR and NOESY spectroscopies, FAB and MALDI-TOF mass spectrometry and viscometry. NMR spectroscopy confirmed complete esterification of the crown ether alcohol; 5b contained 60% monoesterified paraquat units (8) and 40% diester moieties (9), while 6 contained only 24% monoester moieties, which existed primarily as “focal point” units (10), but also as uncomplexed linker units (11). A variety of mechanically linked structures containing cyclic units is possible, including 17–19. Mass spectrometry on 6 detected masses up to 16.3 kDa, corresponding to a degree of polymerization of 11 units of the diester repeat unit in 17 (and/or isomeric structures).
Co-reporter:M. Arunachalam, Harry W. Gibson
Progress in Polymer Science 2014 Volume 39(Issue 6) pp:1043-1073
Publication Date(Web):June 2014
DOI:10.1016/j.progpolymsci.2013.11.005
This review highlights developments since 2008 in the field of polypseudorotaxanes and polyrotaxanes. Progress in synthetic polyrotaxane chemistry has resulted in the preparation of numerous functionalized polymers for various applications in areas such as molecular machines, stimuli responsive materials, supramolecular gels and molecular sensors. This new genre of supramolecular polymers is advancing rapidly with several groups developing novel materials with unique characteristics.
Co-reporter:Mingming Zhang, Xuzhou Yan, Feihe Huang, Zhenbin Niu, and Harry W. Gibson
Accounts of Chemical Research 2014 Volume 47(Issue 7) pp:1995-2005
Publication Date(Web):May 8, 2014
DOI:10.1021/ar500046r
This Account mainly focuses on the application of cryptands in the construction of mechanically interlocked molecules such as rotaxanes and catenanes, and stimuli-responsive host–guest systems such as molecular switches and supramolecular polymers due to their good host–guest properties. These cryptands are bicyclic derivatives of crown ethers, including dibenzo-24-crown-8, bis(m-phenylene)-26-crown-8, dibenzo-30-crown-10, and bis(m-phenylene)-32-crown-10. The length of the third arm has a very important influence on the binding strength of these cryptands with organic guests, because it affects not only the size fit between the host and the guest but also the distances and angles that govern the strengths of the noncovalent interactions between the host and the guest. For example, for bis(m-phenylene)-32-crown-10-based cryptands, a third arm of nine atoms is the best. The environmental responsiveness of these cryptand-based host–guest systems arises from either the crown ether units or the third arms. For example, a dibenzo-24-crown-8 unit introduces potassium cation responsiveness and an azobenzene group on the third arm imbues photoresponsiveness. We believe that studies on stimuli-responsive host–guest systems based on cryptands and organic guests will contribute significantly to future research on molecular devices, supramolecular polymers, and other functional supramolecular materials.
Co-reporter:Harry W. Gibson, Ya Xi Shen, Mukesh C. Bheda, Caiguo Gong
Polymer 2014 Volume 55(Issue 15) pp:3202-3211
Publication Date(Web):25 June 2014
DOI:10.1016/j.polymer.2014.05.007
Reaction of di(p-isocyanatophenyl)methane (MDI, 4) with N,N′-di(2-hydroxyethyl)- (1b) or N,N′-di[2-(2′-hydroxyethoxy)ethyl]-4,4′-bipyridinium di(hexafluorophosphate) (1e) and other diols [oligo(ethylene glycol)s and poly(tetramethylene oxide)s] in the presence of bis(p-phenylene)-34-crown-10 (2) afforded polyurethane (pseudo)rotaxanes as statistical (7P or 7R) and segmented analogs 10P (P = pseudorotaxane, R = rotaxane). In 7R a bulky alcohol was incorporated at the chain ends and in 13R a bulky diol as in-chain units to form polyrotaxanes and preclude the possibility of dethreading. The crown ether 2 in 10P and 13R was shown by 1H NMR spectroscopy to be shuttling between the viologen (paraquat) and urethane sites; in DMSO the crown ether prefers the urethane site, probably because of H-bonding with the N–H moieties and complexation of the pyridinium site by the dipolar solvent, while in acetone at low temperatures the viologen site is preferred by the crown ether, with ΔH = −6.91 kcal/mol and ΔS = −22.9 eu for 13R.
Co-reporter:Harry W. Gibson, Michael A.G. Berg, Terry L. Price Jr., Zhenbin Niu, Minjae Lee, Mason A. Rouser, Carla Slebodnick
Tetrahedron 2014 70(35) pp: 5904-5918
Publication Date(Web):
DOI:10.1016/j.tet.2014.06.027
Co-reporter:Harry W. Gibson, Michael A.G. Berg, Terry L. Price Jr., Zhenbin Niu, Minjae Lee, Mason A. Rouser, Jennifer Clifton Dickson, Carla Slebodnick
Tetrahedron 2013 69(32) pp: 6727
Publication Date(Web):
DOI:10.1016/j.tet.2013.06.035
Co-reporter:Harry W. Gibson, Hong Wang, Zhenbin Niu, Carla Slebodnick, Lev N. Zhakharov, and Arnold L. Rheingold
Macromolecules 2012 Volume 45(Issue 3) pp:1270-1280
Publication Date(Web):January 13, 2012
DOI:10.1021/ma202373x
Attempts to incorporate a tetralactam (3a) into a side-chain polyrotaxane by formation of its complex with p-tritylphenolate and reaction with poly(vinybenzyl chloride) failed, probably due to steric hindrance. However, three new small molecule model rotaxanes (acetal 11a and ethers 14a and 14b) were prepared based on this macrocycle, and one (14b) was characterized by X-ray crystallography. Attempts to prepare poly(ether rotaxane)s such as 18 and 20b via polymerizations in the presence of this tetralactam failed because the mild conditions required for complexation (nonpolar solvent, room temperature) are not sufficient for effective step-growth polymerizations, resulting in only oligomer formation. In an alternative approach, clipping of the tetralactam 23 onto a preformed poly(urethane rotaxane) (27) in CH2Cl2 was successful, and the original biphasic segmented polymer was converted to a single-phase polyrotaxane (28) with a significantly lower glass transition temperature.
Co-reporter:Harry W. Gibson, Michael A.G. Berg, Terry L. Price Jr., Zhenbin Niu, Minjae Lee, Mason A. Rouser, Jennifer Clifton Dickson, Carla Slebodnick
Tetrahedron 2012 68(38) pp: 8052-8067
Publication Date(Web):
DOI:10.1016/j.tet.2012.06.008
Co-reporter:Zhenbin Niu ; Feihe Huang
Journal of the American Chemical Society 2011 Volume 133(Issue 9) pp:2836-2839
Publication Date(Web):February 10, 2011
DOI:10.1021/ja110384v
Two novel bis(m-phenylene)-32-crown-10-based cryptands, one bearing covalent linkages and the other metal-complex linkages, were designed and prepared. By self-assembly of these biscryptands, which can be viewed as AA monomers, and a bisparaquat, which can be viewed as a BB monomer, AA−BB-type linear supramolecular polymers with relatively high molecular weights were successfully prepared.
Co-reporter:Minjae Lee;U. Hyeok Choi;David Salas-de la Cruz;Anuj Mittal;Karen I. Winey;Ralph H. Colby
Advanced Functional Materials 2011 Volume 21( Issue 4) pp:708-717
Publication Date(Web):
DOI:10.1002/adfm.201001878
Abstract
New bis(ω-hydroxyalkyl)imidazolium and 1,2-bis[N-(ω-hydroxyalkyl)imidazolium]ethane salts are synthesized and characterized; most of the salts are room temperature ionic liquids. These hydroxyl end-functionalized ionic liquids are polymerized with diacid chlorides, yielding polyesters containing imidazolium cations embedded in the main chain. By X-ray scattering, four polyesters are found to be semicrystalline at room temperature: mono-imidazolium-C11-sebacate-C6 (4e), mono-imidazolium-C11-sebacate-C11 (4c), bis(imidazolium)ethane-C6-sebacate-C6 (5a), and bis(imidazolium)ethane-C11-sebacate-C11 (5c), all with hexafluorophosphate counterions. The other imidazolium polyesters, including all those with bis(trifluoromethanesulfonyl)imide (TFSI−) counterions, are amorphous at room temperature. Room temperature ionic conductivities of the mono-imidazolium polyesters (4 × 10−6 to 3 × 10−5 S cm−1) are higher than those of the corresponding bis-imidazolium polyesters (4 × 10−9 to 8 × 10−6 S cm−1), even though the bis-imidazolium polyesters have higher ion concentrations. Counterions affect ionic conduction significantly; all polymers with TFSI− counterions have higher ionic conductivities than the hexafluorophosphate analogs. Interestingly, the hexafluorophosphate polyester, 1,2-bis(imidazolium)ethane-C11-sebacate-C11 (5c), displays almost 400-fold higher room temperature ionic conductivity (1.6 × 10−6 S cm−1) than the 1,2-bis(imidazolium)ethane-C6-sebacate-C6 analog (5a, 4.3 × 10−9 S cm−1), attributable to the differences in the semicrystalline structure in 5c as compared to 5a. These results indicate that semicrystalline polymers may result in high ionic conductivity in a soft (low glass tranition temperature, Tg) amorphous phase and good mechanical properties of the crystalline phase.
Co-reporter:Minjae Lee, U Hyeok Choi, Sungsool Wi, Carla Slebodnick, Ralph H. Colby and Harry W. Gibson
Journal of Materials Chemistry A 2011 vol. 21(Issue 33) pp:12280-12287
Publication Date(Web):13 Jul 2011
DOI:10.1039/C1JM10995B
A new class of organic ionic plastic crystals was discovered. A series of alkylene 1,2-bis[N-(N′-alkylimidazolium)] salts with Br− and PF6− anions was prepared and most of the ethylene (C2) linked compounds undergo multiple solid–solid phase transitions as determined by differential scanning calorimetry (DSC). The salts with longer spacers (C3 or C4) between the imidazolium units do not display any solid–solid transitions. The PF6− salts with terminal C10 and C12n-alkyl moieties are “classical” organic ionic plastic crystals by Timmermans' definition; they have low ΔSf (11 J K−1 mol−1 for C10 and 12 J K−1 mol−1 for C12). Additionally, the C2 linked bis-imidazolium salts with n-butyl, n-hexyl, n-heptyl and n-octyl termini all undergo two, three or four solid state transitions and exhibit ΔSf values in the range of 36–49 J K−1 mol−1, which are similar to those of other known ionic plastic crystalline materials. These materials were additionally characterized via ionic conductivity and solid state NMR. These ionic plastic crystals are presumably single-ion conductors, but the ionic conductivities appear to be too low for practical applications. The activation energy for conduction decreases as these compounds are heated through each solid–solid transition. The lack of any change in solid state 2H NMR spectra with temperature indicates that there is no change in phenyl ring flipping, suggesting no change in the imidazolium local environment. A resolution of this apparent dichotomy is perhaps that the counterions reside with the ethylene spacers between imidazolium moieties.
Co-reporter:Zhenbin Niu, Carla Slebodnick, and Harry W. Gibson
Organic Letters 2011 Volume 13(Issue 17) pp:4616-4619
Publication Date(Web):August 3, 2011
DOI:10.1021/ol201837x
The first pseudocryptand-type supramolecular [3]pseudorotaxane was designed and prepared via the self-assembly of a bispicolinate BMP32C10 derivative and a bisparaquat. The complexation behavior was cooperative. In addition, the complex comprised of the BMP32C10 derivative and a cyclic bisparaquat demonstrated strong binding; interestingly, a poly[2]pseudocatenane structure was formed in the solid state for the first time.
Co-reporter:Zhenbin Niu and Harry W. Gibson
Organic & Biomolecular Chemistry 2011 vol. 9(Issue 20) pp:6909-6912
Publication Date(Web):11 Aug 2011
DOI:10.1039/C1OB06299A
Via the self-assembly of two bis(meta-phenylene)-32-crown-10-based cryptands, bearing covalent and metal complex (ferrocene) linkages, with dimethyl paraquat, novel [3]pseudorotaxanes were formed statistically and anticooperatively, respectively.
Co-reporter:Zhenbin Niu, Carla Slebodnick, Feihe Huang, Hugo Azurmendi, Harry W. Gibson
Tetrahedron Letters 2011 Volume 52(Issue 48) pp:6379-6382
Publication Date(Web):30 November 2011
DOI:10.1016/j.tetlet.2011.09.054
A pseudocryptand-type [2]pseudorotaxane was formed via the self-assembly of a dipyridyl bis(meta-phenylene)-32-crown-10 (BMP32C10) derivative and a paraquat derivative. Due to the basicity of the pyridyl group, which forms the third pseudo-bridge of the pseudocryptand, this pseudorotaxane possesses two finite acid–base adjustable association constants.
Co-reporter:Minjae Lee, Robert B. Moore, and Harry W. Gibson
Macromolecules 2011 Volume 44(Issue 15) pp:5987-5993
Publication Date(Web):July 18, 2011
DOI:10.1021/ma201241t
The self-assembly of a polyester containing crown ether host repeat units and a paraquat-terminated polystyrene guest provides a supramolecular pseudorotaxane graft copolymer based on the bis(m-phenylene)-32-crown-10/paraquat recognition motif. NMR chemical shift changes in solution gave evidence of the complexation of the crown ether and paraquat moieties. Solutions of the two polymers at 1:1 host:guest stoichiometry (crown ether:paraquat), but only 3 mass % crown polymer, possessed an intrinsic viscosity 1.9 times that of the paraquat-terminated polystyrene and 2.7 times that of the polyester. Differential scanning calorimetry revealed only one glass transition for a solid from evaporation of a solution at 3:1 host:guest (crown ether:paraquat) stoichiometry; this indicated complete phase mixing, as confirmed by optical microscopy and small-angle laser light scattering.
Co-reporter:Dr. Harry W. Gibson;Dr. Jason W. Jones;Dr. Lev N. Zakharov;Dr. Arnold L. Rheingold;Dr. Carla Slebodnick
Chemistry - A European Journal 2011 Volume 17( Issue 11) pp:3192-3206
Publication Date(Web):
DOI:10.1002/chem.201002522
Abstract
Complexation of anions, cations and even ion pairs is now an active area of investigation in supramolecular chemistry; unfortunately it is an area fraught with complications when these processes are examined in low polarity organic media. Using a pseudorotaxane complex as an example, apparent Ka2 values (=[complex]/{[salt]o−[complex]}{[host]o−[complex]}) for pseudorotaxane formation from dibenzylammonium salts (2-X) and dibenzo-[24]crown-8 (1, DB24C8) in CDCl3/CD3CN 3:2 vary with concentration. This is attributable to the fact that the salt is ion paired, but the complex is not. We report an equilibrium model that explicitly includes ion pair dissociation and is based upon activities rather than molar concentrations for study of such processes in non-aqueous media. Proper analysis requires both a dissociation constant, Kipd, for the salt and a binding constant for interaction of the free cation 2+ with the host, Ka5; Ka5 for pseudorotaxane complexation is independent of the counterion (500 M−1), a result of the complex existing in solution as a free cation, but Kipd values for the salts vary by nearly two orders of magnitude from trifluoroacetate to tosylate to tetrafluoroborate to hexafluorophosphate anions. The activity coefficients depend on the nature of the predominant ions present, whether the pseudorotaxane or the ions from the salt, and also strongly on the molar concentrations; activity coefficients as low as 0.2 are observed, emphasizing the magnitude of their effect. Based on this type of analysis, a method for precise determination of relative binding constants, Ka5, for multiple hosts with a given guest is described. However, while the incorporation of activity coefficients is clearly necessary, it removes the ability to predict from the equilibrium constants the effects of concentration on the extent of binding, which can only be determined experimentally. This has serious implications for study of all such complexation processes in low polarity media.
Co-reporter:Harry W. Gibson;Melvin L. Rasco;Zhenbin Niu
Journal of Polymer Science Part A: Polymer Chemistry 2011 Volume 49( Issue 17) pp:3842-3851
Publication Date(Web):
DOI:10.1002/pola.24821
Abstract
An AA bisphenolic monomer (13), an AB activated fluoro-phenolic monomer (14), and an AB2 alcohol-diester monomer precursor (17), all based on the isoquinoline nucleus, were prepared using Reissert compound chemistry. Additionally, model reactions established the efficiency of condensation of isoquinoline Reissert compounds with dihaloalkanes, providing evidence of the potential of this reaction to produce high molecular weight polymers, which was subsequently realized. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011
Co-reporter:Minjae Lee, U Hyeok Choi, Ralph H. Colby, and Harry W. Gibson
Chemistry of Materials 2010 Volume 22(Issue 21) pp:5814
Publication Date(Web):October 18, 2010
DOI:10.1021/cm101407d
Polymerizable imidazolium acrylates and their polymers with pendant imidazolium cations were synthesized with hexafluorophosphate and bis(trifluoromethanesulfonyl)imide counterions and characterized using calorimetry and dielectric spectroscopy. The ionic polymers containing a diethyleneoxy unit as an N-substituent on the imidazolium cation display higher ionic conductivities than the analogous N-n-butyl polymers. Using a physical model of electrode polarization, we separate the conductivity of single-ion conductors into number density of conducting ions p and their mobility μ. The monomers invariably possess higher conducting ion number density than the polymers, owing to the cation being part of the polymer, but p is insensitive to the N-substituent. In contrast, the diethyleneoxy N-substituent imparts higher mobility than the n-butyl N-substituent, for both monomers and polymers, owing to a lower binding energy between the imidazolium and the counteranions, which is not directly reflected in glass transition temperatures.
Co-reporter:Minjae Lee, Zhenbin Niu, Carla Slebodnick and Harry W. Gibson
The Journal of Physical Chemistry B 2010 Volume 114(Issue 21) pp:7312-7319
Publication Date(Web):May 6, 2010
DOI:10.1021/jp102370j
A series of N,N-alkylene bis(N′-alkylimidazolium) salts with various anions was prepared and characterized. The hydrogen-bonding abilities and ion-pairing strengths of the salts in solution were varied by changing the solvent and anion. Qualitatively, the extent of ion pairing of the 1,2-bis[N-(N′-butylimidazolium)]ethane salts with different anions was determined in acetone-d6 by 1H NMR spectroscopy. Thermal properties of the imidazolium salts were related not only to the nature of anions but also to the spacer length between imidazolium cations. Exceptionally high melting points of 1,2-bis[N-(N′-alkylimidazolium)]ethane bis(hexafluorophosphate)s can be explained by multiple hydrogen bonds observed in the X-ray crystal structures. Moreover, a trans conformation of the ethylene spacer of 1,2-bis[N-(N′-alkylimidazolium)]ethane bis(hexafluorophosphate)s allows good stacking structure in the crystals.
Co-reporter:Harry W. Gibson;Nori Yamaguchi;Zhenbin Niu;Jason W. Jones;Carla Slebodnick;Arnold L. Rheingold;Lev N. Zakharov
Journal of Polymer Science Part A: Polymer Chemistry 2010 Volume 48( Issue 4) pp:975-985
Publication Date(Web):
DOI:10.1002/pola.23861
Abstract
MALDI-TOF MS of the heteroditopic compound 2-(benzylammoniomethyl)dibenzo-24-crown-8 hexafluorophosphate (4) revealed oligomeric “daisy chain” species up to the hexamer. Similar results were obtained for 2-(6′-hydroxyhexylammoniomethyl)dibenzo-24-crown-8 hexafluorophosphate (8). The complexations of two substituted dibenzylammonium salts, 2,2′-dimethyldibenzylammonium hexaflurophosphate (9a) and 2,2′,5-trimethoxydibenzylammonium hexafluorophosphate (9b), with dibenzo-24-crown-8 were examined as models for slippage systems; association constants are reported for these systems. A crystal structure is reported for the new dimethylbenzylammonoium pseudorotaxane. The trimethoxy analog is shown to be capable of slippage formation of a rotaxane, albeit in low yield. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 975–985, 2010
Co-reporter:Harry W. Gibson;Zhongxin Ge;Jason W. Jones;Kim Harich;Adam Pederson ;Harry C. Dorn
Journal of Polymer Science Part A: Polymer Chemistry 2010 Volume 48( Issue 3) pp:
Publication Date(Web):
DOI:10.1002/pola.23856
No abstract is available for this article.
Co-reporter:Harry W. Gibson;Kimberly K. Brumfield;Roger A. Grisle;Christine K. F. Hermann
Journal of Polymer Science Part A: Polymer Chemistry 2010 Volume 48( Issue 17) pp:3856-3867
Publication Date(Web):
DOI:10.1002/pola.24172
Abstract
The chemistry of Reissert compounds has been used to synthesize activated difluorotetraketone monomers containing two coupled isoquinolyl moieties, linked at either the 1,1′- or 4,4′-positions. These monomers offer routes to novel families of poly(heteroarylene ether)s. New 4,4′-coupled bis(Reissert compound) 9 containing 4,4′-diketo moieties failed to afford the desired difluorotetraketo monomer upon attempted rearrangement. However, analogous bis(Reissert compound) 19 containing 4,4′-dibenzyl units did so, via aldehyde condensation, hydrolysis of the intermediate ester and oxidation of the four benzylic moieties to keto groups; thus the novel difluorotetraketone monomer 10 was prepared. Novel bis(Reissert compound)s 24, 28, and 35 were synthesized from diacid chlorides and 4-(p-fluorobenzyl)isoquinoline. Rearrangement of 24 to the diketone 29, followed by oxidation of the 4-benzyl moieties resulted in difluorotetraketone monomer 30 containing a 1,1′-linked bisisoquinoline. The 1,1′-linked bis(isoquinolylfluorodiketo) monomer 38, isomeric with 10, was prepared from 4-(p-fluorobenzyl) Reissert compound 36 by condensation with terephthaldehyde, ester hydrolysis to diol 37, and oxidation. In the course of this effort, a number of new isoquinoline Reissert compounds were synthesized as model systems. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 3856–3867, 2010
Co-reporter:Minjae Lee, Zhenbin Niu, Daniel V. Schoonover, Carla Slebodnick, Harry W. Gibson
Tetrahedron 2010 66(35) pp: 7077-7082
Publication Date(Web):
DOI:10.1016/j.tet.2010.07.010
Co-reporter:Chunying Shu, Wei Xu, Carla Slebodnick, Hunter Champion, Wujun Fu, Jonathan E. Reid, Hugo Azurmendi, Chunru Wang, Kim Harich, Harry C. Dorn and Harry W. Gibson
Organic Letters 2009 Volume 11(Issue 8) pp:1753-1756
Publication Date(Web):March 23, 2009
DOI:10.1021/ol9002527
Two new 6,6-open phenyl-C81-butyric acid methyl ester metallofulleroids, M3N@C80PCBM (M = Sc, Y), were synthesized by diazoalkane addition reactions and fully characterized. The results demonstrate that the reactive sites are the same for M3N@C80 (M = Sc, Y) but dramatically different from that of C60.
Co-reporter:Chunying Shu, Frank D. Corwin, Jianfei Zhang, Zhijian Chen, Jonathan E. Reid, Minghao Sun, Wei Xu, Jae Hyun Sim, Chunru Wang, Panos P. Fatouros, Alan R. Esker, Harry W. Gibson and Harry C. Dorn
Bioconjugate Chemistry 2009 Volume 20(Issue 6) pp:1186
Publication Date(Web):May 15, 2009
DOI:10.1021/bc900051d
A new magnetic resonance imaging (MRI) contrast agent based on the trimetallic nitride templated (TNT) metallofullerene Gd3N@C80 was synthesized by a facile method in high yield. The observed longitudinal and transverse relaxivities r1 and r2 for water hydrogens in the presence of the water-soluble gadofullerene 2 Gd3N@C80(OH)∼26(CH2CH2COOM)∼16 (M = Na or H) are 207 and 282 mM−1 s−1 (per C80 cage) at 2.4 T, respectively; these values are 50 times larger than those of Gd3+ poly(aminocarboxylate) complexes, such as commercial Omniscan and Magnevist. This high 1H relaxivity for this new hydroxylated and carboxylated gadofullerene derivative provides high signal enhancement at significantly lower Gd concentration as demonstrated by in vitro and in vivo MRI studies. Dynamic light scattering data reveal a unimodal size distribution with an average hydrodynamic radius of ca. 78 nm in pure water (pH = 7), which is significantly different from other hydroxylated or carboxylated fullerene and metallofullerene derivatives reported to date. Agarose gel infusion results indicate that the gadofullerene 2 displayed diffusion properties different from those of commercial Omniscan and those of PEG5000 modified Gd3N@C80. The reactive carboxyl functionality present on this highly efficient contrast agent may also serve as a precursor for biomarker tissue-targeting purposes.
Co-reporter:Shijun Li;Feihe Huang;Carla Slebodnick;Mehdi Ashraf-khorassani
Chinese Journal of Chemistry 2009 Volume 27( Issue 9) pp:1777-1781
Publication Date(Web):
DOI:10.1002/cjoc.200990299
Abstract
The complexation between dibenzo-24-crown-8 (1) and diquat (2) was investigated in detail by NMR, MS and X-ray analysis. It was found that dibenzo-24-crown-8 and diquat formed a 1:1 complex 1·2 in acetone with Ka=2.0×102 L·mol−1, but, as shown by X-ray analysis, a crystalline 2:1 host:guest inclusion complex 12·2 was isolated, in which a single molecule of diquat is enclosed in the concave cavity provided by two dibenzo-24-crown-8 host molecules. Both results are different from the previously assumed stoichiometry of the complexation between dibenzo-24-crown-8 and diquat. This result enriches the range of host-guest complexes based on dibenzo-24- crown-8 and provides new opportunities for developing more complicated structures and chemosensors for diquat.
Co-reporter:Minjae Lee, Daniel V. Schoonover, Anthony P. Gies, David M. Hercules and Harry W. Gibson
Macromolecules 2009 Volume 42(Issue 17) pp:6483-6494
Publication Date(Web):August 11, 2009
DOI:10.1021/ma901248m
New polymers that incorporate paraquat or crown ether moieties as chain ends or central units were synthesized by stable free radical polymerization (SFRP) and atom transfer radical polymerization (ATRP). For SFRP, TEMPO derivatives that contain a functional unit were used as initiators. For ATRP, a copper−amine or copper−bipyridine complex was used as a catalyst with radical initiators from halide derivatives of paraquat and crown ethers. These end- or center-functionalized polymers formed pseudorotaxane complexes with complementary small molecules. They also formed reversible pseudorotaxane polymers: chain-extended, star-shaped homopolymers, and block copolymers. Complexation studies to determine stoichiometry and association constants with thermodynamic parameters were performed by NMR spectroscopy and isothermal microcalorimetric (ITC) titration. The association constants of crown ether derivatives and paraquat compounds in chloroform were measured for the first time: Ka = 2.97 × 103 M−1 for paraquat polystyrene and bis(m-phenylene)-32-crown-10 and Ka = 4.38 × 103 M−1 for paraquat polystyrene and crown ether polystyrene. These are 4−6-fold higher than the association constant of the analogous small molecule host−guest system in acetone.
Co-reporter:Harry W. Gibson;Aurica Farcas;Jason W. Jones;Zhongxin Ge;Feihe Huang;Matthew Vergne;David M. Hercules
Journal of Polymer Science Part A: Polymer Chemistry 2009 Volume 47( Issue 14) pp:3518-3543
Publication Date(Web):
DOI:10.1002/pola.23435
Abstract
Dibenzo-24-crown-8-terminated polystyrene (5) was chain extended to “dimeric” 8 by pseudorotaxane formation with a ditopic guest, α,ω-bis[p-(N-benzylammoniomethyl)phenoxy]heptane bis(hexafluorophosphate) (7). The three-armed star polymer 11 was similarly formed by complexation of the dibenzo-24-crown-8-terminated polystyrene (5) with a tritopic secondary ammonium salt, 1,3,5-tris[p-(benzylammoniomethyl)phenyl]benzene tris(hexafluorophosphate) (10). Another three-armed star polymer 13 was self-assembled from dibenzo-24-crown-8-terminated polystyrene (5) and a tetratopic paraquat compound, 1,2,4,5-tetrakis{p-N-[(N′-methyl-4,4′-bipyridinium)methylphenyl]}benzene octakis(hexafluorophosphate) (12). The above chain extension and star polymer formation processes seemed to be cooperative; that is, the second and third complexation steps proceed with stepwise higher efficiencies than statistically expected. Dibenzo-24-crown-8-terminated polystyrene (5) was chain extended with secondary ammonium terminated polystyrene 14, forming 16, and also self-assembled with a secondary ammonium ion terminated polyisoprene 15 to form supramolecular block copolymer 17. These processes were examined by NMR, mass spectrometry and viscometery. Thus, although binding in these systems is not particularly strong (association constants <104 M−1), these examples provide proof-of-principle that pseudorotaxane formation is a viable concept for chain extension and self-assembly of novel types of block copolymers and star polymers. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 3518–3543, 2009
Co-reporter:Harry W. Gibson;Zhongxin Ge;Jason W. Jones;Kim Harich;Adam Pederson;Harry C. Dorn
Journal of Polymer Science Part A: Polymer Chemistry 2009 Volume 47( Issue 23) pp:6472-6495
Publication Date(Web):
DOI:10.1002/pola.23688
Abstract
Multitopic dibenzylammonium derivatives (4) of C60 were prepared by Bingel reactions of C60 with a malonate diester (2) containing two t-BOC protected dibenzylamine moieties, followed by deprotection and protonation. Self-assembly of model pseudorotaxanes 5 from the multidibenzylammonium C60 derivatives with dibenzo-24-crown-8 was studied by 1H NMR spectroscopy and mass spectrometry. Self-assembly of linear and star-shaped pseudorotaxanes 8 with up to 12 arms based on polystyrenes bearing terminal DB24C8 host units (7) and the guest functionalized C60 salts was demonstrated by 1H NMR spectroscopy and solution phase viscometry. These studies provide further evidence of the potential of supramacromolecular chemistry in construction of complex polymeric architectures. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 6472–6495, 2009
Co-reporter:Clairette D. Angeli, Ting Cai, James C. Duchamp, Jonathan E. Reid, Eric S. Singer, Harry W. Gibson and Harry C. Dorn
Chemistry of Materials 2008 Volume 20(Issue 15) pp:4993
Publication Date(Web):June 28, 2008
DOI:10.1021/cm800795q
The availability of pure samples is a major hurdle in the study of trimetallic nitride templated endohedral metallofullerenes (TNT EMFs). Current HPLC methods of purification are costly, nonrecyclable, and take considerable time. Reported herein is a solvent-free reaction of crude soot extract (empty-cage and Sc or Lu TNT EMFs) in molten 9-methylanthracene that affords almost complete conversion of empty-cage fullerenes to 9-methylanthracene adducts, while leaving TNT EMFs unreacted. After the recovered extract is washed with diethyl ether, the washed extract can then be applied to a silica gel column and subsequently flushed with toluene. To illustrate for Sc3N@C80, the sample was purified to ∼60% in a time period of <24 h. Final one-step HPLC purification provided high-purity TNT EMF samples (>99%) in an expeditious and less costly manner.
Co-reporter:Harry W. Gibson;Hong Wang;Chin P. Chng;Thomas E. Glass;Daniel Schoonover;Lev N. Zakharov;Arnold L. Rheingold
Heteroatom Chemistry 2008 Volume 19( Issue 1) pp:48-54
Publication Date(Web):
DOI:10.1002/hc.20393
Abstract
A small cyclophane, bis(5-carbometh-oxy-1,3-phenylene)-14-crown-4 (BCMP14C4, 3) and its diacid, bis(5-carboxy-1,3-phenylene)-14-crown-4 (4), were synthesized and characterized. The solid-state molecular structures of 3 and 4 were determined by X-ray crystallography as ladder or stepped conformations in which the two aromatic rings are antiparallel to each other without overlap and the ethylene tethers both take trans-conformations. Diester 3 is formed in the lowest cyclization yield (under the same reaction conditions) and exhibits the highest melting point compared to its larger ring (20-, 26- and 32-membered) analogs. In CD2Cl2 solution, diester 3 exists predominantly as a nonplanar gauche–gauche structure as deduced by H NMR studies. © 2008 Wiley Periodicals, Inc. Heteroatom Chem 19:48–54, 2008; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.20393
Co-reporter:Ting Cai, Liaosa Xu, Chunying Shu, Jonathan E. Reid, Harry W. Gibson and Harry C. Dorn
The Journal of Physical Chemistry C 2008 Volume 112(Issue 49) pp:19203-19208
Publication Date(Web):November 14, 2008
DOI:10.1021/jp804791e
A non-IPR (isolated pentagon rule) trimetallic nitride endohedral metallofullerene derivative, a diethyl malonate adduct of Sc3N@C68, was synthesized, isolated, and characterized by mass spectrometry and NMR spectroscopy. The proposed addition site of Sc3N@C68 monoadduct was in good agreement with the experimental and calculated 13C NMR spectra, LUMO electron density study, and UV−vis spectroscopy.
Co-reporter:Feihe Huang, Carla Slebodnick, Karen A. Switek and Harry W. Gibson
Chemical Communications 2006 (Issue 18) pp:1929-1931
Publication Date(Web):21 Mar 2006
DOI:10.1039/B600227G
Bis(meta-phenylene)-32-crown-10-based cryptands have been proved to complex diquat much more strongly than bis(meta-phenylene)-32-crown-10 itself; in fact, one containing a pyridyl moiety has one of the highest Ka values yet reported.
Co-reporter:Feihe Huang, Harry W. Gibson
Progress in Polymer Science 2005 Volume 30(Issue 10) pp:982-1018
Publication Date(Web):October 2005
DOI:10.1016/j.progpolymsci.2005.07.003
The latest progress in the field of polypseudorotaxanes and polyrotaxanes is reviewed in four parts: (1) main chain systems, (2) side chain systems, (3) other systems with related structures and (4) properties, potential applications and future directions.
Co-reporter:Feihe Huang, Karen A. Switek and Harry W. Gibson
Chemical Communications 2005 (Issue 29) pp:3655-3657
Publication Date(Web):24 Jun 2005
DOI:10.1039/B504250J
Strong complexation between a pyridine-containing cryptand and paraquat can be reversibly switched off (and back on) by adding acid (and then base).
Co-reporter:Feihe Huang and Harry W. Gibson
Chemical Communications 2005 (Issue 13) pp:1696-1698
Publication Date(Web):03 Feb 2005
DOI:10.1039/B417966H
A supramolecular poly[3]pseudorotaxane was prepared by self-assembly of a homoditopic cylindrical bis(crown ether) host and a bisparaquat derivative in solution by host–guest complexation.
Co-reporter:Feihe Huang, Ilia A. Guzei, Jason W. Jones and Harry W. Gibson
Chemical Communications 2005 (Issue 13) pp:1693-1695
Publication Date(Web):03 Feb 2005
DOI:10.1039/B417661H
Significant improvement of complexation of a bisparaquat guest was achieved by the formation of a pseudocryptand-based [3]pseudorotaxane.
Co-reporter:Feihe Huang, Matthew Lam, Eric J. Mahan, Arnold L. Rheingold and Harry W. Gibson
Chemical Communications 2005 (Issue 26) pp:3268-3270
Publication Date(Web):17 Jun 2005
DOI:10.1039/B503092G
Host folding for the formation of taco complexes can be promoted by introduction of additional interactions between the host and guest as shown by enhanced associations and X-ray crystal structures.
Co-reporter:Harry W. Gibson, Hong Wang, Klaus Bonrad, Jason W. Jones, Carla Slebodnick, Lev N. Zackharov, Arnold L. Rheingold, Bradley Habenicht, Peter Lobue and Amy E. Ratliff
Organic & Biomolecular Chemistry 2005 vol. 3(Issue 11) pp:2114-2121
Publication Date(Web):04 May 2005
DOI:10.1039/B503072M
Two isomers of bis(carbomethoxybenzo)-24-crown-8 (cis-BCMB24C8, 1, and trans-BCMB24C8, 2) were synthesized regiospecifically with acceptable to excellent yields. Cyclization in the presence of a template reagent, KPF6, led to an essentially quantitative yield of the potassium complex of the crown ether 1; the isolated cyclization yield of pure 1 was a remarkable 89%! The methods not only avoid the very difficult separation of the isomers, but also greatly shorten the synthesis time by eliminating syringe pump usage during cyclization. The complexations of the isomeric BCMB24C8 with dibenzylammonium hexafluorophosphate (10) were studied by NMR; association constants (Ka) for 1 and 2 with the dibenzylammonium cation are 190 and 312 M−1, respectively. The X-ray crystal structures of crown ether 2 and the complexes 1·KPF6, 2·KPF6 and pseudorotaxane 2·10 were determined.
Co-reporter:Feihe Huang, Liang Zhou, Jason W. Jones, Harry W. Gibson and Mehdi Ashraf-Khorassani
Chemical Communications 2004 (Issue 23) pp:2670-2671
Publication Date(Web):14 Oct 2004
DOI:10.1039/B411234B
Dimers of inclusion complexes were formed from a new cryptand and viologens (paraquats) driven by dipole–dipole and face-to-face π-stacking interactions as shown by mass spectrometric characterization and X-ray analysis.
Co-reporter:Feihe Huang, Frank R. Fronczek and Harry W. Gibson
Chemical Communications 2003 (Issue 13) pp:1480-1481
Publication Date(Web):20 May 2003
DOI:10.1039/B302682E
The first supramolecular poly(taco complex) was formed in the solid state as shown by X-ray analysis.
Co-reporter:Feihe Huang, Lev N. Zakharov, Arnold L. Rheingold, Jason W. Jones and Harry W. Gibson
Chemical Communications 2003 (Issue 17) pp:2122-2123
Publication Date(Web):24 Jul 2003
DOI:10.1039/B304995G
Water acts as a “molecular clip” to form a supramolecular cryptand structure that improves complexation of a diammonium salt by pseudorotaxane formation, and leads to a novel dimer in the solid state.
Co-reporter:Harry W. Gibson;William S. Bryant;Sang-Hun Lee
Journal of Polymer Science Part A: Polymer Chemistry 2001 Volume 39(Issue 12) pp:1978-1993
Publication Date(Web):26 APR 2001
DOI:10.1002/pola.1174
Poly[(methyl acrylate)-rotaxa-(30-crown-10)] (5) and poly[(methyl methacrylate)-rotaxa-(30-crown-10)] (6) were synthesized by azobisisobutyronitrile-initiated free-radical bulk polymerizations of the respective monomers in the presence of 30-crown-10 (1; equimolar; 5 times the monomer mass). For 5, 3.8 mass % (0.81 mol % with respect to the monomer) of the crown was incorporated versus 1.7 mass % (0.39 mol % with respect to the monomer) for 6. Control reactions with 18-crown-6, which is to small to be threaded, showed that chain transfer to the crown ethers was detectable only for the acrylate but was relatively negligible and spectroscopically distinct. The threading yields were much higher with these systems than with polystyrene, most likely because of the greater compatibility of the crown ether with these polar monomers and polymers and the consequent ability to carry out the polymerizations homogeneously in the absence of added solvent; however, the threading process was still essentially statistical. Therefore, the polymerization of methacrylate monomers 8a–8c based on tetraarylmethane moieties connected via diethyleneoxy or triethyleneoxy spacers was examined in the presence of 1 in the belief that the supramolecular semirotaxane monomer 9 formed statistically insitu could be captured more efficiently and produce higher threading yields, presumably of side-chain polyrotaxanes, than the simple (meth)acrylate monomers. Azobisisobutyronitrile-initiated polymerizations either neat or in toluene produced polyrotaxanes 10 with up to about 1.6 mass % and 2 mol % threaded crown ether, presumably trapped on the pendant stoppered side chains. Although primarily statistical in nature, the latter rotaxane syntheses afforded on a molar basis 3–7 times more efficient incorporation of 1 than styrene (0.33 mol %), methyl acrylate (0.81 mol %), or methyl methacrylate (0.39 mol %) monomers for the preparation of main-chain polypseudorotaxanes and indeed even surpassed the 60-crown-20/polyacrylonitrile system (1.5 mol %). This was presumed to be due to the fact that the loss of the crown ether, once it was threaded onto the monomer to form 9 and the latter was polymerized, was either retarded (by the tetraphenylmethyl stopper in 10a) or prevented completely [by tris(p-t-butylphenyl)phenylmethyl stoppers in 10b and 10c]. © 2001 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 39: 1978–1993, 2001
Co-reporter:Nori Yamaguchi;Lesley Hamilton
Polymers for Advanced Technologies 2000 Volume 11(Issue 8‐12) pp:791-797
Publication Date(Web):7 NOV 2000
DOI:10.1002/1099-1581(200008/12)11:8/12<791::AID-PAT29>3.0.CO;2-H
The formation of pseudorotaxanes from dibenzo-24-crown-8 (DB24C8) and secondary aliphatic ammonium ions was reported by Stoddart et al. Based on that molecular recognition motif several systems have been examined as prototypical examples of (1) self-assembly of dendrimers via pseudorotaxane formation, (2) self-assembly of linear non-covalent polymers of the pseudo- rotaxane type and (3) control of properties of a polymer by pseudorotaxane formation. Attachment of a DB24C8 moiety to the “focal point” of first, second and third generation benzyl ether dendrons (Fréchet type) allowed solution phase self-assembly with a core unit consisting of 1,3,5-tris[p-(N-benzylammoniomethyl)phenyl]benzene to produce the corresponding dendritic pseudoro-rotaxane structures, which are of nanometer scale. Ditopic hosts were prepared by coupling DB24C8 units with difunctional linear species; ditopic guests were similarly constructed by linking two dibenzylammonium ion moieties. At high concentrations in relatively non-polar solvents these complementary building blocks self-assembled into non-covalently bonded (pseudorotaxane) linear arrays, with high viscosity and fiber forming ability. Treatment of polymethacrylates bearing pendant DB24C8 units with dibenzylammonium resulted in changes in properties as a result of formation of side-chain pseudorotaxane units. Copyright © 2000 John Wiley & Sons, Ltd.
Co-reporter:Nori Yamaguchi
Angewandte Chemie 1999 Volume 111(Issue 1‐2) pp:
Publication Date(Web):12 MAR 1999
DOI:10.1002/(SICI)1521-3757(19990115)111:1/2<195::AID-ANGE195>3.0.CO;2-B
Wie aus einem molekularen Baukasten reversibel zusammengesteckt erscheint die Titelverbindung, ein kettenförmiges, supramolekulares Pseudorotaxanpolymer, das in Lösung selbstorganisiert aus zwei komplementären homoditopen Molekülen mit sekundären Ammonio- und Dibenzo[24]krone-8-Gruppen entsteht.
Co-reporter:Nori Yamaguchi
Angewandte Chemie International Edition 1999 Volume 38(Issue 1‐2) pp:
Publication Date(Web):18 JAN 1999
DOI:10.1002/(SICI)1521-3773(19990115)38:1/2<143::AID-ANIE143>3.0.CO;2-Y
As if with a molecular building kit, a linear supramolecular pseudorotaxane polymer was formed with reversible chain extension in solution by self-organization of two complimentary homoditopic molecules with secondary ammonium and dibenzo[24]crown-8 moieties.
Co-reporter:Nori Yamaguchi;Lesley M. Hamilton
Angewandte Chemie International Edition 1998 Volume 37(Issue 23) pp:
Publication Date(Web):23 DEC 1998
DOI:10.1002/(SICI)1521-3773(19981217)37:23<3275::AID-ANIE3275>3.0.CO;2-9
Self-organization is the key. A series of dendritic pseudorotaxanes were efficiently constructed from complementary building blocks—namely, a three-armed, triply charged ammonium salt and the first, second, and third generations of benzyl ether dendrons bearing the dibenzo[24]crown-8 moiety. The pseudorotaxane arising from the third-generation dendron is shown in the picture.
Co-reporter:Nori Yamaguchi;Devdatt S. Nagvekar
Angewandte Chemie International Edition 1998 Volume 37(Issue 17) pp:
Publication Date(Web):17 DEC 1998
DOI:10.1002/(SICI)1521-3773(19980918)37:17<2361::AID-ANIE2361>3.0.CO;2-P
Like the proverbial monkey chain one heteroditopic self-complementary molecule, comprising a crown ether unit and a paraquat unit, catches a second such molecule in solution and thus by self-organization forms novel linear oligo- and polymolecular arrays (shown schematically; the crown ether unit is denoted by the ellipse, and the paraquat unit by the rectangle).
Co-reporter:Nori Yamaguchi;Lesley M. Hamilton
Angewandte Chemie 1998 Volume 110(Issue 23) pp:
Publication Date(Web):12 MAR 1999
DOI:10.1002/(SICI)1521-3757(19981204)110:23<3463::AID-ANGE3463>3.0.CO;2-9
Selbstorganisation ist der Schlüssel zu einer Reihe dendritischer Pseudorotaxane, die sich in guten Ausbeuten aus komplementären Bausteinen aufbauen lassen: aus einem dreiarmigen, dreifach geladenen Ammoniumsalz und aus Benzylether-Dendronen der ersten bis dritten Generation, die die Dibenzo[24]krone-8-Einheit enthalten. Das Pseudorotaxan mit dem Dendron der dritten Generation ist im Bild schematisch dargestellt.
Co-reporter:Nori Yamaguchi;Devdatt S. Nagvekar
Angewandte Chemie 1998 Volume 110(Issue 17) pp:
Publication Date(Web):12 MAR 1999
DOI:10.1002/(SICI)1521-3757(19980904)110:17<2518::AID-ANGE2518>3.0.CO;2-1
Wie Schlangen, die sich gegenseitig in den Schwanz beißen: Eine heteroditope, selbstkomplementäre Verbindung, bestehend aus einer Kronenethereinheit und einer Paraquateinheit, angelt sich in Lösung ihresgleichen und bildet so unter Selbstorganisation neuartige, oligo- und polymolekulare Aggregate (siehe Schema; die Kronenethereinheit ist durch die Ellipse und die Paraquateinheit durch das Rechteck symbolisiert).
Co-reporter:Zhenbin Niu and Harry W. Gibson
Organic & Biomolecular Chemistry 2011 - vol. 9(Issue 20) pp:NaN6912-6912
Publication Date(Web):2011/08/11
DOI:10.1039/C1OB06299A
Via the self-assembly of two bis(meta-phenylene)-32-crown-10-based cryptands, bearing covalent and metal complex (ferrocene) linkages, with dimethyl paraquat, novel [3]pseudorotaxanes were formed statistically and anticooperatively, respectively.
Co-reporter:Minjae Lee, U Hyeok Choi, Sungsool Wi, Carla Slebodnick, Ralph H. Colby and Harry W. Gibson
Journal of Materials Chemistry A 2011 - vol. 21(Issue 33) pp:NaN12287-12287
Publication Date(Web):2011/07/13
DOI:10.1039/C1JM10995B
A new class of organic ionic plastic crystals was discovered. A series of alkylene 1,2-bis[N-(N′-alkylimidazolium)] salts with Br− and PF6− anions was prepared and most of the ethylene (C2) linked compounds undergo multiple solid–solid phase transitions as determined by differential scanning calorimetry (DSC). The salts with longer spacers (C3 or C4) between the imidazolium units do not display any solid–solid transitions. The PF6− salts with terminal C10 and C12n-alkyl moieties are “classical” organic ionic plastic crystals by Timmermans' definition; they have low ΔSf (11 J K−1 mol−1 for C10 and 12 J K−1 mol−1 for C12). Additionally, the C2 linked bis-imidazolium salts with n-butyl, n-hexyl, n-heptyl and n-octyl termini all undergo two, three or four solid state transitions and exhibit ΔSf values in the range of 36–49 J K−1 mol−1, which are similar to those of other known ionic plastic crystalline materials. These materials were additionally characterized via ionic conductivity and solid state NMR. These ionic plastic crystals are presumably single-ion conductors, but the ionic conductivities appear to be too low for practical applications. The activation energy for conduction decreases as these compounds are heated through each solid–solid transition. The lack of any change in solid state 2H NMR spectra with temperature indicates that there is no change in phenyl ring flipping, suggesting no change in the imidazolium local environment. A resolution of this apparent dichotomy is perhaps that the counterions reside with the ethylene spacers between imidazolium moieties.