20
Computers Math. Applic. Vol. 17, No. 1-3, pp. 397-416, 1989 0097-4943/89 $3.00+ 0.00 Printed in Great Britain. All rights reserved Copyright © 1989 Pergamon Press plc CARBON AND ITS NETS A. T. BALABAN Polytechnic Institute, Department of Organic Chemistry, Splaiul Independenlei 313, 77206 Bucharest, Romania Abstract--The uniqueness of C--C bonds, which allows life to exist and gives birth to the infinite number of organic compounds, is analysed. A brief survey of the two known C allotropes, graphite and diamond, stresses how the different structure of the nets leads to such enormous differences in properties. The synthesis of diamond is surveyed. Other possible, hypothetical forms of elemental carbon are math- ematically surveyed depending on the hybridization state of the C atoms, i.e. on the vertex degrees in the infinite lattice. Related nets with other atoms instead of carbon are mentioned. Then the symmetry of hydrocarbons having C skeletons that are fragments from the graphite net (benzenoid hydrocarbons) or from the diamond lattice (diamond hydrocarbons) is discussed, and the connection between symmetry and molecular stability is noted. 1. INTRODUCTION For mankind, C should be the most precious element because all life is based upon it, and because coal and petroleum ensure most of the energy sources. In the preceding volume, Symmetry: Unifying Human Understanding I mentioned the fascinating symmetries of the graphite and diamond lattices [I]. The present paper will review only aspects connected with this topic, and will attempt to survey other possible two- or three-dimensional carbon nets (lattices). Although none of these other nets has yet been prepared, it is of interest to discuss symmetries and relative energies or stabilities of such systems. 2. THE MOST PRECIOUS ATOMS IN THE UNIVERSE, OR WHY C? We start from the well-known fact that all nuclides (isotopes) of the same element share the same electronic configuration. To simplify the discussion, therefore, we shall ignore the composition of the atomic nuclei for the time being, and shall treat the nucleus as a black box containing Z positive charges, around which the Z negatively-charged electrons occupy orbitals of various energies, according to the principles of quantum chemistry. It took chemists more than 20 years to accept Staudinger's idea [2] that vinylic polymers are extremely long chains containing C--C covalent bonds. Likewise, cellulose and starch contain large numbers of six-membered pyranosic rings linked by covalent ethereal (C---O---C) bonds. Familiar- ity with association colloids such as soap micelles (loosely held bundles of smaller molecules or ions) had suggested that polymers possessed some kind of weaker intermolecular forces, but experiments by Staudinger, Mark, Meyer and Kuhn [3] provided evidence for the contrary, and confirmed that polymers had normal covalent bonds. Why is there just one element, then, which forms the basis of life, and which is so ubiquitous in chemistry that more than nine-tenths of all known chemical compounds contain it? What is so unique about C--C bonds that a whole branch of chemistry (organic chemistry, the largest and most important one) is devoted to it, whereas all the remaining elements studied by inorganic chemists give rise to much less numerous and less important compounds? How will the disproportion between organic, inorganic and organo-elemental compounds evolve in the future7 These are important questions for chemists and philosophers of science alike. Let us analyse first what is so unique about the element carbon. Among the 100 + elements (i.e. types of electronic configurations around the nuclei) forming the building blocks of the whole universe, only two may form homoatomic chains whose stability is independent of the chain length. Therefore only these two elements, namely tetravalent carbon and divalent S, may give rise to "infinitely long" homoatomic chains. Though S also forms stable chains or rings of various lengths and shapes, all these substances are just allotropes of the element S; 397 CORE Metadata, citation and similar papers at core.ac.uk Provided by Elsevier - Publisher Connector

Symmetry: Unifying Human Understanding · the long homoatomic C chains in organic polymers are the heteroatomic chains of silicates containing alternating Si and O atoms. Indeed,

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • Computers Math. Applic. Vol. 17, No. 1-3, pp. 397-416, 1989 0097-4943/89 $3.00 + 0.00 Printed in Great Britain. All rights reserved Copyright © 1989 Pergamon Press plc

    C A R B O N A N D ITS N E T S

    A. T. BALABAN Polytechnic Institute, Department of Organic Chemistry, Splaiul Independenlei 313,

    77206 Bucharest, Romania

    Abstract--The uniqueness of C - -C bonds, which allows life to exist and gives birth to the infinite number of organic compounds, is analysed. A brief survey of the two known C allotropes, graphite and diamond, stresses how the different structure of the nets leads to such enormous differences in properties. The synthesis of diamond is surveyed. Other possible, hypothetical forms of elemental carbon are math- ematically surveyed depending on the hybridization state of the C atoms, i.e. on the vertex degrees in the infinite lattice. Related nets with other atoms instead of carbon are mentioned. Then the symmetry of hydrocarbons having C skeletons that are fragments from the graphite net (benzenoid hydrocarbons) or from the diamond lattice (diamond hydrocarbons) is discussed, and the connection between symmetry and molecular stability is noted.

    1. I N T R O D U C T I O N

    For mankind, C should be the most precious element because all life is based upon it, and because coal and petroleum ensure most of the energy sources.

    In the preceding volume, Symmetry: Unifying Human Understanding I mentioned the fascinating symmetries of the graphite and diamond lattices [I]. The present paper will review only aspects connected with this topic, and will attempt to survey other possible two- or three-dimensional carbon nets (lattices). Although none of these other nets has yet been prepared, it is of interest to discuss symmetries and relative energies or stabilities of such systems.

    2. T H E M O S T P R E C I O U S A T O M S I N T H E U N I V E R S E , O R W H Y C?

    We start from the well-known fact that all nuclides (isotopes) of the same element share the same electronic configuration. To simplify the discussion, therefore, we shall ignore the composition of the atomic nuclei for the time being, and shall treat the nucleus as a black box containing Z positive charges, around which the Z negatively-charged electrons occupy orbitals of various energies, according to the principles of quantum chemistry.

    It took chemists more than 20 years to accept Staudinger's idea [2] that vinylic polymers are extremely long chains containing C- -C covalent bonds. Likewise, cellulose and starch contain large numbers of six-membered pyranosic rings linked by covalent ethereal (C---O---C) bonds. Familiar- ity with association colloids such as soap micelles (loosely held bundles of smaller molecules or ions) had suggested that polymers possessed some kind of weaker intermolecular forces, but experiments by Staudinger, Mark, Meyer and Kuhn [3] provided evidence for the contrary, and confirmed that polymers had normal covalent bonds.

    Why is there just one element, then, which forms the basis of life, and which is so ubiquitous in chemistry that more than nine-tenths of all known chemical compounds contain it? What is so unique about C - -C bonds that a whole branch of chemistry (organic chemistry, the largest and most important one) is devoted to it, whereas all the remaining elements studied by inorganic chemists give rise to much less numerous and less important compounds? How will the disproportion between organic, inorganic and organo-elemental compounds evolve in the future7

    These are important questions for chemists and philosophers of science alike. Let us analyse first what is so unique about the element carbon. Among the 100 + elements (i.e. types of electronic configurations around the nuclei) forming the

    building blocks of the whole universe, only two may form homoatomic chains whose stability is independent of the chain length. Therefore only these two elements, namely tetravalent carbon and divalent S, may give rise to "infinitely long" homoatomic chains. Though S also forms stable chains or rings of various lengths and shapes, all these substances are just allotropes of the element S;

    397

    CORE Metadata, citation and similar papers at core.ac.uk

    Provided by Elsevier - Publisher Connector

    https://core.ac.uk/display/82615554?utm_source=pdf&utm_medium=banner&utm_campaign=pdf-decoration-v1

  • 398 A.T. BALABAN

    they cannot give rise to a rich chemistry. When S has higher valency than two, the homoatomic S chains are no longer stable irrespective of the chain length. Chains of O, N, B, P, Si atoms are known but no stable homoatomic chains with more than 20 such atoms could be obtained; the longer the chain, the less stable the compounds.

    Recently, it was discovered, however, that chains ---(SiMe2--SiMePh)n-- may also have kinetic stability; their extreme photolability in the near ultraviolet leads to applications for photoresists [4].

    The reasons for the stability of homoatomic C chains are varied, and the most important factors are: small covalent radius, intermediate electronegativity, large bond energy due to strong overlap between bonding orbitals, lack of alternative bonding possibilities (such as ionic or covalent bonds). In most organic compounds, C is bonded mainly to H atoms (with matching intermediate electronegativity and high bond energy due to small covalent radius, hence to good orbital overlap); however, despite its high electronegativity, F also forms (with C chains) stable fluorocarbons and derivatives owing to its small covalent radius and strong bonds.

    One should include a few words about silicates and about metallic clusters. Almost as stable as the long homoatomic C chains in organic polymers are the heteroatomic chains of silicates containing alternating Si and O atoms. Indeed, most of the earth's crust contains such chains; the abundance of elements on the earth's surface places O (49%) and Si (26%) on the first two places, with C occupying the 14th place (0.1%). The two elements, O and Si, which account for three quarters of the earth's crust (including the atmosphere and the hydrosphere) form with one another a much stronger bond than the O---O and Si--Si bonds. The tetravalency of Si can and, indeed, does give rise to a rich chemistry which will continue to develop. When, in addition to O and Si, other atoms such as C and H are present, a mixed or hybrid kingdom results (in addition to the mineral, vegetable and animal kingdoms): polysiloxanes (the so-called silicones), carbosilanes, etc., with numerous theoretically and practically significant applications.

    During the last decades, metallic clusters have been amenable to detailed studies, mostly in organometallic systems. This area is also rapidly expanding and will undoubtedly cast light on many phenomena related to heterogeneous catalysis.

    Neither of these two areas has, however, the potential of organic chemistry and neither is used in life phenomena, as we know them on earth. Nor has it been possible, so far, to discover such bonding in interstellar molecules. Cosmic abundances are largest for the lightest elements, H and He; O is third, while C and Si occupy the sixth and seventh places, respectively. Many "organic" molecules are known in interstellar clouds but they are quite different from the molecules of life; most of the cosmic molecules have triple bonds; on earth, a characteristic of molecules which are stable at high-temperatures (> 2000°C) is that they also contain triple bonds (CO, N2, HCN, NO, C2H2).

    3. T H E T W O K N O W N A L L O T R O P I C F O R M S OF C: G R A P H I T E A N D D I A M O N D , OR T H E S O F T E S T A N D H A R D E S T C R Y S T A L S

    (a) Graphite There exist two types of graphite: hexagonal and rhombohedral graphite [5-7]; they differ by

    the "vertical" correspondence between C atoms in the parallel planes: in the former this leads to ABAB.. .- type, in the latter to ABCABC.. . - type planes (Fig. 1).

    The facile displacements of "molecular planes" relative to one another explain why graphite is as soft as talcum, the softest mineral, and why it writes in pencils. The Greeks must have used it for writing, hence its name [~p~tq~E~v (to write)], whereas in German pencil is called Bleistift, proving that lead, which is one of the softest metals, was used for writing, too.

    Although diamond and graphite are elemental C, in real crystals of graphite or diamond, the relatively few peripheral atoms are connected to other atoms than C, probably H or O; one may call graphite an "honorary polycyclic benzenoid aromatic hydrocarbon" and diamond an "honorary polycyclic saturated hydrocarbon" or "honorary diamond hydrocarbon", meaning that they represent the limiting (asymptotic) cases of hydrocarbons with increasing numbers of C atoms and decreasing numbers of H.

  • Carbon and its nets 399

    (a)

    t I I

    (b)

    I I I I I I

    I I I I I~

    ! I I

    Fig. 1. The two types of graphite: (a) hexagonal; (b) rhombohedral.

    The "honeycomb lattice" of carbon atoms in graphite is actually a huge planar molecule, wherein all atoms are connected by covalent bonds. The parallel planes are, however, loosely held by van-der-Waals-forces which are much weaker than covalent bonds, accounting for the softness of graphite, whose molecular planes may glide over one another. The strong anisotropy of graphite is also explained by its structure. The linear compressibility of graphite is 104-105 times larger on a direction perpendicular to the molecular planes than along directions within the molecular plane. The thermal and electrical conductivities in single crystals of graphite are about 200 times higher within the molecular planes than in the orthogonal direction. Also, the fact that graphite is black is due to light absorption in the huge aromatic chromophore.

    Interatomic distances give evidence for the different kinds of bonding in graphite: its C-C covalent bond has a length of 141.5 pm, whereas the distance between the molecular planes is 335 pm. This fact leads to a low density (2270 kg.m -3 in the ideal case, usually less).

    All C atoms in a planar graphite lattice have sp 2 hybridization, and all interatomic bond angles are 120 °. The electronic delocalization within the molecular plane makes graphite the most stable allotropic form of C.

    Of particular interest are the so-called intercalation compounds formed by graphite by including between the molecular planes various atoms such as alkaline metals, or molecules such as CrO3, modifying thereby their reactivity, and extending slightly the inter-layer distance in the graphite net [7]. Most such intercalation compounds still conduct electricity.

    The addition of F: to graphite leads to the white non-conducting "perfluorographite" (CF), in which the C atoms with sp3-hybridization cause buckling of the "molecular planes" (Fig. 2).

    Since carbon has two stable nuclides, namely J2C (98.9%) and 13C (1.1%), and since the absorption cross sections for thermal neutron capture are quite low (0.0034 and 0.0009 barn,

    F F F F

    F g F" g

    Fig. 2. Formation of "perfluorographite".

  • 400 A.T. ]3ALABAN

    respectively), much lower than for protons (0.332 barn) or ~70 (0.235 barn), the first nuclear reactors were moderated with graphite.

    Nowadays only few reactors are moderated with graphite; most of the nuclear power reactors use H20 or D20 as moderator and coolant systems. This is because the atomic mass makes C a less effective moderator in elastic interactions with neutrons than protium or deuterium, therefore the power density of the active zone is appreciably lower: U rods in graphite-moderated reactors must be spaced at larger distances than in H20- or D20-moderated reactors. Moreover, graphite may ignite at high temperature, and under neutron bombardment the C atoms are displaced accumulating strain energy (the Wigner effect responsible for the Windscale accident of the nuclear reactor in 1957). The Chernobyl tragedy proves that nuclear reactor accidents may become aggravated by the high temperatures caused by graphite combustion. A further drawback of graphite reactors is their positive void coefficient which by a feed-back loop may cause a catastrophe. In the US, the so-called N-reactor at Handford, Washington (which uses graphite as moderator and, like the Chernobyl-type reactors, lacks a large containment building and is used also for the manufacture of Pu) is reported to be designed so as to have a negative void coefficient [8].

    A major use of graphite is nowadays for C fibers as reinforcement materials for composites. In oriented crystals, graphite can have a high modulus if the force is parallel to the molecular plane. The industrial C fibers used in high-technology composites are produced by pyrolysis of polymers in controlled conditions. The largest single U.S. producer of carbon fibers had sales of $64 million in 1985, despite the pace slackening of this market [9]. Such fibers are manufactured from pitch or polyacrylonitrile fibers; they are oxidized at 250 ° to lose much of the H, then carbonized to an extent of 98% at 800 ° in the absence of 02, when cyclizations may occur, and finally graphitized at 1400-2500°C when loss of H2, N2, HCN, NCCN leaves a graphitic material with few N atoms.

    Composites made from high-tech polymers (e.g. Kevlar, an aramid fiber with high tensile strength obtained from p-phenylenediamine and terephthaloyl chloride) and carbon fibers with high compressive strength, have many potential applications for replacing metals.

    (b ) Diamond In a diamond single crystal (Fig. 3) all C atoms have sp3-hybridization with tetrahedral bond

    angles and interatomic bond lengths 154 pm as in alkanes. The distance between parallel average planes is 140 pro, much smaller than in graphite, which is therefore less dense. Therefore the single crystal is a huge tridimensional molecule, explaining thus its hardness, transparency, high density (ideally 35 l0 kg. m -3) and lack of thermal or electrical conductivity [5, 10].

    Measurements of heats of combustion indicate that diamond is thermodynamically less stable than graphite under normal conditions by 0.2 kcal/mol. However, the diamond-to-graphite conversion has a high energy barrier: only o n heating for a long time at 1500 ° under inert atmosphere at normal pressure is it possible to perform this conversion. Thus at 25 ° and 1 bar, diamond is metastable, i.e. it has a high kinetic barrier.

    The reverse process, namely the production of artificial diamond from graphite, requires high pressure (Le Chatelier's principle accounts for the higher stability of the denser diamond at higher pressures). In the pressure vs temperature diagram (Fig. 4) the hatched area corresponds to the industrially applied process, which includes, in addition to C, a catalyst. Figures 5 and 6 represent the pressure/temperature/composition diagrams with Ni catalysts.

    In manufacturing diamond it is essential to remember that at any given high pressure, the temperature must be high enough for the conversion to proceed at a convenient rate, but not higher

    Fig. 3. The two types of diamond: (a) cubic (with the sphaleritc lattice, having no eclipsed bonds); (b) hexagonal (with the wurtzite lattice).

  • Carbon and its nets 401

    pressure temperature diagram

    /

    T (kbor) - - " - 4001 ~ l i q u i d

    _ i "--...,,.... diamond 30 U. ""'-....

    diamond a n d " ' ' - . . 200' --..

    metasiable g r a p h i t e ' ' ~

    1000 2000 3000 4000 T IK ) Fig. 4, The pressure-temperature diagram of C.

    than the equilibrium indicates because always diamond is metastable relatively to graphite, and at the highest temperatures only graphite is stable. This is clearly seen in Fig. 5.

    The diamond-producing apparatus, catalyst/solvent, and pressure-temperature conditions con- stitute a well-guarded industrial secret. From literature data [11-13] one gets general ideas, e.g. with two pistons one has the "Bridgman opposed anvils" (tetrahedral assemblies of four pistons may also be imagined or hexahedral ones with six pistons)• Hydraulic rams act the pistons. Figure 7 illustrates the widely used "belt apparatus". The gasket material is a hydrated Al silicate (pyrophyllite). The high temperature (1700-1720 K) is obtained in the cell by passing electric current from the top to the bottom piston, after the onset of the required pressure (about 54 kbar, i.e. 5.4 GPa). The heat is evolved mainly in the graphite layers which are less conductive than nickel• Diamond forms at the interface by dissolution of graphite into the molten nickel and precipitation of the less soluble diamond, as crystals. After cooling, the nickel is dissolved in acid leaving the diamond crystals•

    A different procedure for obtaining synthetic diamond involves the deposition of C atoms in vacuum on small diamond crystals. The C atoms are formed by pyrolysis of methane at 1000°C; the resulting larger diamond crystals are freed from graphite impurities by hydrogenation at 50 bar.

    An interesting observation concerns tiny diamond crystals (typical diameters are 5 nm) in carbonaceous meteorites; the isotopic composition of accompanying Xe indicates that the meteorites did not originate in our solar system [14]. It is not certain how the initial diamond crystal resulted; further accretion then occurs as indicated above•

    There exist two forms of diamond: the common (cubic) latticee (identical to the ZnS sphalerite lattice), where all C - -C bonds are in staggered conformation, and the hexagonal (wurtzite) lattice which has Pitzer strain due to eclipsed C- -C bonds. One can see adamantane subunits in the former lattice and iceane [15] subunits in the latter (Fig. 3).

    80

    p(kbar) me,t,n

    diamond.Vy /7~.Jmelt ing J ~/l'*'of N i -

    graphlte ~' ,

    1000 2000 T(K) Fig. 5. The pressure-temperature diagram for the

    C--Ni system.

    20-

    temp

    2000 liq • K _ e

    1000-

    0 % C 100 % Ni

    K 1667 K "" l i q+ d iamond

    ~t . d iamond

    100% C O%Ni

    Fig.

    / , 6 0 0 K

    m.po )raphite

    6. The temperature-composition diagram for the C--Ni system at 54 kbar.

  • 402 A.T. BALABAN

    t

    bert apparatus

    \

    current disk

    another cell for slow growth of large diamond crystals

    ,'" \ k \ Pist°n/ l/', ,' ' ,

    celt

    /heater tube

    white=insulator .=== =seed bed

    Fig. 7. The belt apparatus for synthetic diamond: at the bottom, an enlargement of the central zone is shown.

    4. CARBON M O L E C U L E S

    We shall now discuss two types of relatively small C molecules C,, with n ~< 120, which have been predicted to have some chemical stability.

    ( a ) Buekminster ful lerene, C 6o

    A polyhedral C60 molecule with hexagonal and pentagonal faces, footballene or buckminster- fullerene was discussed [16]; it is still debatable if the experimental evidence is conclusive. In the spherical cavity, a metallic atom (e.g. Ln) may be incorporated. Figure 8 presents this molecule and a planar projection (Schlegel diagram).

    A larger cluster of C atoms (Ct/0, archimedane) corresponding to a semiregular polyhedron with 12 decagonal, 20 hexagonal and 30 square faces, was discussed [17].

    Considering that on replacing some of the six-membered rings in graphite by five-membered rings one induces a curvature in the plane of the lattice which converts it into a polyhedral quasi-spherical/system, one may speculate in abstract mathematical terms if an analogous procedure would be feasible for diamond by replacing some adamantane units in order to arrive at four-dimensional polytopes.

    (a) (b)

    Fig. 8. (a) A view of buekminsterfullervne or footballene, C60; (b) the same molecule in a Scldegel diagram showing all faces.

  • Carbon and its nets 403

    Fig. 9. A C48 cluster of C atoms. Fig. 10. Angles in a regular n-sided polygon, which is drawn with thick lines.

    (b ) A smaller sperhical C~8 molecule

    In a recent paper [18] on a three-connected C net (which will be presented in a subsequent paragraph), a C~ molecule was presented. It has six puckered eight-membered rings in a cubical arrangement, and eight nine-membered rings (Fig. 9). Its C atoms, unlike those of buck- minsterfullerene, have sp- and sp2-hybridization. Similarly to this molecule, it may accommodate a metal atom or ion inside the cavity.

    (c) Cyclocarbyne (cyclokarbin)

    The speculation on smaller molecules having a large ring of sp-hybridized C atoms with small steric strain was published in 1980 [20]. This molecule with either alternating single and triple bonds, cyclo----(~C--), /2, or with cumulenic double bonds, cyclo -~-(~------C~)-~-,/2, might be isolable if the angle strain is divided over a larger number of C atoms; e.g. for n = 50 the deviation from the bond angle of 180 ° becomes ~ = 7.2 °. In the general case one may easily see from Fig. 10 which shows the angles in a regular n-sided polygon that this deviation is ~ = 360°/n: the deviation equals the angle with which the centre of the polygon "sees" each side of the polygon. The black dots in Fig. 10 symbolize the C atoms.

    5. H Y P O T H E T I C A L A L T E R N A T I V E C ALLOTROPES ( I N F I N I T E NETS)

    The various possible nets will be systematically discussed according to the coordination number of C atoms: two-coordinated atoms will be denominated as sp-hybridized, three-coordinated ones as sp:-hybridized, and four-coordinated ones as spLhybridized (even when, owing to steric factors, the hybridization may no longer correspond to these pure states).

    .4. Systems Having sp-Hybridized C ,4toms (Two-connected Nets)

    (a) Polyyne C

    The literature data on a linear macromolecular polyacetylenic (or poly-cumulenic) form of C (also called carbyne, Karbin or chaoite [21]) is still contradictory. It is obtained by oxidative polymerization of acetylene and appears to be different from diamond or graphite. It contains, probably, in addition to the linear chains, portionwise structures which link together several such chains such as those presented in Fig. 11 (of course in these portions the C atoms have sp L or sp 3-hybridizations).

    (b) Other systems containing sp-hybridized C together with other types of C

    In principle it is possible to introduce, into any two- or three-dimensional network, layers of sp-hybridized C atoms. This will cause rings in graphite to become elongated in one direction [Fig. 12(a)] or in several directions [Fig. 12(b)].

    Similarly, a row of C - - C bonds in diamond may be replaced by C - - C ~ - C - - C bonds as hypothesized by Melnichenko et al., affording the so-called polyyne-diamond [22] (Fig. 13).

  • 404 A. T. BALABAN

    (a l {b)

  • Carbon and its nets 405

    i

    I I I IIIi

    Fig. 13. Polyyne-diamond.

    I III

    Fig. 14. Fracture zone in graphite leading, after gliding, to the formation of four- and eight-membered rings along the

    shear zone.

    Newer calculations have been reported by Burdett and Lee [29] using the moments method for the same net {4, 82 } which is known to exist in the anionic part of Ca[B2C2].

    (b ) Networks with pentagons and heptagons

    Bathroom floor tilings corresponding to the previously described networks are fairly common. No tilings with pentagons and heptagons are known to the author, however. One may imagine two such lattices [27]; the first called {52,72) has eight C atoms in the elementary cell (Fig. 16); the second, called {53, 73} has 12C atoms in the elementary cell (Fig. 17).

    A calculation similar to the preceding one [28] on the lattice {52, 7 5 } affords a resonance energy of -4.9kcal/mole. Assuming the least strained geometry (angles in the regular pentagon are fl = 108°;other angles are ~ = 116°,~ = 144°,E = 136.5 °) one calculates a strain energy of + 3.8 kcal/mole resulting in a total energy of - 1.1 kcal/mole, i.e. conjugation energy prevails over strain energy; however, the total energy is higher than for graphite (namely by 5 kcal/mole), but less so than for lattice {4, 82}.

    Burdett and Lee [29] also reported a calculation for the net represented in Fig. 16, which is known to exist in the anionic part of Sc[B2C2]. They also calculated energies for the least stable planar C net (3, 95 } having equilateral triangles and 9-gons (at each sp2-hybridized carbon atom, one triangle and two eneagons meet).

    (c) Other planar nets

    The three-connected net presented in Fig. 18 contains decagons and hexagons (Hiickel-type systems, hatched in the figure) and pentagons. The unit cell contains 12 atoms. The steric strain is small, so that this net is likely to present a favourable calculated total energy, similar to that of the net {52, 72}. Calculations are in progress.

    (d) Tridimensional nets: a polyphenylene net reminiscent of zeolites

    The first attempt to discuss a three-dimensional C net formed from spE-hybridized atoms was published in 1946 and 1950 by Riley and coworkers [30]. The net is formed by infinite chains of non-coplanar benzenoid rings linked as in ortho-tetraphenylene. The result is a three- dimensional net reminiscent of zeolites, including large holes and long channels.

    Lattice .[/,, 82~ .

    Fig. 15. The planar lattice {4, 82 } with a unit cell ( - - - ) . Fig. 16. The planar lattice {5 2, 7 2 } with a unit cell ( - - - ) .

  • 406 A.T. BALABAN

    / •

    / / • ,,

    /

    Fig. 17. The unit cell of the planar lattice {53 , 73}. Fig. 18. A planar net with pentagons, hexagons and decagons; the unit cell, containing 12 C atoms ( .) , is marked

    with dashed lines.

    (e) A hypothetical C aliotrope with possible metallic character

    Interesting three-dimensional networks formed from trigonal sp 2 C atoms were analyzed by Roald Hoffmann et al. [31]: the structure of ThSi2, with C atoms instead of Th and Si. There are infinite polyene chains running along two dimensions with no conjugation along the third dimension. The density is calculated to be 2970 kg. m-3, intermediate between that of diamond and graphite. The smallest rings are 10- and 12-membered. It was calculated that this form should have metallic conductivity, and that the C--C distances should be intermediate between those in graphite and diamond (the non-conjugated bonds should be longer than the polyenic ones). The calculated energy is higher than that of graphite by 0.74 eV/C atom. In order to make comparisons possible, one should keep in mind that 1 eV/atom corresponds to 23 kcal/mole, therefore the net of Fig. 19 is less stable than graphite by about 17 kcal/mole. The authors argue [32, 33] that this net may be able of existence owing to high kinetic barriers characteristic of C--C bonds.

    Fig. 19. A tridimensional net from sp2-hybridized C atoms with possible metallic character.

    (b]

    Fig. 20. A tridimensional net from sp2-hybridizcd C atoms with large cavities: (a) the larger cavity; (b) the smaller

    cavity.

  • Carbon and its nets 407

    ( f ) A tridimensional net with large cavities

    An interesting tridimensional C net formed from sp2-hybridized atoms with large quasi-spherical cavities and smaller quasi-cylindrical ones was discussed [18]. Its large cavities can be visualized by considering the semiregular polyhedron called truncated octahedron; in each of its 24 vertices three edges meet, belonging to two hexagons and a square. On replacing all edges that are common to a square and a hexagon by a sequence of two C- -C bonds, i.e. by a C - - C - - C chain, one obtains the large cavity. It has six puckered (crown) octogons whose average planes are parallel to the faces of a cube, and eight puckered 9-gons (eneagons) whose average planes are parallel to the faces of an octahedron. Each octogonal face is surrounded by four 9-gons, and each 9-gon is surrounded by three 8-gons and three 9-gons.

    In the net, adjacent large cavities share their 8-gonal faces. The "free spaces" between these cubically-packed quasi-spherical large cavities form the smaller cavities; these consist of two parallel puckered 9-gons (which are the bases of the quasi-cylinders) connected by three parallel C - -C bonds.

    The large and smaller cavities are shown in Fig. 20(a) and (b). The net is zeolite-like. The large cavities are isomorphic to the C4s cage molecules discussed above under Section 4.

    The net has 8-gons as the smallest circuits; some of them are boat-shaped (three such 8-gons form a small cavity), other ones are crown-shaped (the faces common to two adjacent large cavities). Next, there are 9-gons, all of which have the same conformation, most easily seen in the representation of the smaller cavity [Fig. 20(b)].

    C. Systems Having Only sp~-Hybridized C Atoms (Four-connected Nets)

    (i) A planar (square) lattice of C atoms with four-fold symmetry would have a high strain energy: all C atoms would be as strained as the central C in fenestrane ("planar C").

    (ii) The tridimensional truncated octahedral lattice contains 12C atoms in the unit cell (Fig. 21). The angle strain in the four-membered rings and the planar six-membered rings amounts to at least 12 kcal/mole of C atoms [27].

    D. Systems with sp 2- and sp3-Hybridized C (Three- and Four-connected Nets)

    The paper published by the present author in 1968 [27] contained only a brief statement on the two lattices shown in Fig. 22 to the effect that the four-connected "planar" C atoms will lead to high strain. Other authors [33] discussed related topics.

    A different approach was recently initiated, by imagining and calculating the energies of three- and four-connected nets possessing little strain; some of the nets may present metallic properties [34]. The densities of such systems (about 3000 kg/m 3) are intermediate between those of graphite and diamond.

    Some of the atoms of Fig. 23(a), namely those which are shown without any double bonds, are forced to become tetracoordinated by forming an extra bond to a similar atom from a string of pentalene units in a parallel plane. For viewing this process, imagine that in a graphite-type stacking of molecular planes as in Fig. 23(a), half of the strings of pentalene units (black points) are translated midway between the molecular planes. Then the new bonds are formed, and the three- and four-connected lattice [Fig. 23(b)] results.

    Fig. 21. The truncated octahedron. Fig. 22. C Nets with tetra- and tricoordinated C atoms; the former atoms are represented by black dots.

    CAMWA 17-I/3---Z

  • (a)

    408 A.T. BALABAN

    (b)

    Fig. 23, (a) The starting planar arrangement of strings of pentalene units; (b) the 3,4-connected net.

    The energy of this net may be calculated by means of the extended Hiickel method in the tight binding approximation. The parameters used for C and the geometries are specified in Ref. [34]. The result is 1.19 eV C atom (27 kcal/mole) above the energy of graphite.

    An energetically more favourable three- and four-connected net results by a similar procedure starting from the planar net containing five- and eight-membered rings [34] as seen in Figs 24 and 25. Again, the black atoms are translated with half the spacing between the molecular planes, and the atoms (black and white) which are shown without double bonds become tetracoordinated.

    (

    (

    ( Fig. 24. The starting arrangement for another 3,4-connected net, similar to that of Fig. 23(a).

  • Carbon and its nets 409

    Fig. 25. Two types of delocalization in the strings of condensed five-membered rings in Fig. 24.

    The energy of this net is calculated by the extended H/ickel method to lie at 0.86-0.97 eV/C atom (20-22 kcal/mole) above the energy of graphite. The range of values is due to the fact that one may conceive a delocalized polyenic system, or two different types of localization (Fig. 25).

    Finally, among the many imaginable nets of such type (discussed by Wells in a geometrical context) a higher-energy system depicted in analogous fashion contains eight- and four-membered rings [34], cf. Fig. 26.

    The energy of this system is calculated to lie 1.27 eV/atom (29 kcal/mole) above the energy of graphite.

    In concluding this chapter, mention should be made of strain-free oligoradicals [35] which contain eight-membered rings.

    6. R ELATED NETS WITH OTHER ATOMS INSTEAD OF C

    Both the diamond and the graphite lattice may be encountered in compounds with other elements than C. Thus the diamond lattice (with lower cohesion because of the longer interatomic distances and smaller bond energies) occurs also in elemental Si (bond length 234 pm), Ge (bond distance 240 pro), and in gray tin (bond length 280 pm). On reaction with concentrated hydrochloric acid, this crystalline form of tin affords SnCI4" 5H20, whereas the white polymorphic tin yields

    (b)

    (a)

    <

    Fig. 26. The starting arrangement for a third 3,4-connected net (a); (b) the resulting net.

  • 410 A.T. BALABAN

    SnCI2.2H20 indicating that in the latter form with longer interatomic distances only two electrons in the valence shell are involved. In addition to these elemental lattices, the diamond structure is encountered also in cubic boron nitride, a compound whose hardness is second only to diamond. Since each pair of BN atoms has eight valence electrons exactly like a pair of C atoms, one might expect that under high temperatures and pressures the diamond-type lattice would result for (BN)x; indeed, about a year after the successful diamond synthesis, Wentorf Jr performed this conversion [36] under conditions similar to those used for synthetic diamond, but using as catalysts/solvents alkali metal nitrides. The Knoop hardness of (BN)~ is 4500 kg/mm 2 (compared to 9000 kg/mm 2 for diamond and 2100 kg/mm 2 for sapphire), therefore cubic (BN)x is used as an abrasive which, unlike diamond, may be heated in 02 and does not react with Fe at high temperatures. This quality is useful in sharpening and shaping hard tool-steels and high-strength alloys.

    On the other hand, hexagonal B nitride has a lattice which is similar to that of graphite, is thermally almost as stable and cleaves similarly to graphite; however, it is white, non-conducting, and does not afford intercalation compounds. It is used for high-temperature crucibles. Apparently, the n-electron delocalization is much smaller than in graphite, owing to the formal charges on quadrivalent B- and N ÷ atoms.

    Passing now to the hypothetical nets discussed above, the trivalency of B and N would convert some infinite lattices into finite molecules. Alternatively, one may imagine new oxides of C; thus the large cavity of Fig. 9 or Fig. 20(a) would become (CO)24 on replacing the vertices of degree two by O atoms, and the pairs of vertices of degree three by C------C double bonds.

    7. SYMMETRY OF FRAGMENTS FROM THE GRAPHITE LATTICE

    All benzenoid polycyclic aromatic hydrocarbons (PAH'S or benzenoids, for brevity) may be considered to represent fragments from the graphite lattice, whose peripheral C atoms are linked to H atoms.

    Considerable interest is associated with these systems because of the proven carcinogenic activity of many compounds from this class possessing "bay-regions". Such a compound is benzo[a]pyrene, which results in many combustion processes and is present in cigarette smoke and in engine exhaust gases, cf. Fig. 27.

    Benzenoids were classified into cata-condensed (catafusenes) and peri-condensed (perifusenes) depending on whether they do not contain or do contain, C atoms belonging to three adjacent six-membered rings (so-called internal C atoms). A newer definition is based upon the notion of dualist graphs [37], whose vertices are centres of hexagons and whose edges connect vertices corresponding to "condensed hexagons", sharing an edge and two C atoms as in naphthalene whose dualist graph has just one edge. One should note that unlike normal graphs where angles between edges are arbitrary, in dualist graphs the geometry is essential; for instance, the dualist graph of anthracene has two collinear edges (with angle 180°), whereas that of the isomeric phenanthrene has two kinked edges with an angle of 120 °. Catafusenes have acyclic dualist graphs (trees), while perifusenes have three-membered rings in their dualist graphs. All catafusenes with the same number of vertices in their dualist graphs are isomeric. By this definition, a third category of benzenoids exists, namely corona-fused system (coronoids) whose dualist graphs possess larger rings than three-membered, which are not contours of aggregates of three-membered rings.

    A "periodic system" of molecular formulas was proposed by Dias [38] for benzenoids CxHy on the basis of two parameters, N (the number of internal C atoms, i.e. zero in catafusenes, two in pyrene etc.) and d (the net disconnection, equal to one plus the number of tertiary C atoms minus the number of internal graph edges): x - y = 2(N + d + 1). The number of tertiary C atoms is twice the number of hexagons minus two.

    ~regions/~~ Fig. 27. Two carcinogenic hydrocarbons with their bay regions: benzo[a]pyrene and benzanthraeene.

    Neither has any symmetry.

  • (o)

    (d)

    tg)

    9

    Carbon and its nets

    ( b ) ~ ( c ~

    (e) (f)

    Fig. 28. The carbon skeletons of polycyclic benzenoid hydrocarbons: (a) anthracene and (b) phenanthrene are two isomeric (Ci4Hi0) non-branched catafusenes; (e) triphenylene (CtsH~2) is a branched catafusene; (d) perylene and (e) pyrene are perifusenes. In all cases a single Kekul6 structure is shown; (f) kekulene, is a coronaphene; (g) is [7] helicene. In the last two cases double bonds are no longer included. The dualist

    graphs are indicated with broken lines.

    411

    Since annelation of catafusenes may be performed in three directions at a marginal benzenoid ring, the numbers Ch of non-branched catafusenes with h six-membered rings are easily calculated [36, 38]:

    4Ch=3h-2+4.3~h-3)/2+ 1, for odd h;

    4Ch = 3 h-2 + 2.3 th-2~/2 + 1, for even h.

    Branched catafusenes can also be counted but complicated formulas result. In these formulas benzenoids include also helicenes, i.e. systems with superposed C atoms whose molecules cannot have a planar structure; helicenes cannot be considered as portions of the graphite lattice. If helicenes are left out from the benzenoid count (as we shall do henceforth), the enumeration becomes difficult even for non-branched catafusenes and only recursive computer algorithms can be employed for this purpose, up to given h values. So far, perifusenes have been counted only by computer algorithms for h ~< 11 [40, 41]. The enumeration of benzenoids is part of a larger unsolved problem in graph theory, namely the "cell growth problem" wherein a cell is a triangle, square or hexagon, and the "animal" consists of condensed cells, sharing an edge.

    A coding system was devised for dualist graphs of catafusenes [37, 39] based upon the three orientations of edges in the graphite lattice, coded by digits 0, 1 and 2 for annelating a terminal benzenoid ring with angles 180 °, 120 ° and 240 °, respectively. The code rules state that coding starts from an endpoint of the dualist graph, that among all possible codes the minimal number is adopted on reading sequentially the digits, and that for branching one uses brackets.

    Figure 28 illustrates catafusenes, perifusenes, coronoids and helicenes. In turn, the symmetry of non-branched catafusenes can be used [37, 40] for systematic generation

    and enumeration. Table 1 presents the numbers of non-branched catafusenes with following symmetries: a =acenes (linear condensation); m = m i r r o r symmetry; c =inversion centre; u = unsymmetrical, cf. Fig. 29.

    Gordon, Cyvin, Gutman, Trinajstir, Knop and other authors [40-42] investigated in detail the numbers of Kekul6 structures for benzenoids using various techniques. Perifusenes have been further classified on the basis of their Kekul6 structures into normal ones (with at least one Kekul6 structure), essentially disconnected ones having some single bonds in all Kekul6 structures, and non-Kekulran systems (radicals or polyradicals), i.e. systems without any Kekul6 structure; in

  • 412 A . T . BALABAN

    Table 1. Numbers of benzenoids according to the type of condensation (cata, peri, corona), branching, symmetry and Kekul6an character

    Catafusenes Perifusenes Coronoids

    Non-branched t Kekul6an Kekul6an Total

    h a m c u Total Branched Total Yes No Yes No benzenoids

    1 1 0 0 0 1 0 1 . . . . 1 2 1 0 0 0 1 0 1 . . . . 1 3 I I 0 0 2 0 2 0 1 - - - - 3 4 1 1 1 1 4 1 5 1 1 - - - - 7 5 1 4 1 4 10 2 12 3 7 - - - - 22 6 1 3 4 16 24 12 36 15 30 - - - - 81 7 1 12 44 50 67 51 118 72 141 - - - - 331 8 1 10 13 158 182 229 411 353 671 1 0 1436 9 1 34 13 472 520 969 1489 1734 3282 3 2 6510

    10 1 28 39 1406 1474 4098 5572 8535 15979 24 19 30129 I I 1 97 39 4111 4248 16867 21115 41764 78350 128 155 141512 12 1 81 116 11998 12196 68925 81121 ? ? 854 1100 ?

    ?a = acenes (linearly condensed); m = mir ror plane perpendicular to the molecular plane; c = centrosymmetric; u = unsymmetric.

    graph-theoretical language, such systems are said to be non-decomposable into 1-factors, or to be devoid of perfect matchings. All systems with odd numbers of C atoms are non-Kekul6an; some peri-condensed systems with even numbers of C atoms are also non-Kekul6an, e.g. diradicals. All catafusenes are Kekul6an.

    Calculations of re-electron energies in large peri-condensed benzenoid systems show [43] that such systems tend gradually towards graphite; these calculations allow to evidence how the topology of peripheral C atoms influences their properties.

    Table 1 presents the numbers of benzenoids, excluding helicenes, according to [42].

    8. SYMMETRY OF FRAGMENTS FROM THE DIAMOND LATTICE

    Cyclic molecules strive to adopt low-energy strain-free conformations. For many molecules possessing saturated six-membered rings, such conformations are portions of the diamond lattice. Thus the chair form of cyclohexane, the two (cis and trans) isomers of decalin, and all the diamond hydrocarbons (adamantane, diamantane, and polymantanes) have carbon skeletons which are fragments of the diamond lattice. This was observed long ago by Wittig [44], Lukes [45] and others.

    We shall pay special attention to the so-called "diamond hydrocarbons" or "polymantanes", consisting of adamantane units fused along a "face" (Fig. 30). Actually, adamantane is a tetrahedrane whose C--C bonds have been replaced by a 3C chain C--CH2--C. The fusion involves joining two such tetrahedra along a face to yield a trigonal bipyramid.

    In collaboration with von R. Schleyer [46], these diamond hydrocarbons have been enumerated and classified according to their symmetry, on the basis of their "tridimensional dualist graph". A coding system was devised based upon the four tetrahedral directions of the dualist graph, each denoted by one of the digits 1, 2, 3 or 4. This coding system was employed also for staggered rotamers of alkanes [39, 47] and was shown to possess a relationship with the three- digit coding of benzenoids [37]. As in the case of benzenoids, catamantanes and perimantanes have been defined according to whether the dualist graph is acyclic or cyclic, respectively. Catamantanes may be linear or branched, according to their dualist graphs. Whereas all catafusenes with the same number h of benzenoid rings are isomeric and have formula C2 ÷,h H4 ÷ 2h, only "regular catamantanes" are isomeric among themselves and have formula C 4 n + 6 H 4 n + ~ 2 ,

    (a) (b) (c}

    Fig . 29. V a r i o u s s y m m e t r i e s o f n o n - b r a n c h e d c a t a f u s e n e s fo r t h r ee i s o m e r i c C t s H t : h y d r o c a r b o n s : (a) acene ; (b) m i r r o r s y m m e t r y ; (c) c en t r i c s y m m e t r y .

  • (a)

    Carbon and its nets

    (b) (c)

    Fig. 30. D i a m o n d hydrocarbons: (a) adamantane; (b) d iamantane (earlier n a m e - - t h e emblem of the 19th IUPAC Congress held in 1963 in L o n d o n - - w a s congressane); (c) t r iamantane.

    413

    where h, n > 1 are integers; they result on annulating a face of a polymantane on replacing three axial hydrogens by a trimethylenemethane group, resulting in a net addition of C4H4. The second kind of catamantanes called "irregular", are obtained by annulation at a face having less than three axial hydrogens, involving the net addition of CxHy, where x, y < 4. The smallest irregular catamantane is [1231]pentamantane, C2sH30. A necessary and sufficient condition for a catamantane to be irregular is to have a code with the same digit separated by two other digits (e.g. the digit 1 in the above code). Table 2 presents all possible catamantanes with n = I-6 adamantane units.

    The coding starts from one end of the dualist graph and goes on registering the orientation of each bond; the orientations of the first two bonds are always l then 2. We adopt the convention of obtaining the minimal number when reading sequentially all digits, among all possible orientations of the dualist graph with respect to the tetrahedral coordinates; therefore the first two digits of the code will always be 12. When branching occurs, the branch is included in brackets; for geminal branching, the two branches are separated by a comma within the bracket. Thus code 121 designates the dualist graph shaped like the C-skeleton of anti-n-butane, code 123 designates its syn-rotamer, code 1(2)3 the C-skeleton of isobutane, and code 1(2, 3)4 the C-skeleton of neopentane.

    In the tridimensional Euclidean space E3, slight deviation from normal bond angles can accomodate such molecules as helicenes or helicene-like staggered C-rotamers of alkanes; however this is no longer possible for diamond hydrocarbons: when two vertices in the dualist graph of a polymantane are at the distance of one edge, these two vertices must be linked in the E3 space, and the dualist graph must be cyclic. It is interesting to speculate about helicene-lik¢ polymantanes in a tetradimensional space E4, where such ring-closure conditions for the dualist graph are not mandatory.

    Table 2. All possible catamantanes with n = 1-6"['~:

    Regular Irregular

    Linear Branched Linear Branched Total

    n Code Sym. Formula Code Sym. Code Sym. Formula Code Sym. No.

    1 - - T d CioHt6 . . . . . . . I

    2 1 D3d C t 4 H 2 0 . . . . . . . 1

    3 12 C~ C,sHu . . . . . . . 1

    4 121 C~ C22H2~ 1(2)3 C~ . . . . . 3 123 C 2

    1212 C2~ 12(1)3 C~ 5 1213 C l C2~ H32 12(3)4 C s 1231 Cs C25 H30 - - - - 7

    1234 C 2 1(2, 3)4 T d

    12121 C~ 121(2)3 C, 12123 Cj 12(1)32 C I 12131 C~ 121(3)4 C, 12(1)31 Cj

    6 12134 Ci 12( 1)34 C, 12132 C I 123( 1)2 C I 12321 Ci C30H3~ 12(1, 3)4 Cj 12314 C 1 C~H~ 123(I)4 C I 23 12324 C2 12(3(12 C~ 12(3)41 C, 12341 C 2 1(2)3(I)2 C~

    1(2)314 C~ 12(3) 14 Ci 1(2)3(I)4 C~

    t ln addition, one perimantane is possible for n = 6; it has code 12312, formula C26Hs0, and symmetry D ~ ~AII systems with point groups C~ or C2 are chiral and give rise to two enantiomers.

  • 414

    H

    H H B A C

    Fig. 31. Imidazole (A) has an unusually high basicity and high acidity because its anion (B) and cation (C) have

    higher symmetries than (,4).

    A. T. BALABAN

    .CH 3 O=C-Ct

    1CH313C C (CH3) 3

    -H20 -He/-H " I i -

    Fig. 32. Formation of a pyrylotrimethinecyanine in its two resonance forms; crosses indicate ten-butyl groups,

    C(CH3)3, and Ph phanyl groups.

    Symmetry considerations have been helpful also in the enumeration of staggered conformers of alkanes [47]; with some restrictions and pruning, these staggered conformations are the dualist graphs of diamond hydrocarbons (polymantanes).

    9. S Y M M E T R Y A N D S T A B I L I T Y O F M O L E C U L E S

    In many cases, high molecular symmetry leads to enhanced stability. This is illustrated by many examples. Thus, isovalent conjugation always corresponds to higher stability than sacrificial conjugation, as illustrated by the high stability of cations and anions derived from azoles (pyrazole, imidazole), cf. Fig. 31. The acidity of the carboxyl group and the basicity of guanidine are due to the symmetry of the corresponding anion and cation, respectively. The spectacular bathoehromic effects associated with the acid-base equilibria of polyarylmethane dyes such as phenolphthalein are also due to the formation of symmetrical ions whose delocalization becomes extended over a large array of atoms. The largest bathochromic shifts in dyestuffs (linearly depending on the number of ~-------CH-- groups) are present in symmetrical polymethine-cyanines with the general formula Ht--,-~CH)~-----Ht +, where Ht is a heterocyclic system: pyridinium, (iso)quinolinium, pyrylium, cf. Fig. 32 [48]. With such cyanines (heptamethine-cyanines or longer cyanines) one may obtain electronic absorption in the infra-red, with obvious applications for photographic sensitization, etc.

    It is interesting to observe that also in neutral molecules high molecular symmetry leads to enhanced stability. Schleyer demonstrated experimentally [49] that adamantane and diamantane, which are highly symmetrical molecules with the lowest energies among their valence isomers due to the lack of steric strain and of eclipsed interactions (Pitzer strain), are the final products in the AICl3-catalysed isomerizations of such valence isomers. Also triamantane and one tetramantane were obtained by such isomerizations [50]. Similarly, dodecahedrane with its very high symmetry, is formed by Rh-catalysed isomerization from an isomeric product named pagodane [51], as seen in Fig. 33. Symmetry, lack of strain and hence higher stability, are the reasons for the presence of adamantane and diamantane in crude oil [52].

    (ALCL3 ~

    (A |Br3)

    Fig. 33. Formation of highly symmetrical, less strained molecules, by isomerization: from top to bottom, adamantane, diamantane, dodecahedrane.

  • Carbon and its nets 415

    10. C O N C L U S I O N

    We have discussed the two known allotropes of elemental carbon (graphite and diamond), the presumed linear carbyne, and other hypothetical finite (molecules) and infinite C forms (allotropes). Despite the higher energy of all these forms than the energy of graphite, some of these nets may have high kinetic barriers for isomerization to graphite, so that they may be able of existence.

    We have also examined related nets with other atoms instead of C. Portions of graphite and diamond lattice, where the marginal C atoms are bonded to H,

    constitute the benzenoid and diamond hydrocarbons, respectively; these were discussed according to their symmetry. Of course, one could also discuss the hydrocarbons which result by considering portions of the other, hypothetical, nets. This has not been done because of space limitations.

    It was pointed our briefly in the preceding section how important for the stability of molecules is their symmetry.

    R E F E R E N C E S

    1, A. T. Balaban, Symmetry in chemical structures and reactions. Comput. Math. Applic. 12B, 999-1020 (1986). Reprinted in Symmetry: Unifying Human Understanding (Ed. I. Hargittai). Pergamon Press, Oxford (1986).

    2. H. Staudinger, l~ber Polymerisation. Bet. dt. chem. Ges. 53, 1073-1085 (1920). 3. V. M. Morawetz, Difficulties in the emergence of the polymer concept--an essay. Angew. Chem. Int. 26, 93-97 (1987)

    and references cited therein. 4. L. D. David, The discovery of soluble polysilane polymers. Chem. Br. 23, 553-556 (1987). 5. J. Donohue, The Structure of the Elements. Wiley, New York (1974). 6. B. T. Kelly, Physics of Graphite. Applied Science Publishers, New Jersey (1981); A. R. Ubbelohde and F. A. Lewis,

    Graphite and Its Crystal Compounds. Oxford Univ. Press, London (1960). 7. M. S. Dresselhaus, G. Dresselhaus, J. E. Fischer and M. J. Moran (Eds), Intercalated Graphite, Materials Research

    Soc. Syrup. Proc., Vol. 20. North-Holland, New York, and Elsevier, Amsterdam (1983); A. S. Whittingham and A. J. Jacobson (Eds), Intercalation Chemistry. Academic Press, New York (1982); M. E. Volpin and Yu. N. Novikov, Graphite as an aromatic ligand, Topics in Nonbenzenoid Aromatic Chemistry (Eds T. Nozoe et aL). Hirokawa Publ., Tokyo (1972).

    8. D. Hanson, D. O, E.'s Hanford reactor to keep operating. Chem. Engng News, 65 (8), 23 (1987). 9. M. C. Reisch, High-performance fibers find expanding military, industrial uses. Chem. Engng News 65, (5), 9-14 (1987);

    E. Fitzer (Ed.), Carbon Fibres and Their Composites, Springer, Berlin (1985). 10. J. E. Field (Ed.), The Properties of Diamond. Academic Press, New York (1979). 11. F. P. Bundy. Superhard materials. Sci. Am. 231(2), 62-70 (1974); Diamond synthesis with nonconventional

    catalyst-solvent, Nature, Lond. 241, 116-118 (1973); The p,T phase and reaction diagram for elemental carbon. J. Geophys. Res. 85, 6930-6936 (1980); Direct conversion of graphite to diamond on static pressure apparatus. J. Chem. Phys. 38, 631-643 (1963); F. P. Bundy, H. M. Strong and R. H. Wentorf Jr, Methods and mechanisms of synthetic diamond growth. Chem. Phys. Carbon 10, 213-263 (1973); F. Bundy and J. S. Kasper, Hexagonal diamond, a new form of carbon. J. Chem. Phys. 46, 3437-3446 (1967).

    12. V. A. Zhorin et aL, Structural changes in graphite induced by the combined action of high pressures and shear deformation. Zh. fiz. Khim. 56, 2486-2490 (1982); Formation of a sp 3 hybridized state in graphite by simultaneous exposure to high pressure and shear deformation, Dokl. Akad. Nauk SSSR 261, 665-668 (1981).

    13. R. H. Wentorf Jr, Solutions of carbon at high pressure. Ber. Bunsenges. Phys. Chem. 70, 975-982 (1966); H. M. Strong and R. M. Chrenko, Diamond growth rates and physical properties of laboratory-made diamond. J. Phys. Chem. 75, 1838-1843 (1971).

    14. R. S. Lewis, T. Ming, J. F. Wacker, E. Anders and E. Steel, Interstellar diamonds in meteorites. Nature, Lond. 326, 160-162 (1987).

    15. C. A. Cupas and L. Hodakowski, Iceane. J. Am. chem. Soc. 96, 4668-4669 (1974). 16. H. W. Kroto, J. R. Heath, S. C. O'Brien, R. F. Curl and R. E. Smalley, C~0. Buckminsterfullerene. Nature, Lond. 318,

    162-163 (1985); J. R. Heath, S. C. O'Brien, Q. Zhang, Y. Liu, R. F. Curl, H. W. Kroto, F. K. Tittel and R. E. Smalley, Lanthanum complexes of spheroidal carbon shells, J. Am. chem. Soc. 107, 7779-7780 (1985); Q. L. Zhang, S. C. O'Brien, J. R. Heath, Y. Liu, R. F. Curl, H. W. Kroto and R. E. Smalley, Reactivity of carbon clusters: spheroidal carbon shells and their possible relevance to the formation and morphology of soot, J. Phys. Chem. 90, 525-528 (1986).

    17. A. D. J. Haymet, Archimedane. Chem. Phys. Lett. 122, 421-424 (1985). 18. A. T. Balaban, On a 3-connected carbon net (infinite tridimensional lattice of sp2-hybridized carbon atoms) and

    congeneric systems. Rev. Roum. Chim, 33, 359-362 (1988). 19. D. J. Klein, W. A. Seitz and T. G. Schmalz, Icosahedral symmetry carbon cage molecules. Nature, Lond. 323, 703-706

    (1986). 20. A. T. Balaban, Is aromaticity outmoded? Pure AppL Chem. 52, 1409-1429 (1980). 21. A. M. Sladkov and Yu. P. Kudryavtsev, Polyacetylenes. Usp, Khim. 32, 509-538 (1963); V. M. Melnichenko,

    A. M. Sladkov and Yu. M. Mikulin, Structure of polymeric carbon. Usp. Khim. 51, 736-763 (1982). 22. V. M. Melnichenko, Yu. I. Nikulin and A. M. Sladkov, Layer-chain carbon. Dokl. Akad, Nauk SSSR 267, 1150-1154

    (1982). 23. M. Randir, Aromaticity and conjugation. J. Am. chem. Soc. 99, 444-450 (1977). 24. N. Trinajsti~, Chemical Graph Theory, Vol. 2, Chap. 3. CRC Press, Boca Raton, Fla. (1983). 25. J. R. Platt, Atomic arrangements and bonding across a twinning plane in graphite, Z. Crystallogr. 109, 226-230 (1957).

  • 416 A.T. BALAaAN"

    26. J. R. Dias, Enumeration of graphite carbon-bond network defects having ring sizes ranging from 3 to 9. Carbon 22, 107-I 14 (1984).

    27. A. T. Balaban, C. C. Renl~ia and E. Ciupitu, Chemical graphs. Part 6. Estimation of the relative stability of several planar and tridimensional lattices for elementary carbon. Rev. Roum. Chim. 13, 231-247 (1968); erratum, p. 1233.

    28. J. Barriol and J. Metzger, Application de la mtthode des orbitales mol~-ulaires au r~seau du graphite. J. chim. phys. 47, 432-436 (1950).

    29. J. K. Burdett and S. Lee, The moments method and elemental structures. J. Am. chem. Soc. 107, 3063-3082 (1985); see also J. K. Burdett, Some structural problems examined using the method of moments. Structure Bonding 65, 29-90 (1987); J. K. Burdett, Topological control of the structures of molecules and solids. Acc. Chem. Res. 21, 189-194 (1988).

    30. R. H. Riley, Chemical and crystallographic factors in carbon combustion. J. Chim. phys. 47, 565-572 (1950). 31. R. Hoffmann, T. Hughbanks, M. Kert6sz and P, H. Bird, a Hypothetical metallic allotrope of carbon. J. Am. chem.

    Soc. 105, 4831-4832 (1983). 32. M. Kert6sz and R. Hoffmann, The graphite to diamond transformation. J. Solid State Chem. 54, 313-319 (1984). 33. I. V. Stankevich, M. V. Nikerov and D. A. Bochvar, Structural chemistry of crystalline carbon: geometry, stability,

    electronic spectrum. Usp. Khim. 53, 1101-1024 (1984); Russ. chem. Revs 53, 640 (1984); A. F. Wells, The Third Dimension in Chemistry. Clarendon Press, Oxford (1956); Three-Dimensional Nets and Polyhedra. Wiley-Interscience, New York (1977).

    34. K. M. Merz Jr, R. Hoffmann and A. T. Balaban, 3,4-Connected carbon nets: through-space and through-bond interactions in the solid state. J. Am. chem. Soc. 109, 6742-6751 (1987).

    35. R. Hoffmann, O. Eisenstein and A. T. Balaban, Hypothetical strain-free oligoradical. Proc, natn. Acad. Sci. 77, 5588-5592 (1980).

    36. R. H. Wentorf Jr., Cubic form of boron nitride, J. Chem. Phys. 26, 956 (1957). 37. A. T. Balaban and F. Harary, Chemical graphs. Part 5. Enumeration and proposed nomenclature of benzenoid

    cata-condensed polycyclic aromatic hydrocarbons, Tetrahedron. 24, 2505-2516 (1968); A. T. Balaban, Chemical graphs. Part 7. Proposed nomenclature of branched cata-condensed benzenoid hydrocarbons. Tetrahedron 25, 2949-2956 (1969); Challenging problems involving benzenoid polycyclics and related systems. Pure appl. Chem. 54, 1075-1096 (1982).

    38. J. R. Dias, A periodic table for polycyclic aromatic hydrocarbons, Part 1. Isomer enumeration of fused polycyclic aromatic hydrocarbons. J. Chem. Inf. Comput. Sci. 22, 15-22 (1982); Part 2. Polycyclic aromatic hydrocarbons containing tetragonal, pentagonal, heptagonal and octagonal rings. J. Chem. Inf. Comput. Sci. 22, 139-152 (1982); Part 3, Enumeration of all the polycyclic conjugated isomers of pyrene having ring sizes ranging from 3 to 9. Math. Chem. 14, 83-138 (1983); Handbook of Polycyclic Hydrocarbons. Part A. Benzenoid Hydrocarbons. Elsevier, Amsterdam, 1987.

    39. A. T. Balaban, Enumeration of catafusenes, diamondoid hydrocarbons and staggered alkane C-rotamers. Math. Chem. 2, 51-61 (1976).

    40. M. Gordon and W. H. T. Davison, Theory of resonance topology of fully aromatic hydrocarbons. I. J. Chem. Phys. 20, 428-435 (1952); S. J. Cyvin and I. Gutman, Number of Kekul6 structures as a function of the number of hexagons in benzenoid hydrocarbons. Z. Naturf. 41a, 1079-1086 (1986); I. Gutman and O. E. Polansky, Mathematical Concepts in Organic Chemistry, p. 59 Springer, Berlin (1986).

    41. J. V. Knop, W. R. Miiller, K. Szymanski and N. Trinajsti~, Computer Generation of Certain Classes of Molecules, p. 109. Kemija u industriji, Zagreb (1985).

    42. A.T. Balaban and I. Tomescu, Algebraic expressions for the number of Kekul6 structures in isoarithmic cata-condensed benzenoid polycyclic hydrocarbons. Math. Chem. 14, 155-182 (1983); Chemical graphs. Part 41, Numbers of conjugated circuits and Kekul6 structures for zig-rag catafusenes and(j, k)-hexes; generalized Fibonacei numbers. Math. Chem. 17, 91-120 (1985); Chemical graphs. Part 40. Three relations between the Fibonacci sequence and the numbers of Kekul6 structures for non-branched cata-condensed polycy¢lic aromatic hydrocarbons. Croat. chem. Acta 54, 391-404 (1984); A. T. Balaban, C. Artemi and I. Tomeseu, Algebraic expressions for Kekul6 structure counts in non-branched regularly cata-condensed benzenoid hydrocarbons. Math. Chem. 22, 77-100 (1987); A. T. Balaban, J. Brunvoll, J. Cioslowski, B. N. Cyvin, S. J. Cyvin, I. Gutman, He Wenchen, He Wenjie, J. V. Knop, M. Kovacevi~, W. R. M~ller, K. Szymanski, R. Tosi6 and N. Trinajsti6, Enumeration of benzenoid and coronoid hydrocarbons Z. Naturf. 4 ~ , 863-870 (1987); A. T. Balaban, J. BrunvoU, B. N. Cyvin and S. J. Cyvin, Enumeration of branched cata-condensed benzenoid hydrocarbons and their numbers of Kekul~ structures. Tetrahedron 44, 221-228 (1988).

    43. S. E. Stein and R. L. Brown, n-Electron properties of large condensed polyaromatic hydrocarbons. J. Am. chem. Soc. 109, 3721-3729 (1987).

    44. G. Wittig, Stereochemie, p. 148, Akademische Verlagsgesellschaft, Leipzig (1930). 45. R. Lukes, Organick~ Chemie, Vol. 1, p. 59. Naklad. Ceskoslovensk. Akad. Ved, Praha (1954). 46. A. T. Balaban and P. von R. Schleyer, Systematic classification and nomenclature of diamond hydrocarbons. I.

    Graph-theoretical enumera~on of polymantanes. Tetrahedron 34, 3599-3609 (1978). 47. A. T. Balaban, Chemical graphs. Part 27. Enumeration and codification of staggered conformation of alkanes. Rev.

    Roum. Chim. 21, 1339-1343 (1976). 48. A. T. Balaban, Synthesis of pyrylotrimethinecyanines. Tetrahedron Lett. 599-600 (1978). 49. P. yon R. Schleyer, A simple preparation of adamantane. J. Am. chem. Soc. 79, 3292 (1957); C. Cupas, P. yon R.

    Schleyer and D. J. Trecker, Congressane. J. Am. chem. Soc. 87, 917-918 (1965). 50. V. Z. Williams Jr., P. von R. Schleyer, G. J. Gleicher and L. B. Rodewald, Triamantane. J. Am. chem. Soc. 88,

    3862-3863 (1966); W. Burns, M. A. McKervey, T. R. B. Mitchell and J. J. Rooney, A new approach to the construction of diamondoid hydrocarbons. Synthesis of anti-tetramantane. J. Am. chem. Soc. 100, 906-911 (1978).

    51. W. D. Fessner, B. A. R: C. Murty, J. W6rth, D. Hunkler, H. Fritz, H. Prinzbach, W. D. Roth, P. von R. Schleyer, A. B. McEwen and W. F. Maier, Dodecahedranes from [1.1.1.1] pagodanes. Angew. chem. Int. 26, 452-454 (1987).

    52. S. H/da, S. Landa and V, Hanus, Isolierung von Tetracyclo[6.3.1.0~6.05a°]dodecan und Pentacyclo- [7.3.1.14,n.02.7.06,n]tetradecan (Diamantan) aus Erd61. Angew. Chem. 78, 1060-1061 (1966).