请输入您要查询的百科知识:

 

词条 Cyclic compound
释义

  1. Structural introduction

  2. Nomenclature

  3. Carbocycles

  4. Inorganic cyclic compounds

  5. Heterocyclic compounds

  6. Aromaticity

  7. Simple, mono-cyclic examples

  8. Stereochemistry

  9. Conformational isomerism

  10. Macrocycles

  11. Principle uses of cyclic structures

  12. Complex and polycyclic examples

  13. Synthetic reactions altering rings

     Important general reactions for forming rings   Ring-closing reactions  Ring-opening reactions   Ring expansion and ring contraction reactions 

  14. See also

  15. References

  16. Further reading

  17. External links

{{Short description|Molecule with a ring of bonded atoms}}

A cyclic compound (ring compound) is a term for a compound in the field of chemistry in which one or more series of atoms in the compound is connected to form a ring. Rings may vary in size from three to many atoms, and include examples where all the atoms are carbon (i.e., are carbocycles), none of the atoms are carbon (inorganic cyclic compounds), or where both carbon and non-carbon atoms are present (heterocyclic compounds). Depending on the ring size, the bond order of the individual links between ring atoms, and their arrangements within the rings, carbocyclic and heterocyclic compounds may be aromatic or non-aromatic, in the latter case, they may vary from being fully saturated to having varying numbers of multiple bonds between the ring atoms. Because of the tremendous diversity allowed, in combination, by the valences of common atoms and their ability to form rings, the number of possible cyclic structures, even of small size (e.g., <17 total atoms) numbers in the many billions.

Adding to their complexity and number, closing of atoms into rings may lock particular atoms with distinct substitution (by functional groups) such that stereochemistry and chirality of the compound results, including some manifestations that are unique to rings (e.g., configurational isomers). As well, depending on ring size, the three-dimensional shapes of particular cyclic structures—typically rings of 5-atoms and larger—can vary and interconvert such that conformational isomerism is displayed. Indeed, the development of this important chemical concept arose, historically, in reference to cyclic compounds. Finally, cyclic compounds, because of the unique shapes, reactivities, properties, and bioactivities that they engender, are the largest majority of all molecules involved in the biochemistry, structure, and function of living organisms, and in the man-made molecules (e.g., drugs, herbicides, etc.).

Structural introduction

A cyclic compound or ring compound is a compound at least some of whose atoms are connected to form a ring.[1]{{rp|unknown}}{{page needed|date=June 2015}} Rings vary in size from 3 to many tens or even hundreds of atoms. Examples of ring compounds readily include cases where:

  • all the atoms are carbon (i.e., are carbocycles),
  • none of the atoms are carbon (inorganic cyclic compounds),[2] or where
  • both carbon and non-carbon atoms are present (heterocyclic compounds).

Common atoms can (as a result of their valences) form varying numbers of bonds, and many common atoms readily form rings. In addition, depending on the ring size, the bond order of the individual links between ring atoms, and their arrangements within the rings, cyclic compounds may be aromatic or non-aromatic; in the case of non-aromatic cyclic compounds, they may vary from being fully saturated to having varying numbers of multiple bonds. As a consequence of the constitutional variability that is thermodynamically possible in cyclic structures, the number of possible cyclic structures, even of small size (e.g., <17 atoms) numbers in the many billions.[3]

Moreover, the closing of atoms into rings may lock particular functional group–substituted atoms into place, resulting in stereochemistry and chirality being associated with the compound, including some manifestations that are unique to rings (e.g., configurational isomers);[4] As well, depending on ring size, the three-dimensional shapes of particular cyclic structures—typically rings of 5-atoms and larger—can vary and interconvert such that conformational isomerism is displayed.[4]

{{refimprove science|date=April 2015}}{{Expert needed|Chemistry|talk=|reason=narrow and therefore unscholarly content for this fundamental article, which at least needs, i, redaction of unimportant material (lists, etc.), ii, sourcing and iii broadening of remaining content to begin to move toward an encyclopedic article|date=April 2015}}

Nomenclature

{{expand section|date=April 2015}}

IUPAC nomenclature has extensive rules to cover the naming of cyclic structures, both as core structures, and as substituents appended to alicyclic structures.{{vague|date=April 2015}}{{citation needed|date=April 2015}} The term macrocycle is used when a ring-containing compound has a ring of 8 or more atoms.[5][6] The term polycyclic is used when more than one ring appears in a single molecule.{{citation needed|date=April 2015}} Naphthalene is formally a polycyclic compound, but is more specifically named as a bicyclic compound. Several examples of macrocyclic and polycyclic structures are given in the final gallery below.

{{anchor|annular atom}}

The atoms that are part of the ring structure are called annular atoms.[7]

Carbocycles

{{expand section|date=April 2015}}

The vast majority of cyclic compounds are organic, and of these, a significant and conceptually important portion are composed of rings made only of carbon atoms (i.e., they are carbocycles).{{citation needed|date=April 2015}}

Inorganic cyclic compounds

{{expand section|date=April 2015}}

Inorganic atoms form cyclic compounds as well. Examples include sulfur, silicon (e.g., in silanes), phosphorus (e.g., in phosphanes and phosphoric acid variants), and boron (e.g., in triboric acid).{{citation needed|date=April 2015}} When carbon in benzene is "replaced" by other elements, e.g., as in borabenzene, silabenzene, germanabenzene, stannabenzene, and phosphorine, aromaticity is retained, and so aromatic inorganic cyclic compounds are known and well-characterized.{{citation needed|date=April 2015}}

Heterocyclic compounds

{{expand section|date=April 2015}}

Cyclic compounds that have both carbon and non-carbon atoms present are termed (heterocyclic compounds);{{citation needed|date=April 2015}} alternatively the name can refer to inorganic cyclic compounds, such as siloxanes and borazines, that have more than one type of atom in their rings.{{citation needed|date=April 2015}} Hantzsch–Widman nomenclature is recommended by the IUPAC for naming heterocycles, but many common names remain in regular use.{{citation needed|date=April 2015}}

Aromaticity

{{expand section|date=April 2015}}

Cyclic compounds may or may not exhibit aromaticity; benzene is an example of an aromatic cyclic compound, while cyclohexane is non-aromatic. In organic chemistry, the term aromaticity is used to describe a cyclic (ring-shaped), planar (flat) molecule that exhibits unusual stability as compared to other geometric or connective arrangements of the same set of atoms. As a result of their stability, it is very difficult to cause aromatic molecules to break apart and to react with other substances. Organic compounds that are not aromatic are classified as aliphatic compounds—they might be cyclic, but only aromatic rings have especial stability (low reactivity).

Since one of the most commonly encountered aromatic systems of compounds in organic chemistry is based on derivatives of the prototypical aromatic compound benzene (an aromatic hydrocarbon common in petroleum and its distillates), the word “aromatic” is occasionally used to refer informally to benzene derivatives, and this is how it was first defined. Nevertheless, many non-benzene aromatic compounds exist. In living organisms, for example, the most common aromatic rings are the double-ringed bases in RNA and DNA. A functional group or other substituent that is aromatic is called an aryl group.

The earliest use of the term “aromatic” was in an article by August Wilhelm Hofmann in 1855.[1] Hofmann used the term for a class of benzene compounds, many of which do have odors (aromas), unlike pure saturated hydrocarbons. Today, there is no general relationship between aromaticity as a chemical property and the olfactory properties of such compounds (how they smell), although in 1855, before the structure of benzene or organic compounds was understood, chemists like Hofmann were beginning to understand that odiferous molecules from plants, such as terpenes, had chemical properties we recognize today are similar to unsaturated petroleum hydrocarbons like benzene.

In terms of the electronic nature of the molecule, aromaticity describes a conjugated system often made of alternating single and double bonds in a ring. This configuration allows for the electrons in the molecule’s pi system to be delocalized around the ring, increasing the molecule's stability. The molecule cannot be represented by one structure, but rather a resonance hybrid of different structures, such as with the two resonance structures of benzene. These molecules cannot be found in either one of these representations, with the longer single bonds in one location and the shorter double bond in another (See Theory below). Rather, the molecule exhibits bond lengths in between those of single and double bonds. This commonly seen model of aromatic rings, namely the idea that benzene was formed from a six-membered carbon ring with alternating single and double bonds (cyclohexatriene), was developed by August Kekulé (see History section below). The model for benzene consists of two resonance forms, which corresponds to the double and single bonds superimposing to produce six one-and-a-half bonds. Benzene is a more stable molecule than would be expected without accounting for charge delocalization.{{citation needed|date=April 2015}}

Simple, mono-cyclic examples

The following are examples of simple and aromatic carbocycles, inorganic cyclic compounds, and heterocycles:

Stereochemistry

{{expand section|date=April 2015}}

The closing of atoms into rings may lock particular atoms with distinct substitution by functional groups such that the result is stereochemistry and chirality of the compound, including some manifestations that are unique to rings (e.g., configurational isomers).[4]

Conformational isomerism

{{expand section|a good, sourced, conceptual introduction to this important, universally taught organic chemistry topic, suitable to an article about many different rings|small=no==date=April 2015|date=June 2015}}{{one source|section|date=June 2015}}{{multiple image
| width = 300
| footer = General description. The structures are shown in line angle representation, though in the image at left, the lines projecting from the cyclohexane are not terminal methyl groups; rather, they indicate possible positions that might be occupied by substituents (functional groups) attached to the ring. In the image at left, those groups projecting upward and downward are termed axial substituents (a), and those groups projecting around the conceptual equator are termed equatorial substituents (e). Note, in general, the axial substituents are closer in space to one another (allowing for repulsive interactions); moreover, in the boat form, axial substituents in directly opposing positions (12 o'clock and 6 o'clock, termed "1,4-") are very close in space, and therefore give rise to even greater repulsion. These and other types of strain are used to explain the observation that the chair conformation of cyclohexanes is the favored conformation.[4]
| image1 = Chair-Boat-Conformation_general.svg
| alt1 = To be supplied
| caption1 = Chair and boat conformers in cyclohexanes. Two conformers of cyclohexane, the chair at left, and the boat at right (in German, respectively, Sessel and Wanne, the latter meaning "bath").
| image2 = Cis14dimethyl cyclohexane2 HD.jpg
| alt2 = To be supplied
| caption2 = cis-1,4-Dimethylcyclohexane, in chair form, minimising steric interactions between the methyl groups in the directly opposing 1,4-positions of the cyclohexane ring.
}}

Depending on ring size, the three-dimensional shapes of particular cyclic structures—typically rings of 5-atoms and larger—can vary and interconvert such that conformational isomerism is displayed.[4] Indeed, the development of this important chemical concept arose, historically, in reference to cyclic compounds. For instance, cyclohexanes—six membered carbocycles with no double bonds, to which various substituents might be attached, see image—display an equilibrium between two conformations, the chair and the boat, as shown in the image.

The chair conformation is the favored configuration, because in this conformation, the steric strain, eclipsing strain, and angle strain that are otherwise possible are minimized.[4] Which of the possible chair conformations predominate in cyclohexanes bearing one or more substituents depends on the substiuents, and where they are located on the ring; generally, "bulky" substituents—those groups with large volumes, or groups that are otherwise repulsive in their interactions{{citation needed|date=June 2015}}—prefer to occupy an equatorial location.[4] An example of interactions within a molecule that would lead to steric strain, leading to a shift in equilibrium from boat to chair, is the interaction between the two methyl groups in cis-1,4-dimethylcyclohexane. In this molecule, the two methyl groups are in opposing positions of the ring (1,4-), and their cis stereochemistry projects both of these groups toward the same side of the ring. Hence, if forced into the higher energy boat form, these methyl groups are in steric contact, repel one another, and drive the equilibrium toward the chair conformation.{{citation needed|date=June 2015}}

Macrocycles

{{expand section|date=April 2015}}{{refimprove science|section|date=April 2015}}

The term macrocycle is used for compounds having a rings of 8 or more atoms.[5][6] Macrocycles may be fully carbocyclic, heterocyclic but having limited heteroatoms (e.g., in lactones and lactams), or be rich in heteroatoms and displaying significant symmetry (e.g., in the case of chelating macrocycles). Macrocycles can access a number of stable conformations, with preference to reside in conformations that minimize transannular nonbonded interactions within the ring (e.g., with the chair and chair-boat being more stable than the boat-boat conformation for cyclooctane, because of the interactions depicted by the arcs shown).{{citation needed|date=April 2015}} Medium rings (8-11 atoms) are the most strained, with between 9-13 (kcal/mol) strain energy, and analysis of factors important in the conformations of larger macrocycles can be modeled using medium ring conformations.[8]{{page needed|date=April 2015}} Conformational analysis of odd-membered rings suggests they tend to reside in less symmetrical forms with smaller energy differences between stable conformations.[9]{{primary source inline|date=April 2015}}

Principle uses of cyclic structures

{{expand section|date=April 2015}}

Because of the unique shapes, reactivities, properties, and bioactivities that they engender, cyclic compounds are the largest majority of all molecules involved in the biochemistry, structure, and function of living organisms, and in the man-made molecules (e.g., drugs, herbicides, etc.) through which man attempts to exert control over nature and biological systems.

Complex and polycyclic examples

The following are examples of cyclic compounds exhibiting more complex ring systems and stereochemical features:

Synthetic reactions altering rings

{{Multiple issues|section=yes|{{unreferenced section|date=April 2015}}{{expand section|1=with general, sourced introductory text to move this from its list-only format|small=no|date=April 2015}}
}}

Important general reactions for forming rings

{{expand section|1=further scholarly and comprehensive list of such reactions as described|small=no|date=April 2015}}

There are a variety of specialized reactions whose use is solely the formation of rings, and these will be discussed below. In addition to those, there are a wide variety of general organic reactions that historically have been crucial in the development, first, of understanding the concepts of ring chemistry, and second, of reliable procedures for preparing ring structures in high yield, and with defined orientation of ring substituents (i.e., defined stereochemistry). These general reactions include:

  • Acyloin condensation;
  • Anodic oxidations; and
  • the Dieckmann condensation as applied to ring formation.

Ring-closing reactions

{{expand section|1=with sourced, scholarly list only of most important examples, referring reader to a source with minor additional examples|small=no|date=April 2015}}

In organic chemistry, a variety of synthetic procures are particularly useful in closing carbocyclic and other rings; these are termed ring-closing reactions. Examples include:

  • alkyne trimerisation;
  • the Bergman cyclization;
  • the Diels-Alder and other cycloaddition reactions;
  • the Nazarov cyclization reaction;
  • various radical cyclizations;
  • ring-closing metathesis reactions, which can be also be used to accomplish a specific type of polymerization;
  • the Ruzicka large ring synthesis; and
  • the Wenker synthesis of aziridines.

Ring-opening reactions

{{expand section|1=with comparably detailed, sourced, scholarly list only of most important examples, again referring reader to source for minor additional examples|small=no|date=April 2015}}

A variety of further synthetic procedures are particularly useful in opening carbocyclic and other rings, generally which contain a double bound or other functional group "handle" to facilitate chemistry; these are termed ring-opening reactions. Examples include:

  • ring opening metathesis, which can be also be used to accomplish a specific type of polymerization.

Ring expansion and ring contraction reactions

{{Main|Ring expansion and ring contraction}}{{empty section|date=April 2015}}

See also

  • Effective molarity
  • Heterocyclic compound
  • Lactone
  • Macrocycle
  • Open-chain compound

References

1. ^{{JerryMarch}}.{{page needed|date=June 2015}}
2. ^{{cite journal |doi=10.1007/BF01141802 |title=Classification of inorganic cyclic compounds |journal=Journal of Structural Chemistry |volume=2 |issue=3 |pages=350–8 |year=1961 |last1=Halduc |first1=I. }}
3. ^{{cite journal |doi=10.1021/ar500432k |pmid=25687211 |title=The Chemical Space Project |journal=Accounts of Chemical Research |volume=48 |issue=3 |pages=722–30 |year=2015 |last1=Reymond |first1=Jean-Louis }}
4. ^William Reusch, 2010, "Stereoisomers Part I," In Virtual Textbook of Organic Chemistry, Michigan State University, see  , accessed 7 April 2015.
5. ^{{cite journal |doi=10.1016/S0040-4020(01)93273-9 |title=Chemical consequences of conformation in macrocyclic compounds |journal=Tetrahedron |volume=37 |issue=23 |pages=3981–96 |year=1981 |last1=Still |first1=W.Clark |last2=Galynker |first2=Igor }}
6. ^J. D. Dunitz. Perspectives in Structural Chemistry (Edited by J. D. Dunitz and J. A. Ibers), Vol. 2, pp. l-70; Wiley, New York (1968)
7. ^{{cite book|last1=Morris|first1=Christopher G.|last2=Press|first2=Academic|title=Academic Press Dictionary of Science and Technology|publisher=Gulf Professional Publishing|isbn=9780122004001|page=120|url=https://books.google.com.au/books?id=nauWlPTBcjIC&pg=PA120|language=en|year=1992}}
8. ^Eliel, E.L., Wilen, S.H. and Mander, L.S. (1994) Stereochemistry of Organic Compounds, John Wiley and Sons, Inc., New York.{{page needed|date=April 2015}}
9. ^{{cite journal |doi=10.1016/S0040-4020(01)90685-4 |title=Conformational analysis of medium-ring ketones |journal=Tetrahedron |volume=30 |issue=12 |pages=1629–37 |year=1974 |last1=Anet |first1=F.A.L. |last2=St. Jacques |first2=M. |last3=Henrichs |first3=P.M. |last4=Cheng |first4=A.K. |last5=Krane |first5=J. |last6=Wong |first6=L. }}{{primary source inline|date=April 2015}}
10. ^{{cite journal |doi=10.1006/jmbi.2001.5077 |pmid=11700061 |title=Refined structure of αβ-tubulin at 3.5 Å resolution |journal=Journal of Molecular Biology |volume=313 |issue=5 |pages=1045–57 |year=2001 |last1=Löwe |first1=J |last2=Li |first2=H |last3=Downing |first3=K.H |last4=Nogales |first4=E }}

Further reading

  • Jürgen-Hinrich Fuhrhop & Gustav Penzlin, 1986, "Organic synthesis: concepts, methods, starting materials," Weinheim, BW, DEU:VCH, {{ISBN|0895732467}}, see [https://books.google.com/books?id=lJbgAAAAMAAJ], accessed 19 June 2015.
  • Michael B. Smith & Jerry March, 2007, "March's Advanced Organic Chemistry: Reactions, Mechanisms, and Structure," 6th Ed., New York, NY, USA:Wiley & Sons, {{ISBN|0470084944}}, see [https://books.google.com/books?isbn=0470084944], accessed 19 June 2015.
  • Francis A. Carey & Richard J. Sundberg, 2006, "Title Advanced Organic Chemistry: Part A: Structure and Mechanisms," 4th Edn., New York, NY, USA:Springer Science & Business Media, {{ISBN|0306468565}}, see [https://books.google.com/books?isbn=0306468565], accessed 19 June 2015.
  • Michael B. Smith, 2011, "Organic Chemistry: An Acid—Base Approach," Boca Raton, FL, USA:CRC Press, {{ISBN|1420079212}}, see [https://books.google.com/books?isbn=1420079212], accessed 19 June 2015. [May not be most necessary material for this article, but significant content here is available online.]
  • Jonathan Clayden, Nick Greeves & Stuart Warren, 2012, "Organic Chemistry," Oxford, Oxon, GBR:Oxford University Press, {{ISBN|0199270295}}, see [https://books.google.com/books?isbn=0199270295], accessed 19 June 2015.
  • László Kürti & Barbara Czakó, 2005, "Strategic Applications of Named Reactions in Organic Synthesis: Background and Detailed Mechanisms, Amsterdam, NH, NLD:Elsevier Academic Press, 2005ISBN 0124297854, see [https://books.google.com/books?isbn=0124297854], accessed 19 June 2015.

External links

  • {{MeshName|Polycyclic+Compounds}}
  • {{MeshName|Macrocyclic+Compounds}}
{{Authority control}}

2 : Molecular geometry|Cyclic compounds

随便看

 

开放百科全书收录14589846条英语、德语、日语等多语种百科知识,基本涵盖了大多数领域的百科知识,是一部内容自由、开放的电子版国际百科全书。

 

Copyright © 2023 OENC.NET All Rights Reserved
京ICP备2021023879号 更新时间:2024/9/22 19:33:03