Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 12 de 12
Filter
Add more filters










Publication year range
1.
Chemistry ; 15(37): 9596-605, 2009 Sep 21.
Article in English | MEDLINE | ID: mdl-19579236

ABSTRACT

The recently developed concept of DNA-based asymmetric catalysis involves the transfer of chirality from the DNA double helix in reactions using a noncovalently bound catalyst. To date, two generations of DNA-based catalysts have been reported that differ in the design of the ligand for the metal. Herein we present a study of the first generation of DNA-based catalysts, which contain ligands comprising a metal-binding domain linked through a spacer to a 9-aminoacridine moiety. Particular emphasis has been placed on determining the effect of DNA on the structure of the Cu(II) complex and the catalyzed Diels-Alder reaction. The most important findings are that the role of DNA is limited to being a chiral scaffold; no rate acceleration was observed in the presence of DNA. Furthermore, the optimal DNA sequence for obtaining high enantioselectivities proved to contain alternating GC nucleotides. Finally, DNA has been shown to interact with the Cu(II) complex to give a chiral structure. Comparison with the second generation of DNA-based catalysts, which bear bipyridine-type ligands, revealed marked differences, which are believed to be related to the DNA microenvironment in which the catalyst resides and where the reaction takes place.


Subject(s)
DNA, Catalytic/chemistry , Ligands , Catalysis , Circular Dichroism , Copper/chemistry , Crystallography, X-Ray , Kinetics , Molecular Conformation , Stereoisomerism
2.
J Am Chem Soc ; 130(35): 11783-90, 2008 Sep 03.
Article in English | MEDLINE | ID: mdl-18681429

ABSTRACT

This study shows that the role of DNA in the DNA-based enantioselective Diels-Alder reaction of azachalcone with cyclopentadiene is not limited to that of a chiral scaffold. DNA in combination with the copper complex of 4,4'-dimethyl-2,2'-bipyridine (Cu-L1) gives rise to a rate acceleration of up to 2 orders of magnitude compared to Cu-L1 catalysis alone. Furthermore, both the enantioselectivity and the rate enhancement prove to be dependent on the DNA-sequence. These features are the main reasons for the efficient and enantioselective catalysis observed with salmon testes DNA/Cu-L1 in the Diels-Alder reaction. The fact that absolute levels of stereocontrol can be achieved with a simple and weak DNA-binding complex like Cu-L1 is a clear demonstration of the power of the supramolecular approach to hybrid catalysis.


Subject(s)
Chalcones/chemistry , Cyclopentanes/chemistry , DNA/chemistry , 2,2'-Dipyridyl/chemistry , Animals , Catalysis , Circular Dichroism , Copper/chemistry , DNA, Single-Stranded/chemistry , Kinetics , Male , Nucleic Acid Conformation , Oligonucleotides/chemistry , Organometallic Compounds/chemistry , Salmon , Stereoisomerism , Substrate Specificity
3.
J Phys Chem B ; 111(19): 5204-11, 2007 May 17.
Article in English | MEDLINE | ID: mdl-17439278

ABSTRACT

The pH-dependent phase behavior and hydroxide-ion adsorption ability of a series of (reduced) carbohydrate-based gemini surfactants were studied between pH 2 and 12. Static and dynamic light scattering were employed to address transitions in the aggregate morphologies and cryo-electron microscopy was used to provide further evidence for the morphologies present in solution. Changes in aggregate structure as a result of a change in solution pH and an accompanying change in protonation state or a change in molecular structure can be rationalized in terms of the variations in the packing parameter. In this paper we have focused our attention on the size of the carbohydrate moiety, the carbohydrate stereochemistry and the nature of the spacer (hydrophobic vs hydrophilic). At near neutral pH, most of the gemini surfactants form vesicles. Upon lowering of the pH, the vesicles undergo a transition toward wormlike micelles followed by a transition to spherical micelles. Upon increasing the solution pH, flocculation occurs due to charge neutralization followed at still higher pH by redispersion and charge reversal of the vesicles through the specific adsorption of hydroxide ions to the vesicular surface. Upon decreasing head group size at constant, but low, degrees of protonation, the packing parameter has a tendency to become larger than one resulting in the formation of inverted phases. Upon further decrease in the head group size, oil droplets are observed. In case of a hydrophobic spacer, the carbohydrate stereochemistry affects the pH of the transitions, but not the type of the transitions. By contrast, for a hydrophilic spacer, the pH of the transitions remains unaffected. Adsorption of hydroxide ions at basic pH follows similar trends, but was only found for vesicles and oil droplets. The large range of structural variations that we have examined allows a better understanding of the requirements for the phase transitions for carbohydrate-based gemini surfactants as well as for the physisorption of hydroxide ions to interfaces in general.

4.
J Phys Chem B ; 110(43): 21694-700, 2006 Nov 02.
Article in English | MEDLINE | ID: mdl-17064128

ABSTRACT

The phase behavior of a series of carbohydrate-based gemini surfactants with varying spacer lengths was studied using static and dynamic light scattering between pH 2 and 12. Cryo-electron microscopy pictures provide evidence for the different morphologies present in solution. The spacer length of the gemini surfactants was varied from two to 12 methylene units. At near neutral pH, spherical vesicles were obtained for gemini surfactants with a spacer shorter than 10 methylene units, whereas nonspherical vesicles were obtained for spacer lengths of 10 and 12. Upon decreasing the pH, the vesicles underwent transitions toward worm-like micelles and spherical micelles for a spacer length of six and larger, whereas for shorter spacers, these transitions are not observed. For the shortest spacer at low pH, perforated vesicles are observed, and vesicles built from the gemini surfactant with a spacer of four methylene units only underwent a transition toward worm-like micelles. Upon increasing the pH to slightly basic values, flocculation followed by redispersion upon charge reversal was observed up to a spacer length of eight methylene units. The redispersal is explained by hydroxide-ion binding to the uncharged vesicular surface. By contrast, vesicles formed from the gemini surfactants with 10 and 12 methylene units only undergo a transition toward inverted phases. The observations can be understood in terms of the packing parameter.

5.
Org Biomol Chem ; 4(19): 3569-70, 2006 Oct 07.
Article in English | MEDLINE | ID: mdl-16990930

ABSTRACT

The mechanism of hydroxide ion binding to nonionic surfaces is explored by variation of the properties of the water-aggregate interface and by variation of the type of the aggregate.

6.
Langmuir ; 22(6): 2558-68, 2006 Mar 14.
Article in English | MEDLINE | ID: mdl-16519455

ABSTRACT

Sugar-based gemini surfactants (GSs) display rich pH-dependent phase diagrams and are considered to be promising candidates as gene- and drug-delivery vehicles for biomedical applications. Several sugar-based GSs form vesicles around neutral pH. The vesicular dispersions undergo transitions toward wormlike micelles and spherical micelles at acidic pH, whereas flocculation followed by redispersion upon charge reversal is observed at basic pH. The influence of various amounts of the double-tailed phospholipids DOPC (1,2-dioleoyl-sn-glycero-3-phosphocholine) and DOPE (1,2-dioleoyl-sn-glycero-3-phosphoethanolamine) and of the single-tailed surfactants lyso-PC (1-palmitoyl-2-hydroxy-sn-glycero-3-phosphocholine) and OTAC (octadecyltrimethylammonium chloride) on the phase behavior of GS1 (1,8-bis(N-octadec-9-yl-1-deoxy-D-glucitol-1-ylamino)3,6-dioxaoctane) was determined as a function of pH, in water and in water at physiological ionic strength. The pH corresponding to the phase transitions and the characteristics of the aggregates were determined by means of a combination of physical techniques: static and dynamic light scattering (SLS and DLS), fluorescence spectroscopy, cryo-TEM and diffusion- and (31)P NMR. The results show that the additives affect the phase behavior of the GS1 dispersions in a pH-dependent fashion. In the presence of double-tailed phospholipids, a higher degree of protonation of GS1 must be reached to observe micelle formation, whereas single-tailed surfactants affect these transitions only slightly. In the presence of increasing amounts of lyso-PC, the pH range of flocculation becomes more narrow, indicating the increased hydration of the vesicles. The pH of redispersion after charge reversal is particularly sensitive to the presence of positively charged additives. It is suggested that the cationic headgroups disturb the hydrogen-bond structure of water at the vesicular surface, hampering OH(-) binding. The effect of an increase in ionic strength to physiological values is found to be modest, except for the dispersions containing the positively charged additives.

7.
Langmuir ; 21(22): 9809-17, 2005 Oct 25.
Article in English | MEDLINE | ID: mdl-16229496

ABSTRACT

The kinetics of the S(N)2 reaction of a series of aromatic alkylsulfonates with water and bromide ions in membrane mimetic media have been investigated. These media include vesicles formed from only synthetic amphiphiles, vesicles composed only of phospholipids and mixtures of these components. Special focus is placed on the influence of the addition of n-dodecyl-beta-glucoside as a mimic for glycolipids. The kinetic data have been analyzed by using the pseudophase model for bimolecular reactions. Contrary to previous results on a base-catalyzed E2 reaction (Org. Biomol. Chem. 2004, 2, 1789-1799), the presence of n-dodecyl-beta-glucoside at the vesicular surface does not lead to large rate accelerations for the S(N)2 reaction. In fact, when present at 50 mol % (i.e., the additive covers 34% of the vesicular surface) these glycolipid mimics appear not to affect the bimolecular rate constants, but they only decrease the local water concentration by about 40%. The reactivity of water at the surface of vesicles that are formed from cationic amphiphiles appears to be increased about 10-fold relative to the reactivity of water in the bulk liquid, whereas in zwitterionic vesicles the reactivity is comparable to that in bulk water. The obtained rate constants are also compared to micellar rate constants.


Subject(s)
Glycolipids/chemistry , Bromides/chemistry , Catalysis , Glucosides/chemistry , Ions , Kinetics , Lipid Bilayers , Magnetic Resonance Spectroscopy , Micelles , Models, Chemical , Models, Statistical , Phospholipids/chemistry , Substrate Specificity , Temperature , Water/chemistry
8.
Nature ; 435(7043): 746-7, 2005 Jun 09.
Article in English | MEDLINE | ID: mdl-15944683
9.
Org Biomol Chem ; 2(12): 1789-99, 2004 Jun 21.
Article in English | MEDLINE | ID: mdl-15188048

ABSTRACT

Vesicles formed from synthetic, double-tailed amphiphiles are often used as mimics for biological membranes. However, biological membranes are a complex mixture of various compounds. In the present paper we describe a first attempt to study the importance of additives on vesicular catalysis. The rate-determining deprotonation of 5-nitrobenzisoxazole (Kemp elimination) by hydroxide ion is efficiently catalysed by vesicles formed from dimethyldi-n-octadecylammonium chloride (C(18)C(18)(+)) as a result of (partial) dehydration of the reactants (especially the hydroxide ion) at the vesicular binding sites. Gradual addition of linear alcohols, such as n-decanol (C(10)OH), n-octadecanol (C(18)OH) and batyl alcohol (C(18)GlyOH) leads to a decrease in the observed catalysis. By contrast, gradual addition of oleyl alcohol, n-dodecyl-beta-glucoside (C(12)Glu) and n-dodecyl-beta-maltoside (C(12)Mal) leads to an increase in the observed catalysis. A detailed kinetic analysis, taking into account substrate binding site polarities, counterion binding percentages and binding affinity of the kinetic probe, suggests that the catalytic changes depend strongly on subtle changes in the structure of the additive. Whereas the C(12)Glu-induced effect can be explained by an increase in the vesicular rate constant, the effect of C(12)Mal can only be explained by an increase in the binding constant of the kinetic probe. However, for these pyranoside-containing vesicles others factors, such as a more extensive dehydration of the hydroxide ion, and micelle formation have to be considered. For the linear alcohols, besides a decrease in the counterion binding, changes in the vesicular rate constant and the binding constant should be taken into account. These two parameters change to a different extent for the different alcohols. The kinetic analysis is supported by differential scanning calorimetry (DSC), E(T)(30) absorbance data and Nile Red, Laurdan, ANS and pyrene fluorescence measurements. The overall kinetic results are illustrative for the highly complex mix of factors which determines catalytic effects on reactions occurring in biological cell membranes.


Subject(s)
Alcohols/chemistry , Biomimetic Materials/chemistry , Membranes, Artificial , Pyrans/chemistry , Quaternary Ammonium Compounds/chemistry , Calorimetry, Differential Scanning , Catalysis , Cations/chemistry , Isoxazoles/chemistry , Kinetics , Molecular Structure , Nitrobenzenes/chemistry , Phase Transition , Phosphatidylcholines/chemistry , Sodium Hydroxide/chemistry , Spectrometry, Fluorescence
10.
J Org Chem ; 69(3): 660-4, 2004 Feb 06.
Article in English | MEDLINE | ID: mdl-14750789

ABSTRACT

The (13)C NMR spectra of two different iodoalkynes, 1-iodo-1-hexyne (1) and diiodoethyne (2), exhibit a strong solvent dependence. Comparisons of the data with several common empirical models, including Gutmann's Donor numbers, Reichardt's E(N)(T), and Taft and Kamlet's beta and pi, demonstrate that this solvent effect arises from a specific acid-base interaction. Solvent basicity measures such as Donor numbers and beta values correlate well with the alpha-carbon chemical shift of 1, but polarity measures such as E(N)(T) and pi do not correlate. The similarity of the solvent effect for 1 and 2 suggests that carbon-carbon bond polarization may not play a role in the change in chemical shift, as previously hypothesized.

11.
J Am Chem Soc ; 125(7): 1825-33, 2003 Feb 19.
Article in English | MEDLINE | ID: mdl-12580609

ABSTRACT

The rate-determining deprotonation of 5-nitrobenzisoxazole (Kemp elimination) by hydroxide is efficiently catalyzed by vesicles formed from dimethyldioctadecylammonium chloride (C(18)()C(18)()(+)()). Gradual addition of sodium didecyl phosphate (C(10)()C(10)()(-)()) leads to the formation of catanionic vesicles, which were characterized by cryo-electron microscopy, and their main phase transition temperatures (DSC) and zeta-potentials. Increasing percentages of C(10)()C(10)()(-)() in the vesicular bilayers decrease the catalysis of the Kemp elimination. A detailed kinetic analysis, supported by consideration of substrate binding site polarities and counterion binding percentages, suggest that the catalytic effects of C(18)()C(18)()(+)()/C(10)()C(10)()(-)() catanionic vesicles are primarily determined by the binding of catalytically active hydroxide ions to the vesicular surface area. The formation of neutral microdomains between 10 and 30 mol % of C(10)()C(10)()(-)() in the bilayer, as revealed by DSC, is not apparent from the catalytic effects found for these vesicles. Interestingly, the catalytic effects observed for 50 mol % C(10)()C(10)()(-)() in the catanionic vesicles indicate an asymmetric distribution of C(18)()C(18)()(+)() and C(10)()C(10)()(-)() over the bilayer leaflets. The overall kinetic results illustrate the highly complex mix of factors which determines catalytic effects on reactions occurring in biological cell membranes.


Subject(s)
Alkanes/chemistry , Biomimetic Materials/chemistry , Membranes/chemistry , Micelles , Phosphates/chemistry , Quaternary Ammonium Compounds/chemistry , Calorimetry, Differential Scanning , Catalysis , Isoxazoles/chemistry , Kinetics , Microscopy, Electron , Nitrobenzenes/chemistry , Sodium Hydroxide/chemistry , Static Electricity , Surface Properties
12.
J Colloid Interface Sci ; 226(1): 76-82, 2000 Jun 01.
Article in English | MEDLINE | ID: mdl-11401349

ABSTRACT

Titration microcalorimetry and steady-state fluorescence spectroscopy have been used to study the aggregation of mono-endcapped hydrophobically modified poly(sodium acrylate)s in aqueous solution. Polymers with molecular weights varying between 800 and 31,700 were synthesized by radical polymerization using an initiator and chain transfer agent. The resulting polymers form hydrophobic microdomains in aqueous solutions. The following conditions were applied: no salt and pH 5 and 9, respectively; 1 M sodium citrate and pH 9. At pH 5 the critical aggregation concentration (CAC, the concentration at which microdomains are formed) increases with increasing molecular weight of the polymers. The concentration range for aggregation is about 0.2-2.4 mM. At pH 9 the carboxylic acid groups are deprotonated and electrostatic repulsions are introduced; therefore the concentration for aggregation rises to about 80 mM. Interestingly, in case of polymers having M(n)<1400 the CAC decreases with increasing molecular weight due to a counterion-concentration gradient toward the hydrophobic microdomain. Near the microdomain the counterion binding is increased, reducing the electrostatic repulsions and allowing for lower aggregation concentrations. In the presence of 1 M sodium citrate this anomalous trend is suppressed to a large extent; since the overall counterion binding is increased and the CAC is lower. The concentration for aggregation is then in the same range as at pH 5 in the absence of salt. Copyright 2000 Academic Press.

SELECTION OF CITATIONS
SEARCH DETAIL
...