<<

Perspective

Supramolecular templating in thermodynamically controlled synthesis

Ricardo L. E. Furlan*, Sijbren Otto†‡, and Jeremy K. M. Sanders†‡

*Universidad Nacional de Rosario, Facultad de Ciencias Bioquı´micasy Farmace´uticas, Suipacha, 531, 2000 Rosario, Argentina; and †Department of , , Lensfield Road, Cambridge CB2 1EW,

his perspective has a simple message: lar recognition has become a versatile tool TCovalent synthesis under thermody- that enables control over equilibrium sys- namic control can be a powerful tool for tem. Selective recognition of one of the preparing mechanically interlocked struc- components in an equilibrium system will tures and specific macrocycles in high result in the creation of an additional well in yield and at the same time serve as an the Gibbs energy landscape (Fig. 1). If this engine for the creation of dynamic com- well is deep enough, the recognized com- binatorial libraries (DCLs), which are re- pound will dominate the mixture at the sponsive to their environment. Here we expense of undesired products. In practice will highlight some examples that illus- this outcome requires addition of a species trate the potential of thermodynamically (a template) that selectively recognizes, controlled covalent chemistry and we will binds, and stabilizes the desired compound. point out some exciting directions for the In a sense, this template plays a role similar future. A more comprehensive review of to that of a catalyst in kinetically controlled the area has been provided elsewhere (1). synthesis. Most of the reactions used in organic Despite the fact that equilibrium sys- synthesis are irreversible in practice. The tems seem much more tractable than ki- Fig. 1. Free energy landscape of an equilibrium outcome of a given procedure is determined netically controlled transformations, rela- mixture showing the effect of adding a template that strongly and selectively binds to one of the by the kinetics of the desired reaction as well tively little use has been made of them so equilibrating species. as the kinetics of the undesired competing far in synthesis. Although there are good processes. Hence, yields and product distri- reasons not to strive for equilibrium dis- PERSPECTIVE butions reflect the relative differences be- tributions in many instances, there are are summarized in Fig. 2. They can be tween the various initial and transition-state other cases where thermodynamically divided into two categories, depending on Gibbs energies. Optimization of such kinet- controlled synthesis is the way forward. In the symmetry of the reversible linkage. ically controlled processes involves changing particular, thermodynamic control is exchange (2, 3) and the Gibbs energies of initial states or, more highly advantageous in the synthesis of metathesis (4) are symmetrical. Disulfide frequently, transition states. Impressive macrocycles and mechanically interlocked exchange is mediated by a catalytic achievements have been made following compounds. Also, combinatorial chemis- amount of thiolate anion attacking the this approach mainly through the use of try is more versatile when operated under existing disulfide bonds. Exchange pro- SPECIAL FEATURE catalysts. Nevertheless, improving kineti- thermodynamic control. As we will show ceeds readily under neutral to mildly basic cally controlled reactions has been and still below, dynamic combinatorial chemistry conditions, whereas acidifying the me- remains a challenge, largely as a result of the enables high-yield synthesis of, for in- dium causes protonation of the thiolate elusive nature of the activated complex, stance, a host in one step with- anion and turns off exchange (2, 3). which does not lend itself to direct study. out prior knowledge of its exact structure. Alkene metathesis (4) is controlled The situation is further complicated by Reversible Covalent Chemistry through addition or removal of a catalyst. the fact that effects on the Gibbs energy The usefulness of this reaction has been of the initial state need to be taken into A major reason why the use of reversible increasing rapidly, as new, more efficient, consideration. covalent chemistry is not more widespread and general catalysts become available. For thermodynamically controlled reac- is the rather limited number of reversible Reversible reactions that feature un- tions the situation is much less complex. reactions available. Furthermore, many re- symmetrical linkages include ex- Product distributions now depend on only actions are reversible only under conditions change (5), transesterification (6–8), as the relative Gibbs energies (stabilities) of that are not compatible with the subtle well as exchange of (9, 10) and the products themselves. Of course, chem- noncovalent interactions involved in tem- ists have long made use of the fact that plating. Finally, after the equilibrium distri- derivatives thereof. Only one example of product distributions of some reactions are bution has been reached it must be possible amide exchange has been reported (5), under kinetic control under one set of con- to turn off the reversible reaction to isolate making use of enzymes to cleave and ditions, whereas the thermodynamic prod- and handle the desired product. In practice reform the amide bonds. Transesterifica- uct prevails under other conditions. But why this is most often accomplished by removal tion has received more attention. Unfor- not go one step further and actively manip- of the catalyst responsible for reversibility or tunately, this reaction generally requires ulate the Gibbs energies of the products so by altering the pH of the solution. harsh conditions that are likely to inter- as to shift the equilibria in any desired The reversible reactions that have thus direction? With the increased understand- far been used successfully in templated ‡To whom reprint requests should be addressed. E-mail: ing of supramolecular interactions, molecu- thermodynamically controlled synthesis [email protected] or [email protected].

www.pnas.org͞cgi͞doi͞10.1073͞pnas.022643699 PNAS ͉ April 16, 2002 ͉ vol. 99 ͉ no. 8 ͉ 4801–4804 Downloaded by guest on September 24, 2021 plate recognition. (11) and hydra- also respond to exterior influences. The zones (12) do not suffer from this disadvan- more fundamental of those influences, such tage. In these cases, exchange requires acid as temperature, pressure, and effects of and is switched off after neutralization to solvents and extents of ionization of the yield a robust product. , has received relatively little at- tention in the area of small molecule syn- Manipulating Equilibria thetic chemistry. Yet these parameters Reactions that are most likely to benefit might well prove to be useful handles in from being controlled by thermodynamics optimizing the proportion of desired prod- are processes such as macrocyclizations uct in equilibrium mixtures. where several products are accessible that More specific manipulation of an equi- differ in size rather than connecting chem- librium composition is possible by exposing istry. Ring-closure reactions performed un- the system to templates. Two types of tem- der kinetic control are notorious for giving plating approaches can be distinguished. In relatively large amounts of unwanted oligo- the synthesis of mechanically interlocked meric side products even under high- structures the template will usually be part dilution conditions. Under thermodynamic of the final product.§ Alternatively the tem- Fig. 2. Reversible reactions that have been used control, the product distribution of the same plate can be used simply as a mould, in successfully in thermodynamically controlled reaction is governed only by the stabilities of which case it will be removed once the synthesis. the different macrocycles. These stabilities equilibrium is frozen, giving access to the are partly determined by interactions within free receptor molecule. Even though these the macrocycles (e.g., ring strain) and partly approaches lead to topologically different fere with the weak interactions underlying by the concentration at which the reaction is products, they share two necessary features: recognition by a template (6, 7). In the carried out. Disregarding ring strain, rela- (i) a template effect that stabilizes the de- special case of allyl , Pd-catalyzed tively dilute conditions (less than 10 mM) sired product via supramolecular interac- transesterification proceeds under mild will favor smaller ring sizes (13). Higher tions and (ii) one or more reversible con- conditions (8). oligomers might still be formed in the initial nections allowing for proofreading and Reactions based on exchange around the stages of the reaction but they will be con- editing of undesired products (Fig. 3). C ϭ N bond are more versatile. The parent sumed as equilibrium is approached. As imines can be used, but they suffer from an concentrations are raised, the product dis- Thermodynamically Controlled Synthesis of inherent instability. Reduction of the imines tribution will gradually shift toward higher Mechanically Interlocked Structures. The in- to is generally required to stop the oligomers, culminating in polymer forma- terest in mechanically interlocked structures equilibration process and allow for isolation tion for very concentrated solutions (for a like catenanes (interlocked rings) and rotax- of the product (9, 10). These final amines are review of thermodynamically controlled po- anes (one or more rings mechanically locked geometrically and electronically different lymerizations, see ref. 1). onto a linear thread by bulky stoppers) has from the imines that are recognized by the Equilibrium systems are adaptive; apart been growing rapidly because of their po- template, potentially leading to loss of tem- from a concentration dependency they will tential for the development of nanoscale devices or molecular machines (14–16). The current strategies for the construc- tion of these mechanically interlocked struc- tures make use of noncovalent interactions that bring together the components of the superstructure in a specific orientation. The final step in the synthesis is the formation of one or more covalent bonds that fix the product in the interlocked orientation. In most of the syntheses reported to date, this final reaction is carried out under kinetic conditions. As a result, if the reaction does not proceed via a threaded precursor, un- desired noninterlocked products are pro- duced that cannot be rescued, reducing the yield of the target molecule. As a thermo- dynamically controlled approach includes bond breaking and reformation, it permits recycling of these side products to produce the correct structure. We explored the use of alkene metathesis under thermodynamic conditions for the preparation of [2]catenane 4 (Fig. 4) (17), exploiting ␲-donor͞␲-acceptor interactions between ␲-deficient aromatic diimides and ␲-rich aromatic diethers to stabilize the

§Synthesis of mechanically interlocked structures can also be performed by using removable templates like metal Fig. 3. Overview of different structures that benefit from thermodynamic control. . See ref. 14.

4802 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.022643699 Furlan et al. Downloaded by guest on September 24, 2021 stable interlocked product. The attractive feature in Stoddart’s system is that reduction of the bonds is faster in the desired rotaxane than in the nonthreaded macro- cycles (like 9); this results in a further con- structive shift in the equilibrium between the threaded and unthreaded macrocycles leading to the formation of the stable [2]ro- taxane 12 in 70% isolated yield.

Templating Dynamic Combinatorial Libraries (20–22). Allowing thermodynamics to con- trol the composition of combinatorial librar- ies can give truly remarkable results and Fig. 4. Thermodynamically controlled preparation of a [2]catenane by using ␲-donor͞␲-acceptor provides the most compelling demonstra- interactions and alkene metathesis. tion of the potential of reversible covalent chemistry. We have developed dynamic combinatorial libraries (6) with a view to desired product. The ring-closing metathe- if only the macrocycle had been exposed to generating macrocyclic receptors. The li- 1 the catalyst at that concentration. sis of dialkene leads to the formation of braries are generated by linking macrocycle 2, accompanied presumably by a The thermodynamic preference of this building blocks together through reversible variety of linear and cyclic oligomeric spe- system can be drastically affected by bonds. A guest molecule is then introduced cies (Fig. 4). When half a molar equivalent switching off the -bonding ability and allowed to select its ideal host from of crown 3 and additional catalyst were of the macrocycle by trifluororoacetyla- among the available macrocycles in the mix- added, [2]catenane 4 was produced as a tion of the amide functionality. In the ture. Binding of the guest to its preferred mixture of cis͞trans isomers in 50% isolated absence of the catalyst, the trifluoroacety- host leads to stabilization and concomitant yield. lated [2]catenane is stable while in its shift of the equilibrium toward this species: A more subtle example of the thermody- presence the catenane unthreads to form the best binder will be amplified. Thus, namic approach has been reported by Leigh the noninterlocked macrocycles. introduction of the guest will convert an and coworkers (18) who prepared [2]cat- The saturated analog of [2]catenane 6 initially diverse library into the desired mol- enane 6. The amide macrocycle 5 is self- had been previously prepared by using ecule, ideally in high purity, thereby com- complementary and can dimerize by form- kinetically controlled condensation of bining identification and isolation of the ing four hydrogen bonds between the amide acid chlorides and masked phenols, but as target receptor into one step. hydrogen and the carbonyls undesired products could not be recycled Our initial attempts to generate dynamic the highest yield obtained was 18% (19). combinatorial libraries focused on transes- (Fig. 5). In addition, it contains an internal PERSPECTIVE Stoddart and colleagues (10) reported the alkene available for metathesis under re- terification (6, 7, 23) but we have since template-directed preparation of [2]rotax- versible thermodynamic control. Under ap- moved on to the more practical ane 12 under thermodynamic conditions. In propriate ring-opening and ring-closing alk- and disulfide exchange reactions. Generat- this case, the interaction between a proton- ene metathesis conditions, macrocycle 5 ing diverse dynamic combinatorial libraries ated dialkylammonium center and a crown based on these chemistries is straightfor- self-assembles to give [2]catenane 6 in was used as recognition motif for the ward. Hydrazone libraries form spontane- Ͼ95% yield. The amount of catenane stabilization of the rotaxane, whereas imine ously after adding acid to a solution of formed depends on the initial concentration formation͞breaking was used as ‘‘correct- building blocks such as 13 containing a

of the macrocycle used. At low concentra- ing’’ reversible reaction (Fig. 6). hydrazide and an (protected as an SPECIAL FEATURE tion (0.0002 M), only the noninterlocked On mixing of dialdehyde 7 and diamine 8 ) functionality (12). Dynamic libraries macrocycle is present in the reaction mix- in chloroform, an interconverting mixture of of macrocyclic form readily simply ture, whereas above 0.2 M higher cyclic macrocycle 9 and higher cyclic and linear after dissolving building blocks in oligomers become increasingly abundant. oligomers is obtained. Addition of the water at pH 7 and stirring the solution in an That the reactions are under thermody- thread (10) that binds 9 with a binding open flask. from the air is sufficient namic control was demonstrated by taking constant of 300 MϪ1, shifts the equilibrium to convert the to disulfides. While the product mixture obtained at one con- of the interconverting imines toward the oxidation is taking place the remaining centration and reexposing it to the catalyst formation of the [2]rotaxane 11. Because mediates the exchange process responsible at a different concentration. In all cases, the imine bonds are kinetically labile, a fixation for equilibration (2). product distribution adjusts to that obtained step by reduction is necessary to obtain a Exposure of these libraries to templates has resulted in dramatic shifts in the distri- bution of the library in a number of cases (6, 24–27). In one instance (26), cyclization of building block 13 derived from L--L- phenylalanine in chloroform solution ini- tially produced a mixture of at least 10 interconverting macrocycles. When the dy- namic mixture was exposed to LiI it trans- formed to essentially pure trimer 14 (Fig. 7). Binding studies demonstrated that the se- lected trimer forms a 1:1 complex with Liϩ with a binding constant of 4 ϫ 104 MϪ1. Fig. 5. Thermodynamically controlled preparation of a [2]catenane by using hydrogen bonds and alkene This example clearly illustrates the power of metathesis. dynamic combinatorial chemistry: a new

Furlan et al. PNAS ͉ April 16, 2002 ͉ vol. 99 ͉ no. 8 ͉ 4803 Downloaded by guest on September 24, 2021 Complementary to our work, others have studied dynamic combinatorial libraries by using hosts to template the formation of guest molecules (9, 28). When using pro- teins or nucleic acids as the host molecules, this approach has considerable potential for drug development (for a recent review, see ref. 22). Outlook The attractions of the use of reversible co- valent chemistry under thermodynamic conditions have been illustrated by the prep- aration of several supramolecular architec- tures obtained through rational design or with a combinatorial approach. In these examples a bond breakage process is part of the synthetic strategy. Although bond Fig. 6. Thermodynamically controlled preparation of a [2]rotaxane by using ammonium͞crown ether breakage is an intrinsic part of thermody- interaction and transimination. namically controlled chemistry, it would be challenging to use this strategy in a kineti- cally controlled system. receptor molecule was identified while at more, it is unlikely that the selected receptor Although the thermodynamic approach the same time the thermodynamically con- would have been the outcome of any ratio- has in a short time earned its place among trolled chemistry immediately provided an nal design. In fact, even knowing that 1 binds the more useful tools available to supramo- efficient synthetic route toward an other- strongly to Liϩ we are still rather puzzled as lecular chemists, we believe that it is not yet wise difficult to access molecule. Further- to how it actually manages to do so. used to its full potential. For instance, ther- modynamically controlled chemistry seems the method of choice for the synthesis of capsule-like molecules from relatively easily prepared trifunctional building blocks. First steps on this road have recently been made (29, 30). Furthermore, dynamic combinato- rial chemistry seems a highly promising method for the identification of new su- pramolecular catalysts. Analogous to the catalytic approach, using a transi- tion state analogue as a template it should be possible to identify catalytically active receptors. These targets are now being ac- tively pursued.

We are grateful for support from the Marie Curie Individual Fellowships Program (HPMF- CT-1999-00069) and the Royal Society Univer- sity Research Fellowships Program (to S.O.), the Fundacio´n Antorchas and the Consejo Na- cional de Investigaciones Cientı´ficas y Tecnicas (Argentina) (to R.L.E.F.), and the Engineering Fig. 7. A dynamic combinatorial library of hydrazone-based pseudopeptide macrocycles. Shown are the and Physical Sciences Research Council HPLC traces of the library before (Left) and after (Right) addition of Liϩ template. (to J.K.M.S.).

1. Rowan, S. J., Cantrill, S. J., Cousins, G. R. L., Sanders, 10. Glink, P. T., Oliva, A. I., Stoddart, J. F., White, A. J. P. & 20. Karan, C. & Miller, B. L. (2000) Drug Discov. Today 5, 67–75. J. K. M. & Stoddart, J. F. (2002) Angew. Chem. Int. Ed. Williams, D. J. (2001) Angew. Chem. Int. Ed. Engl. 40, 21. Lehn, J.-M. & Eliseev, A. V. (2001) Science 291, 117–125. Engl. 41, in press. 1870–1874. 22. Otto, S., Furlan, R. L. E. & Sanders, J. K. M. (2002) Drug 2. Otto, S., Furlan, R. L. E. & Sanders, J. K. M. (2000) J. Am. 11. Polyakov, V. A., Nelen, M. I., Nazarpack-Kandlousy, N., Discov. Today 7, 24–32. Chem. Soc. 122, 12063–12064. Ryabov, A. D. & Eliseev, A. V. (1999) J. Phys. Org. Chem. 12, 23. Brady, P. A., Bonar-Law, R. P., Rowan, S. J., Suckling, C. J. 3. Ramstro¨m, O. & Lehn J.-M. (2000) ChemBioChem. 1, 357–363. & Sanders, J. K. M. (1996) Chem. Commun., 319–320. 41–48. 12. Cousins, G. R. L., Poulsen, S.-A. & Sanders, J. K. M. 24. Furlan, R. L. E., Cousins, G. R. L. & Sanders, J. K. M. 4. Grubbs, R. H. & Chang, S. (1998) Tetrahedron 54, 4413– (1999) Chem. Commun., 1575–1576. (2000) Chem. Commun.1761–1762. 4450. 13. Ercolani, G., Mandolini, L., Mencarelli, P. & Roelens, S. 25. Cousins, G. R. L., Furlan, R. L. E., Ng, Y.-F., Redman, 5. Swann, P. G., Casanova, R. A., Desai, A., Frauenhoff, (1993) J. Am. Chem. Soc. 115, 3901–3908. J. E. & Sanders, J. K. M. (2001) Angew. Chem. Int. Ed. M. M., Urbancic, M., Slomczynska, U., Hopfinger, A. J., 14. Sauvage, J.-P. (1990) Acc. Chem. Res. 23, 312–327. Engl. 40, 423–428. Le Breton, G. C. & Venton, D. L. (1996) Biopolymers 40, 15. Amabilino, D. B. & Stoddart, J. F. (1995) Chem. Rev. 95, 26. Furlan, R. L. E., Ng, Y.-F., Otto, S. & Sanders, J. K. M. 617–625. 2725–2828. (2001) J. Am. Chem. Soc. 123, 8876–8877. 6. Brady, P. A. & Sanders, J. K. M. (1997) J. Chem. Soc. 16. Balzani, V., Credi, A., Raymo, F. M. & Stoddart, J. F. 27. Furlan, R. L. E., Ng, Y.-F., Cousins, G. R. L., Redman, Perkin Trans. 1, 3237–3253. (2000) Angew. Chem. Int. Ed. Engl. 39, 3348–3391. J. E. & Sanders, J. K. M. Tetrahedron (2002) 58, 771–778. 7. Rowan, S. J. & Sanders, J. K. M. (1998) J. Org. Chem. 63, 17. Hamilton, D. G., Feeder, N., Teat, S. J. & Sanders, J. K. M. 28. Karan, C. & Miller, B. L. (2001) J. Am. Chem. Soc. 123, 1536–1546. (1998) New J. Chem. 22, 1019–1021. 7455–7456. 8. Kaiser, G. & Sanders, J. K. M. (2000) Chem. Commun., 18. Kidd, T. J., Leigh, D. A. & Wilson, A. J. (1999) J. Am. 29. Tam-Chang, S.-W., Stehouwer, J. S. & Hao, J. (1999) J. 1763–1764. Chem. Soc. 121, 1599–1600. Org. Chem. 64, 334–335. 9. Huc, I. & Lehn, J.-M. (1997) Proc. Natl. Acad. Sci. USA 94, 19. Leigh, D. A., Moody, K., Smart, J. P., Watson, K. J. & Slawin, 30. Ro, S., Rowan, S. J., Pease, A. R., Cram, D. J. & Stoddart, 2106–2110. A. M. Z. (1996) Angew. Chem. Int. Ed. Engl. 35, 306–310. J. F. (2000) Org. Lett. 2, 2411–2414.

4804 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.022643699 Furlan et al. Downloaded by guest on September 24, 2021