Skip to main content
Log in

Emergences of supramolecular chemistry: from supramolecular chemistry to supramolecular science

  • Original Article
  • Published:
Journal of Inclusion Phenomena and Macrocyclic Chemistry Aims and scope Submit manuscript

Abstract

We describe the field of supramolecular chemistry as a consequence of the progress of chemistry from its premises to recent achievements. Supramolecular chemistry has been claimed to be an emergent field of research taking its roots in chemistry. According to the definitions of emergences related to hierarchy or more recently to scope, supramolecular chemistry is shown to have bottom-up or top-down emergences. The bottom-up emergence, directly related to hierarchy by definition, opens up the world of nanochemistry and nanomaterials while the top-down one, attributable to scope due to the implication of supramolecular chemistry in other fields of research, open the world of supramolecular biochemistry. Both emergences lead supramolecular chemistry to become a supramolecular science. Combining supramolecular chemistry with biology opens new direction in the study of life and it origin.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. On many levels, and to many people, chemistry and art are not easily correlated. From the DaVinci Project a conceptual framework for the development of project materials and activities was identified. This framework was based on three levels of complexity of ideas: structures, interactions between and among structures, and applications of structures and interactions to the everyday, observable environment. The development of higher thinking skills was fostered during the entire project, especially during the development of the third phase of the DaVinci Multimedia System. The conclusion that chemists and artists behave similarly creative for reasons evidenced during the DaVinci Project is given in Simonson and Schlosser [110].

  2. One example can be found of supramolecular chemistry giving rise to art works due to the fascinating shapes of molecules: Balzani et al. [111].

  3. Words of French philosopher Michel Foucault describe this type of thinking: ‘J’aurais aimé m’apercevoir qu’au moment de parler une voix sans nom me précédait depuis longtemps: il m’aurait suffi alors d’enchaîner, de poursuivre la phrase, de me loger, sans qu’on y prenne bien garde, dans des interstices, comme si elle m’avait fait signe en se tenant, un instant, en suspens. De commencement il n’y en aurait donc pas’ [112].

  4. A recent account deals with the matter symbolized by alchemists and the evolution of alchemy to chemistry: see L. Fabbrizzi [113].

  5. According to H. H. Kubbinga [114] two first ‘molecular’ theories have been conceived by Isaac Beeckman (1588–1637), a Dutch philosopher and scientist, and Sebastian Basso (1573–?), a French physician and natural philosopher, introducing the concept of ‘substantial individuals’ (Beeckman, 1620) and ‘substantial species’ (Basso, 1620). Beeckman and Basso are also credited by Kubbinga to have anticipated the concept of isomers. Their theories gave quite important consequent concepts such as the ‘minima sui generis’ of Daniel Sennert (1592–1637), a German physician, the ‘particles’ of René Descartes (1596–1650), a French philosopher, mathematician and physicist, and Christian Huygens (1629–1695), a Dutch mathematician, astronomer, and physicist, the ‘monads’ of Gottfred Leibniz (1649–1716), a German philosopher and mathematician, the ‘molecules’ of Georg Ernst Stahl (1659–1734), a German chemist and physician, and the ‘particulae ultimate compositionis’ of Isaac Newton (1643–1727) an English physicist, mathematician, astronomer and chemist. In the same paper, H. H. Kubbinga also gives a reference indicating that the word ‘molecule’ has been introduced by Gassendi.

  6. The existence of the ‘phlogiston’ (Greek φλογιστόν phlŏgistón = burning up, from φλόξ phlóx = fire) as the fifth element with air, earth, fire and water, was first proposed in 1667 by German physician and chemist Johann Joachim Becher (1635–1682) to explain processes such as combustion and the rusting of metals. He published his Physical Education, which was the first mention of what would become the phlogiston theory. In his book, Becher eliminated fire and air from the classical element model and replaced them with three forms of earth: terra lapidea, terra fluida, and terra pinguis. Terra pinguis was the element which imparted oily, sulfurous, or combustible properties. Terra pinguis was a key feature of combustion and was released when combustible substances were burned. In 1703, Georg Ernst Stahl (1659–1734), a German physician and chemist, proposed a variant of the theory in which he renamed Becher's terra pinguis to phlogiston, and it was in this form that the theory probably had its greatest influence. Lavoisier and before him Mikhail Lomonosov (1711–1765) a Russian chemist who had expressed his ideas during 1748 and proved them by experiments, showed that in fact phlogiston does not exist and that combustion consumes the oxygen of the air. The German philosopher Immanuel Kant (1724–1804) was first in favor of the phlogiston theory of Stahl but later supported the Lavoisier’s proposal of oxygen consuming. The no-need of phlogiston was expressed in the Méthode de Nomenclature Chimique of de Morveau, Lavoisier, Bertholet and de Fourcroy in 1787 and in the Traité Elémentaire de Chimie of Lavoisier in 1789.

  7. The Méthode de Nomenclature Chimique and the proposal made by Berzelius of designating the elements by the one or two first letters were not only useful for the communication between scientists but also they provide the memory with what is named a double mediation: tools and organs. This is mentioned in the preface of the books by Lavoisier who noted that due to the simplification of new naming of substances will provide the chemists with a new common language with no possible misinterpretation. This new naming is necessary to learn and give a mathematic-like logic to chemical science in such a way that names or words—directly based on experiments—are giving birth to ideas. That means at the same time that memory is organised and organising. And this leads Lavoisier to write 2 years later the Traité Elémentaire de Chimie in which he developed the concepts of modern chemistry. Examples can be found in other fields than chemistry. For example, it has been shown that the effect of the appearing of script was not to memorise oral culture but to introduce a graphic rationality (http://www.mediologie.org). See also Merzeau [115] and Goody [116].

  8. By the middle of the 19th century chemists generally understood that chemical elements can be grouped together in separate classes according to similarities and dissimilarities in their properties. For example alkali metals are flammable and form cations while halogens are poisonous and form anions. A recent article has appeared on the subject of atoms and molecules, and periodicity: Babaev and Hefferlin [117].

  9. A covalently formed bond can be considered to be as strong as an irreversible linkage between atoms. This property has been synthetically exploited, from forming one bond at a time, to methodically building larger and larger covalent structures from smaller molecular starting materials. For a long time, this was the only available method to produce a molecule with a desired shape and function, with common molecular targets having less than 100 covalent bonds and molecular weights of several hundred Daltons. Some of the largest structures synthesized at the upper limits of covalent synthesis, palytoxins, have molecular weights of several thousand Daltons and lengths of around one nm. See as an example Armstrong et al. [118].

  10. Attempts to rationalize the periodic table have included reduction to quantum mechanics as well as approaches from mathematical chemistry. However quantum mechanics does not provide a conclusive means of classifying certain elements like hydrogen and helium into their appropriate groups. An alternative approach using atomic number triads is proposed and the validity of this approach is defended in the light of some prediction made via the information theoretical approach that suggests a connection between nuclear structure and electronic structure of atoms [119].

  11. The molecular structure hypothesis—that a molecule is a collection of atoms linked by a network of bonds—was forged through experiments during the nineteenth century. It has continued to serve as the principal means of ordering and classifying the observations of chemistry. However, this hypothesis was not related directly to quantum mechanics which governs the motions of the nuclei and electrons that make up the atoms and the bonds. Indeed there was, and with some there still is, a prevailing opinion that these fundamental concepts, while unquestionably useful, were beyond theoretical definition. Chemists have an understanding based on a classification scheme that is both powerful and at the same time, because of its empirical nature, limited. Richard Feynman and Julian Schwinger have given a reformulation of physics that enables one to pose and answer the questions ‘what is an atom in a molecule and how does one predict its properties?’ It was demonstrated that this new formulation of physics, when applied to the topology of the distribution of electronic charge in real space, yields a unique partitioning of some total system into a set of bounded spatial regions. The form and properties of the groups so defined faithfully recover the characteristics ascribed to the atoms and functional groups of chemistry. By establishing this association, the molecular structure hypothesis is freed from its empirical constraints and the full predictive power of quantum mechanics can be incorporated into the resulting theory—a theory of atoms in molecules (AIM) and crystals.

    AIM recovers the central operational concepts of the molecular structure hypothesis, that of a functional grouping of atoms with an additive and characteristic set of properties, together with a definition of the bonds that link atoms and impart the structure. Not only does the theory thereby quantify and provide the physical understanding of the existing concepts of chemistry, it makes possible new applications of theory as for example enabling one to perform on a computer, in parallel to experiment, everything that can now be done in the laboratory thus linking together the functional groups of theory. AIM enables one to take advantage of the single most important observation of chemistry, that of a functional group with a characteristic set of properties. This outlines and illustrates the topological basis of the theory and its relation to the quantum mechanics of an open system. However AIM cannot be directly observed by experiment, nor can one measure enough properties of an atom in a molecule to define it unambiguously. They are multiple ways to partition molecules into atoms that are consistent with various observed chemical trends and experimental data. And some authors emphasized that because of this ambiguity atoms in molecules remains a noumenon in the sense given to this word by Immanuel Kant [120]. Similarly, ‘Few people seriously argue that atoms and molecules are fictitious, but there is no picture or model of a atom that is equivalent to a photograph of an object at the human scale’. These words are taken from a publication of Chris Toumey dealing with the relation between an object and an image of the object [121]. He asks the questions about representations of atoms and molecules by STM and AFM techniques which are seen as indirect techniques: What is a faithful reproduction? How do technical processes affect the image? Nanoscale images evoke these issues. To enhance our visual knowledge of nanoscale objects he revisits earlier cubist theory. This lead to suggestions in a neo-cubist spirit for making and seing nanoscale structures [121, 122].

  12. Marcellin Berthelot noted that ‘chemistry creates its own objects’. These words are important in the sense that due to the possibility given to chemists to develop reactions for synthesis from simple elements—and one can say now from the table of elements of Mendeleiev—the work of chemists has been the reverse to the achievement of alchemists and chemists until the beginning of the 20th century. In other words chemists have analysed and separated chemical substances into simple elements that have been used further to prepare molecules with higher structures, successively going from complex to simple and from simple to complex.

  13. For further reading on ‘molecular beauty and chemists’ imagination see also: Spector and Schummer [123].

  14. The study of interlocked molecules is now an emerging field of research dealing with catenanes and rotaxanes. Catenanes are chemical structures in which two or more molecules are interlocked while in rotaxanes one or more macrocycles are mechanically prevented from dethreading from a liner unit by bulky ‘stoppers’. Such interlocked systems that at the beginning were curiosity or challenge of chemists are now related to molecular machines and present applications in molecular devices. A wide and important review on these molecular systems has been published by Kay et al. [124].

  15. Paul John Flory (1910–1985) is an American chemist who was known for his prodigious volume of work in the field of polymers and macromolecules. He was a leading pioneer in understanding the behavior of polymers in solution, and won the Nobel Prize in Chemistry in 1974 ‘for his fundamental achievements, both theoretical and experimental, in the physical chemistry of macromolecules.’

  16. Dendrimers (greek dendron = tree) are macromolecules with a tree-like structure. They are synthetically built from a core with repeating units as molecular branches and terminated by end groups. The preparation of such branched structures demands the use of particular building blocks with appropriate stereochemistry and multiple, equivalent reaction centres. See for example Newkome et al. [125].

  17. This sentence of Sismour and Benner is reminiscent of the concept of double mediation that is memory is organised, organising and in the same run involved in the creation processes of synthesis of molecules. Another very important concept used by the memory of organic chemists is the use of curly arrows to represent the movement of electrons. Curly arrows are currently used for both explaining reaction mechanisms and to anticipate reactions. Robert Robinson (1886–1975), Nobel Prize for Chemistry in 1947 is credited for the invention of this useful tool [126]. The atomic theory developed during the first half of the twentieth century was not readily accepted by all organic chemists and its acceptance has been made difficult by French organic chemist [127].

  18. Van’t Hoff is not only one of the fathers of 3D-chemistry as we know it today, but he is probably also the father of Molecular Origami’s. Origami is a Japanese word to designate the art of folding (oru) the paper (kami). Molecular origami’s use this art and the way of cutting and folding to represent molecules or crystalline solids. It is related to the knowing of matter by X-ray diffraction. It is used by chemists and biologists. For example the work of R. M. Hanson (Ed.) Molecular Origami Mass Scattered Paper Models proposes the precise-scale construction in 3D with angles and distances with variously coloured papers. Molecular Origami’s images are also used by softwares to calculate molecular structures as large as the ones of nanochemistry. This type of modelisation-representation has been extended to DNA and related molecules which can be seen with CAChe programs [128].

  19. One can as the question that if matter is made of atoms and molecules why the three dimensions of crystals has not been taken into account to figure out that molecules being the smallest part of them have three dimensions too. Crystals have long been collected, sold, manufactured, and admired for their regularity, but they were not investigated scientifically until the 17th century. Johannes Kepler (1571–1630), a German mathematician, described in Strena seu de Nive Sexangula (1611) the hexagonal symmetry of snowflake crystals as a regular packing of spherical water particles. Nicolas Steno (1638–1686), a Danish anatomist and geologist, showed in De solido intra solidum naturaliter contento dissertationis prodromus (1669) that the angles between the faces are the same in every sample of a particular type of crystal opening the road to crystal geometry or crystallography. Later on, René Just Haüy (1743–1822), a French mineralogist, reported in Essai d’une théorie sur des crystaux (1784) that every face of a crystal can be described by simple stacking patterns of blocks he named ‘molécules intégrantes’ of the same shape and size: crystals are a regular three-dimensional array of atoms and molecules in which a single unit cell is repeated indefinitely along three principal directions. As a consequence in 1839, William Hallowes Miller (1801–1880), a British mineralogist and crystallographer, was able to give each face a unique label of three small integers, the Miller indices which are still used today for identifying crystal faces. In the 19th century, a complete catalogue of the space groups of a crystal was elaborated by J. Hessel, A. Bravais, Y. Fyodorov, and A. Schönflies. In 1880, William Barlow (1845–1934), an English crystallographer, proposed structural models of NaCl and CsCl that were later confirmed using X-ray crystallography. All these studies were made on mineral crystals. One can assume that no link was made at that time between three dimensions of mineral matter and organic molecules. The studies on crystals were mainly based on mathematics and reasoning while studies on the matter was mainly based on obtaining a chemical structure by analysis. The first structures of organic molecules were lately published by William Lawrence Bragg (1890–1971), an English physicist in 1921 thanks to X-ray techniques [129].

  20. Curiously, the concept of molecular graph was first conceived by Arthur Cayley (1821–1895), a British mathematician in 1847. Two types of molecular graphs were proposed: plerograms and kenograms. Plerograms are molecular graphs in which all the atoms are represented by vertices. Kenograms are what nowadays is referred to as hydrogen-suppressed or hydrogen-depleted molecular graphs: See Caley [35, 130, 131].

  21. Picture or visual thinking is the ability of thinking through images and not through words using the part of the brain that is emotional and creative to organize information in an intuitive and simultaneous way. Thinking in pictures, is one of a number of other recognized forms of non-verbal thought such as kinesthetic, musical and mathematical thinking. It is nonlinear and often has the nature of a computer simulation, in the sense that a lot of data is put through a process to yield insight into complex systems, which would be impossible through language alone. ‘Visual thinking calls for the ability to see visual shapes as images of the patterns of forces that underlie our existence—the functioning of minds, of bodies or machines, the structure of societies or ideas.’ are the words of Arnheim [38]. A recent book has appeared in which A. J. Rocke emits the thesis that chemists have been progressing in defining molecular structures because ‘human minds work far more visually, and less purely linguistically than we realize’ [132].

  22. The use of ‘key and lock’ complementariness was also often used by German physiologist Paul Ehrlich (1854–1915; Nobel Prize in 1908). This is mentioned in 1946 by Pauling [133] in an article showing how many basic problems of biology—nature of growth, mechanism of duplication of viruses and genes, action of enzymes, mechanism of physiological activity of drugs, hormones, and vitamins, structure and action of nerve and brain tissue—find answers in the knowledge of molecular structure and intermolecular reactions. Many examples are given which are explained by the underlying concept of molecular recognition involving ‘shape and size and the detailed nature of intermolecular forces.’ And in his paper dated of 1946, Pauling uses words and expressions that will be used later on to describe in some parts contemporary chemistry, as exemplified: molecular architecture, size and shape, intermolecular forces, duplication, storage battery, machinery, ring gear, brake pedal, hydrogen bonds, Van der Waals forces, surface regions, complementary in structure, fit into the molecule of the recipient combination, specificity, and, fit into the cavity [133].

  23. Molecular engineering was first introduced by A. I. Kitaigorodskii as the concept of close packing for molecular crystals giving to molecules shape and volume [134, 135]. Molecular recognition in organic crystals is nowadays in various subjects of research [136].

  24. In 1888, the Austrian chemist Friedrich Reinitzer (1857–1927), working in the Institute of Plant Physiology at the University of Prague, discovered a strange phenomenon. He was determining the melting point of a derivative of cholesterol and he noticed two melting points. At 145.5 °C the solid melted into a cloudy liquid which existed until 178.5 °C where the cloudiness disappeared, giving way to a clear transparent liquid. He turned for help to the German physicist Otto Lehmann (1855–1922) who realized that the cloudy liquid is new state of matter (three states were known: solid, liquid, and gas) and coined the name ‘liquid crystal’, illustrating that it was something between a liquid and a solid. In liquids the properties are isotropic, i.e. the same in all directions. In liquid crystals they are not. Liquid crystals were already submitted to supramolecular concepts.

  25. Self-organisation is manifested in any complex system, which is in any system rich in interactions between its elements. Such systems are found in the physics of condensed matter, chemistry, biology, economics, social sciences, computer sciences etc. It is the complexity that is essential for the emergence of new phenomena, and self-organisation is the most striking but not the only example. The word ‘complexity’ comes from the Latin roots: ‘com’ meaning ‘together’, and ‘plectere’ meaning ‘to plait’. Like most of the concepts in chemistry, the concept of complexity seems rather fuzzy and even subjective. Complexity is so wide-ranging that nobody knows quite how to define it, or even where its boundaries lie. In fact there are different kinds of complexity and no single concept could embrace all the aspects of complexity. Scientists generally agree that the more complex the system, the less predictable it is. A typical complex system is one for which at least some of its global behaviors ‘that result cannot be predicted simply with the rules of the underlying interactions’: see the excellent and informative chapter of Bonchev and Seitz [137].

  26. Some molecular interactions were known from earlier 20th century and hydrogen bonding was termed as a weak bond in the case of water and ammonium hydroxide [138, 139].

  27. ‘It is impossible to dissociate language from science or science from language, because every natural science always involves three things: the sequence of phenomena on which the science is based; the abstract concepts which call these phenomena to mind; and the words in which the concepts are expressed. To call forth a concept a word is needed; to portray a phenomenon, a concept is needed. All three mirror one and the same reality’. Antoine Laurent Lavoisier, Traité Elémentaire de Chimie.

  28. According to J.-M. Lehn: ‘Definitions have a clear, precise core but often fuzzy borders, where interpenetration between areas takes place. These fuzzy regions in fact play a positive role since it is often there that mutual fertilization between areas may occur. This certainly is also true for the case at hand, the case of supramolecular chemistry and its language’ and language seems to be one of the driving forces that allows ideas to come. For the evolution and need of concepts and new names for chemistry to advance see also: Shaik [140] and Childs [141].

  29. Metal–Organic Frameworks (MOF’s) are crystalline compounds consisting of metal ions or clusters coordinated to often rigid organic molecules to form one-, two-, or three-dimensional structures that can be porous. In some cases, the pores are stable to elimination of the guest molecules (often solvents) and can be used for the storage of gases such as hydrogen and carbon dioxide. Other possible applications of MOFs are gas purification, gas separation, catalysis and sensors. This field of research is in full development. Several names are associated to this research which can be found in review articles [142144]. We would like to take the opportunity of this special issue dedicate to Pr Leonard F. Lindoy to mention one of his papers dealing with this chemistry [145].

  30. Supramolecular chemistry has a vocabulary borrowed from other disciplines, mostly in human sciences. For example the word entropy is as much a part of the language of the physical sciences as it is of the human sciences. Fortunately or unfortunately, physicists, chemists, and sociologists use indiscriminately a number of terms that they take to be synonymous with entropy, such as disorder, probability, noise, random mixture, heat. And all use terms such as information, complexity, organisation, order, selection, etc. But they are used in different sense. For example a ‘supramolecular complex system’ is different from the concept of ‘complex system’. One can make a short glossary of the words in use and see that the confusion and mixing of the words have lead in parts to the ideas and concepts of supramolecular chemistry, the experimental evidence of the concepts being obviously given by chemists:

    • Self-organisation is a process where the organisation of a system spontaneously increases without the control by the environment or an external system.

    • Selection is the quantity of variety: some of the possibilities or alternatives are eliminated, others are retained. The result is a constraint: a limitation of a number of possibilities.

    • Systems Theory is the transdisciplinary study of the abstract organisation of phenomena, independent of their substance, type, or spatial or temporal scale of existence. It investigates both the principles common to all complex entities, and the (usually mathematical) models which can be used to describe them.

    • Constraint is a measure of the reduction of variety or reduction of freedom.

    • Complexity (not a supramolecular complex!) has many definitions all falling short in one respect or another, classifying something as complex which we intuitively would see as simple, or denying an obviously complex phenomenon the label of complexity. These definitions are either only applicable to a very restricted domain, such as computer algorithms or genomes, or so vague as to be almost meaningless. Complexity comes from the Latin word complexus, which means ‘twisted together’. A complex is made of two or more objects, joined in such a way that it is difficult to separate them. Here is found the basic duality between parts which are at the same time distinct and connected. A system would be more complex if more parts could be distinguished, and if more connections between them existed.

    • Statistical entropy is a probabilistic measure of uncertainty or ignorance and information is a measure of a reduction in that uncertainty.

    • Fit is an assumed property of a system that determines the probability that that system will be selected, i.e. that it will survive, reproduce or be produced.

    • Dissipative structures characterizes a system that exits far from thermodynamic equilibrium, hence efficiently dissipates the heat generated to sustain it, and has the capacity of changing to higher levels of borderlines. Many definitions and useful explanations can be found on the site: http://pespmc1.vub.ac.be.

    At this level of such very simple glossary it would be to emphase that supramolecular chemists and cybernetics elaborate similar new concepts to improve their ability to communicate. For, example, L. M. Rocha introduced the concept of ‘Selected self-organization’ [146]. Self-organization is the spontaneous formation of well organized structures. They possess a large number of elements and variables and thus very large state spaces. But starting from some initial conditions they tend to converge in small areas of this space. It seems that when supramolecular chemists directed the synthesis of a system towards a desire self-organisation they are describing experimentally the concept of selected self organization. Such a selected self-organization concept has been named by Lehn in one of his papers ‘Self-organization by selection’ [147]. Self-organization by selection occurs through a two-level self-assembly with components selection driven by the formation of a specific product in a ‘self-design’ fashion [147].

  31. van Helmont, J. B.: Ortus Medicinae, Amsterdam, 108–109 (1648). The exact reference is given by Pross [97].

  32. In 1665, Robert Hooke (1635–1703), an English scientist, mathematician and architect published an important work titled Micrographia: Physiological Descriptions of Minute Bodies made by Magnifying Glasses and described an observation that changed basic biological theory and research. While examining a dried section of cork tree with a microscope, he observed small ‘monk’s chambers’ and coined the term ‘cell’ (from the Latin ‘cellula’ which means small compartment). Over the next 175 years, research led to the formation of the Cell Theory first proposed by Theodore Schwann (1810–1882) and Matthias Jacob Schleynden (1804–1881) both German physiologists by explicit claim that ‘there is only universal principle of development for the elementary parts, of organisms, however different, and this principle is the formation of cells’ [148]. Around 1833 Robert Brown (1773–1858) a Scottish botanist reported the discovery of nucleus (or areola as he called it) as ‘an opaque spot’ in the course of microscopic studies of epidermis of orchids. German doctor, Rudolph Ludwig Karl Virchow (1821–1902) was one of the first to give credit and plagiarise the work of Robert Remak (1815–1865), a Polish/German physiologist who showed that origins of cells was division of pre-existing cells.

  33. Origin of life has been the subject of wide research since the philosophers asked first the question. Origin of life can find some explanation in chemistry from two points of view. The first is the synthesis of the molecules the second is how the molecules could survive and evolve to life. In a publication entitled: ‘Chemistry and Selection’, Christian de Duve has reported how chemists have given rise to these two points of view [149].

  34. Recently a new step has been made by chemists to reproduce as near as possible natural molecules with desired functions. They started from the challenge of chemists in learning from Nature which combines four themes: chemical structure, function, size, and molecular shape. While structure and function are better understood, as can be seen in this work, size and shape remain challenging in synthesis. Thus, they prepared, by precise construction, a dendronized polymer that approximates the size and cylindrical shape of the tobacco mosaic virus pushing the chemical frontiers [150].

References

  1. Balian, R.: La radioactivité apprivoisée. Hôtel du Département des Bouches du Rhône, Septembre 9 (1994)

  2. Cintas, P.: The road to chemical names and eponyms: discovery, priority, and credit. Angew. Chem. Int. Ed. 43, 5888–5894 (2004)

    Article  CAS  Google Scholar 

  3. Klein, E.: Galilée et les indiens. Flammarion, Paris (2008)

    Google Scholar 

  4. Toumey, C.: Plenty of room, plenty of history. Nat. Nanotechnol. 4, 783–784 (2009)

    Article  CAS  Google Scholar 

  5. Lehn, J.-M.: From supramolecular chemistry towards constitutional dynamic and adaptative chemistry. Chem. Soc. Rev. 36, 151–161 (2007)

    Article  CAS  Google Scholar 

  6. Lehn, J.-M.: From matter to life: chemistry?!. Resonance 1996, 39–53 (1996)

    Article  Google Scholar 

  7. Stoddart, J.F.: Thither supramolecular chemistry? Nat. Chem. 1, 14–15 (2009)

    Article  CAS  Google Scholar 

  8. Pedersen, C.J.: The discovery of crown ethers (Nobel lecture). Angew. Chem. Int. Ed. Engl. 27, 1021–1027 (1988)

    Article  Google Scholar 

  9. Lehn, J.-M.: Supramolecular chemistry—scope and perspectives—molecules, supermolecules, and molecular devices (Nobel lecture). Angew. Chem. Int. Ed. Engl. 27, 89–112 (1988)

    Article  Google Scholar 

  10. Cram, D.J.: The design of molecular hosts, guests, and their complexes (Nobel lecture). Angew. Chem. Int. Ed. Engl. 27, 1009–1020 (1988)

    Article  Google Scholar 

  11. Marggraf, A.S.: A German chemist, first obtained concentrated formic acid by distillation of red ants and produced sugar substances from beets in 1746 (1709–1782)

  12. Russell, C.A.: Advances in organic chemistry over the last 100 years. Annu. Rep. Prog. Chem. B 100, 3–31 (2004)

    Article  CAS  Google Scholar 

  13. Lestel, L., Tkatchenko, I.: Chemical pathways: 150 years of evolution of chemistry. Eur. J. Org. Chem. 2007, 3064–3067 (2007)

    Article  Google Scholar 

  14. Lehn, J.-M.: Cryptates: the chemistry of macropolycyclic inclusion complexes. Acc. Chem. Res. 11, 49–57 (1978)

    Article  CAS  Google Scholar 

  15. Lehn, J.-M.: The concepts and language of supramolecular chemistry. In: Kisakürek, V. (ed.) Organic Chemistry: Its Language and Its State of the Art, pp 77–91. VCH, Weinheim (1993)

  16. Clayton, P., Davies, P.: The Re-Emergence of Emergence: the Emergentist Hypothesis from Science to Religion. Oxford University Press, Oxford (2006)

    Google Scholar 

  17. Nicolaou, K.C.: Joy of molecules. J. Org. Chem. 70, 7007–7016 (2005)

    Article  CAS  Google Scholar 

  18. Nicolaou, K.C.: Inspirations, discoveries, and future perspectives in total synthesis. J. Org. Chem. 74, 951–972 (2009)

    Article  CAS  Google Scholar 

  19. Woodward, R.B.: The total synthesis of vitamin B 12. Pure Appl. Chem. 33, 145–177 (1973)

    Article  CAS  Google Scholar 

  20. Newman, M.S., Lednider, D.: The synthesis and resolution of hexahelicene. J. Am. Chem. Soc. 78, 4765–4770 (1956)

    Article  CAS  Google Scholar 

  21. Hoffman, R.: Molecular beauty. Interdiscipl. Sci. Rev. 16, 301–311 (1991)

    Google Scholar 

  22. Wasserman, E.: The preparation of interlocking rings: a catenane. J. Am. Chem. Soc. 82, 4433–4434 (1960)

    Article  CAS  Google Scholar 

  23. Frisch, H.L., Wasserman, E.: Chemical topology. J. Am. Chem. Soc. 83, 3789–3795 (1961)

    Article  CAS  Google Scholar 

  24. Harrison, I.T., Harrison, S.: The synthesis of a stable complex of a macrocycle and a threaded chain. J. Am. Chem. Soc. 89, 5323–5324 (1967)

    Google Scholar 

  25. Flory, P.J.: Molecular size distribution in three dimensional polymers. i. gelation. J. Am. Chem. Soc. 63, 3082–3090 (1941)

    Google Scholar 

  26. Flory, P.J.: Molecular size distribution in three dimensional polymers. II. trifunctional branching units. J. Am. Chem. Soc. 63, 3091–3096 (1941)

    Article  CAS  Google Scholar 

  27. Flory, P.J.: Molecular size distribution in three dimensional polymers. III. Tetrafunctional branching units. J. Am. Chem. Soc. 63, 3096–3100 (1941)

    Article  CAS  Google Scholar 

  28. Sismour, A.M., Benner, S.A.: Synthetic biology. Expert Opin. Biol. 5, 1409–1414 (2005)

    Article  CAS  Google Scholar 

  29. Pedersen, C.J.: Cyclic polyethers and their complexes with metal salts. J. Am. Chem. Soc. 89, 2495–2496 (1967)

    Article  CAS  Google Scholar 

  30. Goodman, C.: Jean-Marie Lehn. Nat. Chem. Biol. 3, 685–688 (2007)

    Article  CAS  Google Scholar 

  31. Laszlo, P.: Playing with molecular models. Hyle 6, 85–97 (2000)

    Google Scholar 

  32. Cooke, H.: A historical study of structures for communication of organic chemistry information prior to 1950. Org. Biomol. Chem. 2, 3179–3191 (2004)

    Article  CAS  Google Scholar 

  33. Hoffman, R., Laszlo, P.: Representation in chemistry. Angew. Chem. Int. Ed. Engl. 30, 1–16 (1991)

    Article  Google Scholar 

  34. Goodwin, W.M.: Structural formulas and explanation in organic chemistry. Found. Chem. 10, 117–127 (2008)

    Article  CAS  Google Scholar 

  35. Cayley, A.: On the theory of the analytical forms called tree. Phil. Mag. XIII, 172–176 (1857)

    Google Scholar 

  36. Ourisson, G.: L’information moléculaire. Bull. Bibl. Fr. Paris 32(3) (1987)

  37. Ourisson, G.: Le langage universel de la chimie: les idéogrammes. L’actualité chimique, Janvier-Février (1986)

    Google Scholar 

  38. Arnheim, R.: Visual Thinking, p. 315. University of California Press, Berkeley (1969)

    Google Scholar 

  39. Jansen, M., Schön, J.C.: “Design” in chemical synthesis—an illusion? Angew. Chem. Int. Ed. 45, 3406–3472 (2006)

    Article  CAS  Google Scholar 

  40. Pauling, L., Corey, R.B.: A proposed structure for the nucleic acids. Proc. Natl. Acad. Sci. USA 39, 84–97 (1953)

    Article  CAS  Google Scholar 

  41. Watson, J.D., Crick, F.H.C.: A structure for deoxyribose nucleic acid. Nature 171, 737–738 (1953)

    Article  CAS  Google Scholar 

  42. Wilkins, A.R., Stokes, A.R., Wilson, H.R.: Molecular structure of deoxypentose nucleic acids. Nature 171, 738–740 (1953)

    Article  CAS  Google Scholar 

  43. Franklin, E.R., Gosling, R.G.: Molecular configuration in sodium thymonucleate. Nature 171, 740–741 (1953)

    Article  CAS  Google Scholar 

  44. Franklin, E.R., Gosling, R.G.: Evidence for 2-chain helix in crystalline structure of sodium deoxyribonucleate. Nature 172, 156–157 (1953)

    Article  CAS  Google Scholar 

  45. Watson, J.D., Crick, F.H.C.: Genetical Implications of the structure of deoxyribose nucleic acid. Nature 171, 964–967 (1953)

    Article  CAS  Google Scholar 

  46. Chargaff, E.: Chemical specificity of nucleic acids. Experientia VI, 201–209 (1950)

  47. Avery, O.T., MacLeod, C.M., McCarty, M.: Studies on the chemical transformation of pneumococcal types. J. Exp. Med. 79, 137–158 (1944)

    Article  CAS  Google Scholar 

  48. Hershey, A.D., Chase, M.: Independent functions of viral proteins and nucleic acid in growth of bacteriophage. J. Gen. Physiol. 36, 39–56 (1952)

    Article  CAS  Google Scholar 

  49. Eschenmoser, A.: One hundred years lock and key principle. Angew. Chem. Int. Ed. Engl. 33, 2363–2374 (1994)

    Article  Google Scholar 

  50. Lichtenthaler, F.W.: 100 years “schlüssel-schloss-prinzip”: what made emil fischer use this analogy? Angew. Chem. Int. Ed. Engl. 33, 2364–2373 (1994)

    Article  Google Scholar 

  51. Koshland Jr., D. E.: The key-lock theory and the induced fit theory. Angew. Chem. Int. Ed. Engl. 33, 2375–2378 (1994)

    Article  Google Scholar 

  52. Koshland, D.E. Jr.: Application of a theory of enzyme specificity to protein synthesis. Proc. Natl. Acad. Soc. USA 44, 98 (1958)

    Google Scholar 

  53. Williamson, R.J.: Induced fit in RNA-protein recognition. Nat. Struct. Biol. 7, 834–837 (2000)

    Article  CAS  Google Scholar 

  54. Rebek, J. Jr: Contorsions of encapsulated alkyl groups. Chem. Commun. 2777–2789 (2007)

  55. Rebek Jr, J.: Molecular behavior in small spaces. Acc. Chem. Res. 42, 1660–1668 (2009)

    Article  CAS  Google Scholar 

  56. Monod, J., Wyman, J., Changeux, J.-P.: On the nature of allosteric transitions: a plausible model. J. Mol. Biol. 12, 88–118 (1965)

    Article  CAS  Google Scholar 

  57. Changeux, J.-P., Edelstein, S.J.: Allosteric mechanisms of signal transduction. Science 308, 1424–1428 (2005); and references therein

    Google Scholar 

  58. Kovbasayuk, L., Krämer, R.: Allosteric supramolecular receptors and catalysts. Chem. Rev. 104, 3161–3188 (2004)

    Article  CAS  Google Scholar 

  59. Cohen, M.D., Schmidt, J.M.G.: Topochemistry. I. J. Chem. Soc. 1964, 1996–2000 (1964)

    Article  Google Scholar 

  60. Schmidt, J.M.G., et al: Published on Topochemical Reactions in the Crystal State from ‘Topochemistry. I.’ (1964) to ‘Topochemistry. Part XL’ (1978)

  61. de Gennes, P.G.: Nematic fluctuations and light scattering. C. R. Acad. Sci. Paris 266, 15–28 (1968)

    Google Scholar 

  62. Eigen, M., Schuster, P.: The hypercycle. Die Naturwissenschaften 64, 541–565 (1977)

    Article  CAS  Google Scholar 

  63. Kuhn, H.: Versuche zur Hernstellung einfacher organisieter System von Molekülen. Verhandlunghen der Schweirzerischen Natuforschenden Gesellschaft 1965, 245–266 (1965)

    Google Scholar 

  64. Kuhn, H.: Self-organization of molecular systems and evolution of the genetic apparatus. Angew. Chem. Int. Engl. 11, 798–820 (1972)

    Article  CAS  Google Scholar 

  65. Prigogine, I.: Nonequilibrium Statistical Mechanics. Wiley, New York (1962)

    Google Scholar 

  66. Thom, R.: Structural Stability and Morphogenesis. Benjamin, Reading, MA (1975)

    Google Scholar 

  67. Zeeman, C.: Catastrophe Theory. Addison-Wesley, Reading, MA (1977)

    Google Scholar 

  68. Thom, R.: Modèles mathématiques de la morphogenèse. 10/18. Union générale d’éditions, Paris (1974)

    Google Scholar 

  69. Mann, S.: The chemistry of form. Angew. Chem. Int. Ed. 39, 3392–3406 (2000)

    Article  CAS  Google Scholar 

  70. Cramer, F., Hettler, H.: Inclusion compounds of cyclodextrins. Die Naturwissenschaten 24, 625–632 (1967)

    Article  Google Scholar 

  71. Simmons, H.E., Park, C.H.: Macrobicyclic amines. I. Out-in isomerism of 1,(k + 2)-diazabicyclo(k.l.m)alkanes. J. Am. Chem. Soc. 90, 2428–2429 (1968)

    Article  CAS  Google Scholar 

  72. Park, C.H., Simmons, H.E.: Macrobicyclic amines. II. Out–out ↔in–in prototropy in 1,(k + 2)-diazabicyclo(k.l.m)alkaneammonium ions. J. Am. Chem. Soc. 90, 2429–2431 (1968)

    Article  CAS  Google Scholar 

  73. Park, C.H., Simmons, H.E.: Macrobicyclic amines. III. Encapsulation of halide ions by in–in 1,(k + 2)-diazabicyclo(k.l.m)alkaneammonium ions. J. Am. Chem. Soc. 90, 2431–2432 (1968)

    Article  CAS  Google Scholar 

  74. Lehn, J.-M.: Toward complex matter: supramolecular chemistry and self-organisation. Proc. Natl. Acad. Sci. USA 99, 4763–4768 (2002)

    Article  CAS  Google Scholar 

  75. Lehn, J.-M.: Supramolecular chemistry: from molecular information towards self-organisation and complex matter. Rep. Prog. Phys. 67, 249–265 (2004)

    Article  Google Scholar 

  76. Gale, P.: Supramolecular chemistry: from complexes to complexity. Phil. Trans.: Math. Phys. Eng. Sci. 358, 432–453 (2000)

    Google Scholar 

  77. Douglas, S.M., Dietz, H., Liedl, T., Hogberg, B., Graf, F., Shih, W.M.: Self-assembly of DNA into nanoscale three-dimensional shapes. Nature 459, 414–418 (2009)

    Article  CAS  Google Scholar 

  78. Nykypanchuk, D., Maye, M.M., van der Lelie, D., Gang, O.: DNA-guided crystallization of colloidal nanoparticles. Nature 451, 549–552 (2008)

    Article  CAS  Google Scholar 

  79. Park, S.Y., Lytton-Jean, A.K., Lee, B., Weigand, S., Schatz, G.C., Mirkin, C.A.: DNA-programmable nanoparticle crystallization. Nature 451, 553–556 (2008)

    Article  CAS  Google Scholar 

  80. Seeman, N.C.: DNA in a material world. Nature 421, 427–431 (2003)

    Article  CAS  Google Scholar 

  81. Vincent, A., Suresh, B., Heckert, E., Dowding, J., Hirst, S.M., Inerbaev, T.M., Self, W.T., Reilly, C.M., Masunov, A.E., Rahman, T.S., Seal, S.: Protonated nano particle surface governing ligand tethering and cellular targeting. ACS Nano. doi:10.1021/nn9000148

  82. Müller, A.: Bringing inorganic chemistry to life. Chem. Commun. 2003, 803–806 (2003)

    Article  CAS  Google Scholar 

  83. Jones, S.: Emergence and organisation. Towards a taxonomy of organizing relations. NeuroQuantology 3, 219–236 (2004)

    Google Scholar 

  84. Halley, J.D., Winkler, D.A.: Classification of emergence and its relation to self-organization. Complexity 14, 10–15 (2008)

    Article  Google Scholar 

  85. Corning, P.A.: The re-emergence of “emergence”: a venerable concept in search of a theory. Complexity 7, 18–30 (2002)

    Article  Google Scholar 

  86. Harned, R.L., Hidy, P.H., Corum, C.J., Jones, K.I.: Nigericine, a new crystalline antibiotic from an “unidentified streptomicine”. Antibiot Chemoter. 1, 496–594 (1951)

    Google Scholar 

  87. Van Beylen, M., Bywater, S., Smets, G., Szwarc, M., Worsfold, D.J.: Developments in Anionic Polymerization. Advances in Polymer Science, vol. 86. Springer, Berlin/Heidelberg (1988)

    Google Scholar 

  88. Ryan, A.J.: Emergence is coupled to scope, not level. Complexity 13, 67–77 (2007)

    Article  Google Scholar 

  89. Ludlow, R.F., Otto, S.: Systems chemistry. Chem. Soc. Rev. 37, 101–108 (2008)

    Article  CAS  Google Scholar 

  90. Desiraju, G.R.: The middle kingdom. Curr. Sci. 88, 374–380 (2005)

    Google Scholar 

  91. Whitesides, G.M., Ismagilov, R.F.: Complexity in chemistry. Science 284, 89–92 (1999)

    Article  CAS  Google Scholar 

  92. Ball, P.: What chemists want to know. Nature 442, 500–502 (2006)

    Article  CAS  Google Scholar 

  93. Wächtershäuser, G.: The origin of life and its methodological challenge. J. Theor. Biol. 187, 483–494 (1997)

    Article  Google Scholar 

  94. Löb, W.: Ueber das Verhalten des Formamids undter der Wirking der stillen Entlandung. Ein Beitrag zur Frage Sticktoff-Assimilation. Berichte der deutschen chemischen gesellschaft 46, 684–697 (1913)

    Article  Google Scholar 

  95. Miescher, F.: Ueber die Chemische Zusammensetzung der Eiterzellen. Medecinische-chemische Untersuchungen 4, 441–460 (1871)

    Google Scholar 

  96. Mendelejeff, D.: Ueber die Heglenchungen der Eigenschaffen zur den Atomweglohten der Elemente. Zeitsch. Chem. 12, 405–406 (1869)

    Google Scholar 

  97. Pross, A.: Seeking the chemical roots of Darwinism: Bridging between chemistry and biology. Chem. Eur. J. 15, 8374–8381 (2009)

    Article  CAS  Google Scholar 

  98. Voytek, S.B., Joyce, G.F.: Niche partitioning in the coevolution of 2 distinct RNA enzymes. Proc. Natl. Acad. Sci. USA 106, 7780–7785 (2009)

    Article  CAS  Google Scholar 

  99. Pioda, L.A.P., Wachater, H.A., Dohner, R.E., Simon, W.: Helv. Chim. Acta 50, 1373–1381 (1967)

    Article  CAS  Google Scholar 

  100. Wipf, H.K., Pioda, L.A.P., Sefara, Z., Simon, W.: Helv. Chim. Acta 51, 377–391 (1968)

    Article  CAS  Google Scholar 

  101. Steed, J.W., Atwood, J.L.: Supramolecular Chemistry, 2nd edn. John Wiley & Sons, United Kingdom (2009)

    Book  Google Scholar 

  102. Graham Cairns-Smith, A.: Seven Clues to the Origin of Life: A Scientific Detective Story. Cambridge University Press, Cambridge, UK (1985)

    Google Scholar 

  103. Cleland, C.E., Chyba, C.F.: Defining ‘life’. Origins Life Evol. Biosph. 31, 387–393 (2002)

    Article  Google Scholar 

  104. Szostak, J.W., Bartel, D.P., Luisi, P.L.: Synthesizing life. Nature 409, 387–390 (2001)

    Article  CAS  Google Scholar 

  105. Joyce, G.F., Orgel, L.E.: Progress toward understanding the origin of the RNA world. In: Gesteland, R.F., Cech, T.R., Atkins, J.F. (eds.) Progress toward Understanding the Origin of the RNA World, in The RNA World, 3rd edn. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York (2006)

    Google Scholar 

  106. Zhu, T.F., Szostak, J.W.: Coupled growth and division of model protocell membranes. J. Am. Chem. Soc. 131, 5705–5713 (2009)

    Article  CAS  Google Scholar 

  107. McCray, W.P.: Will small be beautiful? Making policies for our nanotech future. Hist. Technol. 21, 177–203 (2005)

    Article  Google Scholar 

  108. Taniguchi, N.: On the basic concept of ‘nano-technology’. In: Proc. Intl. Conf. Prod. Eng. Tokyo, Part II. Japan Society of Precision Engineering (1974)

  109. Bronowski, J.: Science and Human Values. Harper & Row, New York (1956)

  110. Simonson, M., Schlosser, C.: The DaVinci Project: Multimedia in Art and Chemistry. TechTrends, April/May 1998, pp 19–22 (1998)

  111. Balzani, V., Credi, A., Venturi, M.: The bottom-up approach to molecular-level devices and machines. Chem. Eur. J. 8, 5525–5532 (2002)

    Google Scholar 

  112. Foucault, M.: in L’ordre du discours. Leçon inaugurale. Collège de France (1971)

  113. Fabbrizzi, L.: Communicating about matter with symbols: evolving from alchemy to chemistry. J. Chem. Ed. 85, 1501–1511 (2008)

    Article  CAS  Google Scholar 

  114. Kubbinga, H.H.: Les premières théories «moléculaires»: Isaac Beeckman (1620) et Sébastien Basson (1621). Le concept d’ «individu substantiel» et d’ «espèce substantielle». Rev. Hist. Sci. XXXVII(3–4), 215–233 (1984)

    Google Scholar 

  115. Mezeau, L.: Mémoire. Médium No 9 (2007)

  116. Goody, J.: La Raison Graphique. Editions de Minuit (1979)

  117. Babaev, E.V., Hefferlin, R.: The concepts of periodicity and hyperperiodicity: from atoms to molecules. In: Rouvray, D. H. (ed.) Concepts in Chemistry—a Contemporary Challenge, Chap. 3. John Wiley & Sons Inc., New York, USA (1997)

  118. Armstrong, R.W., Beau, J.-M., Cheon, S.H., Christ, W.J., Fujioka, H., Ham, W.H., Hawkins, L.D., Jin, H., Kang, S.H., Kishi, Y., Martinelli, M.J., McWhorter, W.W.J., Mizuno, M., Nakata, M., Stutz, A.E., Talamas, M., Tino, J.A., Ueda, K., Uenishi, J.-I., White, J.B., Yonaga, M.: Total synthesis of a fully protected palytoxin carboxylic acid. J. Am. Chem. Soc. 111, 7525–7530 (1989)

    Article  CAS  Google Scholar 

  119. Scerri, E.: Explaining the periodic table, and the role of chemical triads. Found. Chem. doi:10.1007/s10698-010-9082-9

  120. Parr, R.G., Ayers, P.W., Nalewajski, R.F.: What is an atom in a molecule? J. Phys. Chem. A 109, 3957–3959 (2005)

    Article  CAS  Google Scholar 

  121. Toumey, C.: Truth and beauty at the nanoscale. Leonardo 42, 151–155 (2009)

    Article  Google Scholar 

  122. Toumey, C.: Cubism at the nanoscale. Nat. Nanotechnol. 2, 587–589 (2007)

    Article  CAS  Google Scholar 

  123. Spector, T.M., Schummer, J.: Aesthetics and visualisation in chemistry. HYLE-Int. J. Philos. Chem. 9, 3–118 (2003)

    Google Scholar 

  124. Kay, E.R., Leigh, D.A., Zerbetto, F.: Synthetic molecular motors and mechanical machines. Angew. Chem. Int. Ed. 46, 72–191 (2007)

    Article  CAS  Google Scholar 

  125. Newkome, G.R., Moorefield, C.N., Vögtle, F.: Dendritic Molecules. VCH, Weinheim, Germany (1996)

  126. Kermack, W.O., Robinson, R.: An explanation of the properties of induced polarity of atoms and an interpretation of the theory of partial valencies on an electronic basis. J. Chem. Soc. 121, 427–440 (1922)

    CAS  Google Scholar 

  127. Bram, G., Anh, N.T.: The difficult marriage of theory and French organic chemistry in the 20th century. J. Mol. Struct. (TheoChem) 424, 201–206 (1998)

    Article  CAS  Google Scholar 

  128. Hanson, R. M.: Molecular Origami—Precision Scale Models. University Science Books, Sausalito (1995)

  129. Bragg, W.H.: The structure of organic crystals. Proc. R. Soc. Lond. 34, 33–75 (1921)

    CAS  Google Scholar 

  130. Cayley, A.: On the analytical forms called trees, with applications to the theory of chemical combinations. Rep. Brit. Ass. 257–305 (1875)

  131. Gutman, I., Vidovic, D., Popovic, L.: Graph representation of organic molecules. Cayley’s plerograms vs. his kenograms. J. Chem. Soc. Faraday Trans. 94, 857–860 (1998)

    Article  CAS  Google Scholar 

  132. Rocke, A.J.: Image and Reality: Kékulé, Kopp, and the Scientific Imagination. University of Chicago Press,   (2010)

    Google Scholar 

  133. Pauling, L.: Molecular architecture and biological reactions. Chem. Eng. News 24, 1375–1377 (1946)

    Article  CAS  Google Scholar 

  134. Kitaigorodskii, A.I.: Organic Chemical Crystallography. Consultants Bureau, New York (1961)

    Google Scholar 

  135. Kitaigorodskii, A.I.: Molecular Crystals and Molecules. Academic Press, New York (1973)

    Google Scholar 

  136. Dunitz, J.D., Gavezzotti, A.: Molecular recognition in organic crystals: directed intermolecular bonds or nonlocalized bonding? Angew. Chem. Int. Ed. 44, 1766–1787 (2005)

    Article  CAS  Google Scholar 

  137. Bonchev, D., Seitz, W.A.: The concept of complexity in chemistry. Rouvray, D. H. (ed.) In: Concepts in Chemistry, Chap. 11. John Wiley & Sons Inc., New York (1997)

  138. Moore, T.S., Winmill, T.F.: The state of maines in aqueous solution. J. Chem. Soc. 101, 1635–1676 (1912)

    CAS  Google Scholar 

  139. Latimer, W.M., Rodebush, W.H.: Polarity and inonization from standpoint of the Lewis theory of valence. J. Am. Chem. Soc. 42, 1419–1433 (1920)

    Article  CAS  Google Scholar 

  140. Shaik, S.: Is my chemical universe localized or delocalized? Is there a future for chemical concepts? New J. Chem. 31, 2015–2028 (2007)

    Article  CAS  Google Scholar 

  141. Childs, P.: Chimlingo. Educ. Chem. 30, 65 (1993)

    CAS  Google Scholar 

  142. Stang, P.: Molecular architecture: coordination as the motif in the rational design and assembly of discrete supramolecular species—self-assembly of metallacyclic polygons and polyhedra. Chem. Eur. J. 4, 19–27 (1998)

    Article  CAS  Google Scholar 

  143. Breault, G.A., Hunter, C.A., Mayers, P.C.: Supramolecular topology. Tetrahedron 55, 5265–5293

  144. Amijs, C.H.M., van Klink, G.P.M., van Koten, G.: Dalton Trans. 308–327 (2006)

  145. Clegg, J.K., Iremonger, S.S., Hayter, M.J., Southon, P.D., Macquart, R.B., Duriska, M.B., Jensen, P., Turner, P., Jolliffe, K.A., Kepert, C.J., Meehan, G.V., Lindoy, L.F.: Angew. Chem. Int. Ed. 49, 1075–1078 (2010)

    Article  CAS  Google Scholar 

  146. Rocha, L.M.: Selected self-organization. In: Salthe, S., van de Vijver, G., Delpos, M. (eds.) Evolutionary Systems: Biological and Epistemological Perspectives on Selection and Self-Organization, pp 341–358. Kluwer Academic Publishers, Dordrecht (1998)

  147. Nitschke, J.R., Lehn, J.-M.: Self-organization by selection: generation of a metallosupramolecular grid architecture by selection of components in a dynamic library of ligands. Proc. Natl. Acad. Sci. USA 100, 11970–11974 (2003)

    Article  CAS  Google Scholar 

  148. Schwann, T., Schleyden, M.J.: Microscopical Investigations into the Accordance in the Structure and Growth of Plants and Animals. Sydenham Society, London (1847)

  149. de Duve, C.: Chemistry and selection. Chem. Biodivers. 4, 574–583 (2007)

    Article  Google Scholar 

  150. Zhang, B., Wepf, R., Fischer, K., Schmidt, M., Besse, S., Lindner, P., King, B.T., Sigel, R., Schurtenberger, P., Talmon, Y., Ding, Y., Krüger, M., Halperin, A., Schlüter, A.D.: The largest synthetic structure with molecular precision: towards a molecular object. Angew. Chem. Int. Ed. 50, 737–740 (2011)

    Article  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jacques Vicens.

Additional information

This is an extended version of a previous article: Vicens, J., Vicens, Q.: Origins and emergences of supramolecular chemistry. J. Incl. Phenom. Macrocycl. Chem. 65, 221–235 (2009).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Vicens, J., Vicens, Q. Emergences of supramolecular chemistry: from supramolecular chemistry to supramolecular science. J Incl Phenom Macrocycl Chem 71, 251–274 (2011). https://doi.org/10.1007/s10847-011-0001-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10847-011-0001-z

Keywords

Navigation