You are on page 1of 9

Article

Cite This: J. Chem. Educ. 2019, 96, 2467−2475 pubs.acs.org/jchemeduc

Representation of Three-Center−Two-Electron Bonds in Covalent


Molecules with Bridging Hydrogen Atoms
Gerard Parkin*
Department of Chemistry, Columbia University, New York, New York 10027, United States
*
S Supporting Information

ABSTRACT: Many compounds can be represented well in


terms of the two-center−two-electron (2c−2e) bond model.
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

However, it is well-known that this approach has limitations;


for example, certain compounds require the use of three-
center−two-electron (3c−2e) bonds to provide an adequate
Downloaded via UNIV DE ALICANTE on October 14, 2022 at 17:53:06 (UTC).

description of the bonding. Although a classic example of a


compound that features a 3c−2e bond is provided by
diborane, B2H6, 3c−2e interactions also feature prominently in transition metal chemistry, as exemplified by bridging hydride
compounds, agostic compounds, dihydrogen complexes, and hydrocarbon and silane σ-complexes. In addition to being able to
identify the different types of bonds (2c−2e and 3c−2e) present in a molecule, it is essential to be able to utilize these models to
evaluate the chemical reasonableness of a molecule by applying the octet and 18-electron rules; however, to do so requires
determination of the electron counts of atoms in molecules. Although this is easily achieved for molecules that possess only 2c−
2e bonds, the situation is more complex for those that possess 3c−2e bonds. Therefore, this article describes a convenient
approach for representing 3c−2e interactions in a manner that facilitates the electron counting procedure for such compounds.
In particular, specific attention is devoted to the use of the half-arrow formalism to represent 3c−2e interactions in compounds
with bridging hydrogen atoms.
KEYWORDS: Second-Year Undergraduate, Graduate Education/Research, Inorganic Chemistry, Analogies/Transfer,
Problem Solving/Decision Making, Covalent Bonding, Lewis Structures, Main-Group Elements, Organometallics, Transition Elements

■ INTRODUCTION
The octet1,2 and 18-electron2a,3 rules, together with the
deficient”5,7 to indicate that there is an insufficient number of
electrons for all bonds to be 2c−2e.
construction of so-called “Lewis structures”,4 are of pivotal The structures of so-called “electron deficient” compounds
importance because they provide a rudimentary means to may, however, be rationalized by the use of multicenter
evaluate the chemical reasonableness of a molecule. Central to bonding descriptions, of which the three-center−two-electron
the application of these rules is the necessity of determining (3c−2e) model is the simplest.8−10 In view of this explanation
the electron count (or electron number, EN) of the atoms in
the molecule of interest. For many molecules (e.g., H2 and
CH4), the bonding can be described well in terms of a two-
center−two-electron (2c−2e) model, and the electron counts
of atoms in such molecules are typically easily evaluated from
the Lewis structure. In contrast, the bonding in some
molecules cannot be described by using only 2c−2e bonds,
as exemplified by those with bridging hydrogen atoms, of
which diborane, B2H6 (Figure 1),5,6 is an excellent example.
For example, whereas ethane, C2H6, has a total of 14 valence
Figure 1. Comparison of the structures of C2H6 and B2H6. C2H6 has
electrons derived from two carbon atoms and six hydrogen
14 valence electrons, which is sufficient to form seven 2c−2e bonds,
atoms (i.e., (2 × 4) + (6 × 1)) and can hence possess seven whereas B2H6 has only 12 valence electrons and can only form a
2c−2e bonds to give rise to the familiar structure shown in maximum of six 2c−2e bonds; therefore, B2H6 cannot adopt the same
Figure 1, B2H6 has a total of only 12 valence electrons (i.e., (2 structure as C2H6, and multicenter bonding is required to rationalize
× 3) + (6 × 1)) and can therefore only support a maximum of its structure. The dotted lines connecting the bridging hydrogen and
six 2c−2e bonds; as such, the number of electrons is boron atoms are intended to indicate that these are not 2c−2e bonds.
insufficient for B2H6 to adopt a structure akin to that of
C2H6. Therefore, B2H6 adopts a different type of structure in Received: August 10, 2019
which two of the hydrogen atoms bridge the two boron centers. Revised: September 7, 2019
Compounds of this type are often referred to as “electron Published: October 9, 2019
© 2019 American Chemical Society and
Division of Chemical Education, Inc. 2467 DOI: 10.1021/acs.jchemed.9b00750
J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

Figure 4. Two alternative representations for a dative bond, as


illustrated for H3NBH3. Despite the different appearances, the
representations are not resonance structures because the electrons
are in the same location (recall that resonance structures differ in the
locations of the electrons).

■ REVIEW OF THE COVALENT BOND


CLASSIFICATION AND THE REPRESENTATION OF
2C−2E BONDS
Prior to discussing 3c−2e interactions, it is first pertinent to
review the representation of 2c−2e bonds. The molecular
Figure 2. Molecular orbital diagram for H2. The bonding orbital (σ) orbital description of a 2c−2e bond is straightforward: the
is occupied by two electrons, so each hydrogen is assigned a duet
overlap of two atomic orbitals on adjacent atoms results in the
configuration. The + and − symbols on the orbitals refer to their
relative phases, as do their colors. formation of bonding and antibonding orbitals, of which the
former is occupied by two electrons, as illustrated by H2
(Figure 2). However, although this is a common feature of all
of the bonding, Rundle once quipped, “There are no such 2c−2e bonds, there are two fundamentally different types of
things as electron deficient compounds, only theory deficient covalent bond that are differentiated by the number of
chemists.”11,12 electrons that each partner contributes to the interaction
However, although the bonding in “electron deficient” (Figure 3). Thus, a normal covalent bond is one in which each
compounds such as B2H6 may be understood by the neutral partner contributes one electron to the bonding
application of theoretical methods, this approach lacks the orbital,14 whereas a dative covalent (or coordinate) bond is
convenience of using simple electron counting procedures to one in which a single atom provides both electrons.15
evaluate the chemical reasonableness of a covalent molecule. Common examples of species that provide one electron to a
Unfortunately, the electron counts for atoms in molecules with bond are H, Cl and CH3 (i.e., radicals), whereas species that
3c−2e interactions are not as easily determined as those with provide two electrons are illustrated by H2O, NH3, PR3, and
2c−2e interactions because (i) the structural representations CO (i.e., neutral closed shell molecules with lone pairs).
employed often lack clarity, and (ii) there is also the need to Normal covalent and dative covalent bonds are represented
evaluate the electron count of more than one atom. The by different symbols. Thus, a normal covalent bond is depicted
purpose of this article is, therefore, to describe how to as a line between two atoms, whereas a dative covalent bond is
represent the bonding in molecules with 3c−2e bonds in such represented as an arrow with the head pointing to the acceptor
a manner that it is simple to evaluate the electron count of atom (Figure 4). The latter can also be denoted by an
each atom that is involved in the interaction, and thereby equivalent form that uses a line but also depicts formal charges
utilize effectively the octet and 18-electron rules for classes of on the atoms, as illustrated for H3NBH3 (Figure 4).16,17 It
molecules such as bridging hydride, dihydrogen, and agostic must be emphasized that the two depictions of the dative
complexes, which are often encountered in undergraduate covalent bond are not resonance structures because the
inorganic chemistry courses.13 electrons are in the same location (i.e., between N and B;

Figure 3. Covalent bond classification (CBC) of L, X, and Z ligands. Each partner provides one electron for a normal covalent bond (b), whereas a
single partner provides both electrons for a dative covalent bond (a, c).

2468 DOI: 10.1021/acs.jchemed.9b00750


J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

Figure 4); as such, the two depictions are just different dative covalent bond is provided by the ammonia−borane
representations of the same bonding description, and the form adduct, H3N→BH3, in which the boron is classified as MLX3
that is used is often just a matter of convenience. and the nitrogen is classified as MX3Z. Thus, both the boron
The covalent bond classification18 provides a means to (i.e., EN = 3 + 2 + (3 × 1) = 8) and nitrogen (i.e., EN = 5 + (3
categorize covalent molecules and, for this approach, ligands × 1) = 8) possess octet configurations. Alternatively, a
are classified according to the number of electrons that they procedure that is often more convenient for compounds with
contribute to the bonding orbital. Thus, a neutral two-electron multidentate ligands is to determine the electron count by the
donor is classified as L, a one-electron donor is X, and a zero- expression EN = m + Σ(ligand electrons) − Q±, where
electron donor is Z (Figure 3). Σ(ligand electrons) is the sum of the contributions of
Many ligands possess more than one L, X or Z function, or individual ligands. For example, the electron count for
combination thereof, and in such cases, the ligand is classified Cp2TaH3 is EN = 5 + (2 × 5) + (3 × 1) = 18; likewise,
as [LlXxZz], where l, x, and z are the respective numbers of L, Cp2W(CO) also has an 18-electron configuration (Figure 7).
X, and Z functionalities. For example, an η6-benzene ligand is
classified as L3 on the basis that each of the “double bonds”
may be conceptually viewed as an L donor akin to that of C2H4
(Figure 5). Correspondingly, an η5-cyclopentadienyl ligand is

Figure 7. Alternative procedure for determining the electron counts


(EN) for atoms in some compounds with multidentate ligands (EN =
m + Σ(ligand electrons) − Q±).

Figure 5. [LlXxZz] classifications of two common multidentate


ligands. A cyclopentadienyl ligand is classified as L2X because the
Lewis structure contains two double bonds (L) and an alkyl moiety MOLECULAR ORBITAL DESCRIPTION OF 3C−2E
(X), whereas a benzene ligand is classified as L3. INTERACTIONS AND STRUCTURE−BONDING
REPRESENTATIONS
classified as L2X on the basis that the cyclopentadienyl radical The simplest molecule that possesses a 3c−2e interaction is
may be conceptually considered to be composed of two [H3]+, which has been spectroscopically observed but not
“double bonds” (each of which is an L donor) and an alkyl isolated. Both linear and triangular structures may be
radical (which is an X donor). In terms of overall electron considered, and the molecular orbitals for the two forms,
count, cyclopentadienyl and benzene are respectively consid- which are related by an orbital correlation diagram, are
ered to be five-electron and six-electron donors (Figure 5). illustrated in Figure 8.20 Since the system contains a single pair
The electron count about an atom (M) in a molecule of the of electrons, the preferred structure is determined only by the
type [MLlXxZz]Q±, where l, x, and z are the respective numbers occupancy of the fully bonding orbital (i.e., all orbitals in
of L, X, and Z functions, and Q± is the charge on the molecule, phase), which is lowest in energy for the triangular form
is, therefore, given by the expression EN = m + 2l + x − Q±, because it maximizes the overlap of the three hydrogen 1s
where m is the number of valence electrons on a neutral M
atom. For the transition metals in groups 3−11,19 m
corresponds to the Periodic Table group number, (e.g., Sc,
3; Ti, 4; V, 5; etc.), whereas for the main group elements in
groups 12−18, the value of m is the last digit.
As an illustration of the determination of the electron count,
the carbon atom in CH4 is classified as MX4 and therefore
possesses an octet configuration (i.e., EN = 4 + (4 × 1) = 8,
Figure 6). A simple example of a molecule that possesses a

Figure 8. Molecular orbitals for linear and triangular [H3]+. Each


Figure 6. Procedure for determining the electron counts (EN) for hydrogen possesses a duet configuration. The + and − symbols on the
atoms in some simple compounds (EN = m + 2l + x − Q±). orbitals refer to their relative phases, as do their colors.

2469 DOI: 10.1021/acs.jchemed.9b00750


J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

represented by the structures shown in Figure 9. Evaluation of


each hydrogen atom in turn illustrates how the duet
configuration may be derived. Thus, HA and HB have a duet
configuration because they share a pair of electrons, whereas
HC has a duet configuration because HC+ (with no electrons)
shares the pair of electrons in the HA−HB bond.
This approach for representing a 3c−2e interaction can also
be used if the three atoms are not part of an equilateral triangle
such that the interactions between all pairs of atoms are not
equal. However, an alternative way to represent this bonding
situation is to employ the so-called “half-arrow” notation that
Figure 9. (a) Connectivity and (b) three equivalent structure− was introduced in 1980 by Green for bridging hydride
bonding representations of [H3]+. compounds (Figure 10).21,22 With this notation, a half-arrow
is drawn from a central bridging hydrogen atom to an outer
atom. This representation is very similar to the well-known use
of a full-arrow for bridging halide compounds (e.g., Al2Cl6) to
indicate the donation of a lone pair of electrons (Figure 10);
the purpose of using a half-arrow (rather than a full-arrow) for
the bridging hydrogen atom is to emphasize that the arrow
does not correspond to donation of a lone pair but rather
donation of a bond pair.22 Otherwise, there is no difference in
the impact of a half-arrow or full-arrow on the electron count
of the outer atom.
Figure 10. Half-arrow representation for a single 3c−2e interaction
It is important to emphasize that the two representations for
for a bridging hydride compound (top) and comparison with the
representation for a bridging chloride ligand that has two 2c−2e the 3c−2e interaction (i.e., a full-arrow from the center of the
bonds because of the availability of lone pairs on chlorine (bottom). bond and a half-arrow from the central atom) are intended to
convey the same information with respect to electron counting
purposes, and the form that is used is simply a matter of
orbitals. Since the pair of electrons in the bonding orbital of
convenience, similar to the manner in which a dative covalent
[H3]+ is associated with all of the hydrogen atoms, each
hydrogen atom possesses a duet configuration; this situation is bond can represented as either A→B or A+−B−. For example,
analogous to both hydrogen atoms in H2 being assigned a duet the full-arrow from the center of the bond is most commonly
configuration because they share a pair of electrons (Figure 2). employed if the angle at the central atom is distinctly nonlinear
Thus, with respect to electron counting, an important feature and the two “outer” atoms are considered to be within a
of a 3c−2e interaction is that each atom is assigned a shared bonding distance (i.e., a “closed” system23), whereas the half-
pair of electrons. It should be noted that the same concept arrow representation is used if the angle at the central atom is
would apply if the linear form of [H3]+ (or a bent form with an significantly obtuse and there is little interaction between the
H−H−H angle not equal to 60°) were to be more stable (i.e., outer atoms.24
each atom would still be assigned a duet configuration).
Rather than having to construct an MO diagram to
determine the electron counts for atoms involved in a 3c−2e
■ EXAMPLES OF 3C−2E INTERACTIONS
Molecules that feature 3c−2e interactions are rather common,
interaction, however, it is more convenient to use structure− and the application of the above structure−bonding
bonding figures that allow one to evaluate the electron counts. representations to these molecules is illustrated below by
Consider, for example, [H3]+ being derived from two hydrogen consideration of three classes of compounds, namely, (i)
atoms (HA and HB) and a proton (HC+), as illustrated in the bridging hydride compounds, (ii) agostic compounds, and (iii)
center structure of Figure 9. In valence bond terms, the [H3]+ dihydrogen compounds.13
molecule can be viewed as an adduct of H2 and H+, which
differs from a conventional adduct in the sense that the donor Bridging Hydride Compounds
pair of electrons is derived from an H−H bond rather than a Bridging hydride compounds constitute the largest class of
lone pair on a single atom. The bonding can, therefore, be molecules with 3c−2e interactions, as exemplified by B2H6, as

Figure 11. Representations of the 3c−2e B−H−B interaction in diborane. Coordination of BH3 by a B−H bond of another molecule allows boron
to achieve an octet configuration. The 3c−2e interaction can be represented by either an arrow from the center of the B−H bond or an equivalent
form in which a half-arrow is drawn from the hydrogen atom.

2470 DOI: 10.1021/acs.jchemed.9b00750


J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

means to determine whether any additional direct bonding


exists between the outer A atoms. For example, if A was a
transition metal, an apparent electron count of EN would
indicate that each A atom would require an additional (18 −
Figure 12. Often encountered incorrect representation of B2H6 that EN) A−A bonds in order to achieve an 18-electron
appears to depict two 2c−2e bonds to hydrogen. configuration, because each A−A bond would increase the
electron count by one. As an illustration, the two rhenium
noted above. The bonding in B2H6 may be considered in terms atoms of [CpReH2]2(μ-H)2 would possess 17-electron
of its formation by dimerization of BH3. Thus, whereas the configurations, and so a direct Re−Re bond is required to
boron atom in BH3 has only a sextet configuration, achieve an 18-electron configuration (Figure 14). In contrast,
coordination by the pair of electrons in a B−H bond of a [(CO)4Re]2(μ-H)2 possesses an 18-electron configuration for
second molecule results in an octet configuration (Figure 11). each rhenium atom, and so no additional Re−Re bond is
The half-arrow representation provides a very convenient required (Figure 14).
means to evaluate the electron count of each boron center. It A good illustration of the use of the half-arrow to represent
also provides a clear depiction of the 3c−2e nature of the the structure of bridging hydride compounds is provided by
interaction, in contrast to representations that simply draw two Os3(CO)10(μ-H)2, which has been depicted in the literature by
lines from each hydrogen atom,25 which can thereby give the a variety of different structures that have ambiguous bonding
misleading impression that hydrogen is forming two normal descriptions, with Os−Os bond orders that vary from zero to
covalent bonds (Figure 12). two (Figure 15). The ambiguity is also highlighted by the fact
Transition metal compounds also feature bridging hydride
that some representations simply utilize ill-defined dotted lines
ligands, and some examples are presented in Figures 13 and
to represent the bonding. The half-arrow representation,
14.26 For example, [(CO)5W(μ-H)W(CO)5]− can be depicted
by the two structure−bonding representations shown in Figure however, clearly indicates that each osmium center associated
13, in which both tungsten centers possess 18-electron with the bridging hydrogens adopts an 18-electron config-
configurations. uration without requiring an additional Os−Os bond.
Significantly, this description is in accord with theoretical
calculations, which indicate that there is no OsOs double
bond and that the interaction is best represented as two 3c−2e
interactions.27
With respect to the origin of the different bonding
representations of Os3(CO)10(μ-H)2 in Figure 15, it is
pertinent to note that the derivation of an OsOs double
bond is a consequence of essentially apportioning only “half of
an electron” of a bridging hydrogen atom to each metal
center.28 The problem with this approach can be readily
Figure 13. Half-arrow bonding representations for [(CO)5W(μ- ascertained by recognizing that this method would also predict
H)W(CO)5]−. The 3c−2e interaction allows each tungsten center to a BB double bond in B2H6 in order for boron to achieve an
possess an 18-electron configuration without requiring an M−M octet configuration, which is clearly not feasible. More
bond. generally, this approach predicts the number of metal−metal
bonds (m#) for transition metals by using the formula m# =
(18n − N)/2, where n is the number of M atoms, and N is the
total valence electron count for the molecule. However, this
simple formula does not take into account the 3c−2e nature of
bridging hydride interactions, and so the predicted M−M bond
order is greater than that determined by computational
analysis; in fact, the bond order predicted by the formula
does not actually depend on whether the hydrogen atoms are
terminal or bridged, which is clearly not a satisfactory situation.
In contrast, the half-arrow representation provides a bonding
description that is much more in accord with that predicted
theoretically.
In addition to the above compounds that feature hydrogen
bridging between two boron centers and between two
Figure 14. Half-arrow bonding representations for [(CO)4Re]2(μ- transition metals centers, hybrid compounds are known in
H)2 and [CpReH2]2(μ-H)2. Whereas the two 3c−2e interactions in which hydrogen bridges boron and a metal. For example, there
[(CO)4Re]2(μ-H)2 allow the rhenium centers to achieve an 18- is a large variety of transition metal borohydride compounds in
electron configuration, [CpReH2]2(μ-H)2 achieves only a 17-electron which the metal and boron are bridged by one (κ1), two (κ2),
configuration; therefore, a Re−Re bond is required to obtain an 18- or three (κ3) hydrogen atoms (Figure 16).29 The bonding in
electron configuration for the latter. these compounds may likewise be described in terms of the
half-arrow representations, which provides a convenient means
One aspect of evaluating the electron count of atoms to evaluate the electron count of a metal center, as illustrated
involved in an A−H−A 3c−2e interaction is that it provides a for Cp2Nb(κ2-H2BH2) in Figure 17.
2471 DOI: 10.1021/acs.jchemed.9b00750
J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

Figure 15. Structural representations for Os3(CO)10(μ2-H)2. Only E and F provide good representations of the bonding.

interaction is provided by the observation that the titanium


ethyl compound, (dmpe)TiEtCl3, exhibits a very unusual acute
Ti−C−C bond angle of 85.9° because of the interaction of the
titanium with the β-C−H bond. This interaction is a
consequence of the titanium center of (dmpe)TiEtCl3 being
Figure 16. Coordination modes of metal borohydride compounds. highly electron deficient (with a formal electron count of 12 in
the absence of an interaction), and the electronic unsaturation
is partially relieved by interaction of the titanium center with
the electron density associated with a C−H bond.
Dihydrogen Compounds and Other σ-Complexes
Dihydrogen compounds compose an extremely interesting
class of molecules, first discovered by Kubas,31 in which the
H−H bond of H2 remains intact upon coordination to a metal
center, [M](η2-H2), as illustrated in Figure 19. Prior to Kubas’s
Figure 17. Example of a κ2-borohydride complex. discovery, the only known interaction between H2 and a metal
center was oxidative addition resulting in a dihydride
Agostic Compounds compound in which the H−H bond has been cleaved,
An interesting class of compounds is one in which there is an [M]H2. The 3c−2e interaction can be viewed as donation of
interaction between a C−H bond and a metal center. Such the pair of electrons associated with the H−H bond into an
compounds, and their interactions, are referred to by the term empty orbital on the metal center (cf. [H3]+); for this reason,
“agostic”.13,30 Agostic interactions are recognized to be an dihydrogen compounds are typically represented by drawing
important feature of organometallic chemistry, and they can an arrow from the center of the H−H bond to the metal center
play important roles in modifying the structure and stability of (Figure 20).
a molecule in which the metal center would otherwise be The ability to isolate a dihydrogen compound, rather than a
electron deficient. Classic examples of compounds that possess dihydride derivative, requires that there be little backbonding
agostic interactions are provided by the bis-
from the metal center into the H−H σ* orbital, because such
(dimethylphosphino)ethane titanium alkyl compounds
transfer of electron density promotes cleavage of the H−H
(dmpe)TiMeCl3 and (dmpe)TiEtCl3, in which the titanium
centers interact, respectively, with α- and β-C−H bonds bond. For this reason, dihydrogen compounds are typically
(Figure 18).30 The significance of the agostic C−H−M associated with metal centers that have ligands that remove
electron density via backbonding (e.g., CO).32

Figure 18. Examples of compounds that feature α- and β-agostic Figure 19. Dihydrogen and dihydride tautomers. The former
interactions. Coordination of the C−H bond to Ti increases the possesses one 3c−2e interaction, whereas the latter possesses two
electron count by two. 2c−2e interactions.

2472 DOI: 10.1021/acs.jchemed.9b00750


J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

Figure 20. Examples of transition metal dihydrogen complexes. The


dihydrogen molecule donates its bonding electrons to the metal to Figure 22. Example of a bridging methyl compound with 3c−2e
result in 18-electron configurations. interactions.

Dihydrogen complexes are a subset of a class of molecules


referred to as σ-complexes, so named because the compounds
compounds, which include bridging hydride and agostic
are considered to be derived by donation of the electron
complexes, and furthermore provides a means to evaluate
density of an X−H σ-bond to a metal. Specific examples are
metal−metal bond orders in bridging hydride compounds.


provided by hydrocarbon33 and silane34 adducts, in which
C−H and Si−H bonds interact with a metal center. However, ASSOCIATED CONTENT
such compounds are much less commonly encountered than
dihydrogen complexes. Alkane σ-complexes are particularly *
S Supporting Information

unstable but are often invoked as intermediates in reactions The Supporting Information is available on the ACS
involving the oxidative-addition and reductive-elimination of Publications website at DOI: 10.1021/acs.jchemed.9b00750.
C−H bonds. By comparison, silane σ-complexes are more Worksheets to help students practice the concepts
stable than alkane σ-complexes, as illustrated by (PDF)
Mo(dppe)2(η2-HSiH2Ph), which has been isolated in the Worksheet answers and comments for instructors (PDF)


solid state. Since the 3c−2e interactions are asymmetric, with
longer M−C and M−Si bonds than M−H bonds, these AUTHOR INFORMATION
compounds are frequently represented by using either the half- Corresponding Author
arrow or full-arrow representations, although both are intended
to convey the same information with respect to electron *E-mail: parkin@columbia.edu.
counting purposes (Figure 21). ORCID
Gerard Parkin: 0000-0003-1925-0547
Notes
The author declares no competing financial interest.

■ ACKNOWLEDGMENTS
The National Science Foundation (CHE-1465095) is thanked
for support.

Figure 21. Two alternative representations of silane σ-complexes,


both of which convey the same information with respect to the
■ REFERENCES
(1) (a) Lewis, G. N. Valence and The Structure of Atoms and
purpose of electron counting. Molecules; The Chemical Catalog Company, 1923. (b) Lewis, G. N.
The atom and the molecule. J. Chem. Phys. 1933, 1, 17−28.
(2) (a) Langmuir, I. Types of valence. Science 1921, 54, 59−67.
Other Examples of 3c−2e Interactions (b) Langmuir, I. The structure of atoms and the octet theory of
Many examples of compounds with 3c−2e interactions feature valence. Proc. Natl. Acad. Sci. U. S. A. 1919, 5, 252−259.
a bridging hydrogen because the nondirectional spherical (3) (a) Jensen, W. B. The origin of the 18-electron rule. J. Chem.
nature of the 1s orbital facilitates an interaction with more than Educ. 2005, 82, 28−28. (b) Pyykkö, P. Understanding the eighteen-
electron rule. J. Organomet. Chem. 2006, 691, 4336−4340.
one atom. Despite this observation, a bridging hydrogen atom
(c) Rasmussen, S. C. The 18-electron rule and electron counting in
is not a requirement for the existence of a 3c−2e interaction, as transition metal compounds: Theory and application. ChemTexts
illustrated by Al2Me6, which possesses bridging methyl groups 2015, 1, 10. (d) Mingos, D. M. P. Complementary spherical electron
(Figure 22). In addition, although less common, a variety of density model and its implications for the 18 electron rule. J.
other ligands, such as PR3 and MeCN, are also known to Organomet. Chem. 2004, 689, 4420−4436. (e) Mingos, D. M. P.
participate in 3c−2e interactions.8


Development of bonding models based on the Periodic Table. Chimia
2019, 73, 152−164. (f) Landis, C. R.; Weinhold, F. 18-Electron rule
SUMMARY and the 3c/4e hyperbonding saturation limit. J. Comput. Chem. 2016,
In summary, the half-arrow notation provides a means to 37, 237−241.
(4) Gillespie, R. J.; Robinson, E. A. Gilbert N. Lewis and the
represent the structures of molecules with 3c−2e interactions chemical bond: The electron pair and the octet rule from 1916 to the
in such a manner that the electron count of each atom can be present day. J. Comput. Chem. 2007, 28, 87−97.
conveniently evaluated. Therefore, in conjunction with the (5) (a) Wade, K. Electron Deficient Compounds; Springer, 1971.
octet and 18-electron rules, the approach affords a simple (b) Wade, K. Boranes: Rule-breakers become pattern-makers. New
process to evaluate the chemical reasonableness of a molecule. Scientist, June 6, 1974, pp 615−617 (c) Stone, F. G. Electron-deficient
This notation has broad applicability to a variety of classes of compounds. Endeavour 1961, 20, 61−67.

2473 DOI: 10.1021/acs.jchemed.9b00750


J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

(6) Laszlo, P. A diborane story. Angew. Chem., Int. Ed. 2000, 39, (20) Albright, T. A.; Burdett, J. K.; Whangbo, M.-H. Orbital
2071−2072. Interactions in Chemistry, 2nd ed.; John Wiley & Sons, Inc.: New York,
(7) Rundle, R. E. Electron deficient compounds. J. Phys. Chem. 1957, NY, 2013.
61, 45−50. (21) Berry, M.; Cooper, N. J.; Green, M. L. H.; Simpson, S. J.
(8) (a) Green, J. C.; Green, M. L. H.; Parkin, G. The occurrence and Synthesis and reactions of binuclear molybdenocene and tungsteno-
representation of three-centre two-electron bonds in covalent cene derivatives. J. Chem. Soc., Dalton Trans. 1980, 29−41.
inorganic compounds. Chem. Commun. 2012, 48, 11481−11503. (22) Arrows are used in a multitude of ways in chemistry. For
(b) Green, M. L. H.; Parkin, G. The Covalent Bond Classification example, in addition to a dative bond, a straight arrow can be used to
method and its application to compounds that feature three-center indicate a reaction, evolution of a gas, precipitation of a solid, or an
two-electron bonds. Struct. Bonding 2016, 171, 79−140. electron in a molecular orbital diagram. Likewise, a straight, half-
(9) (a) Bridgeman, A. J.; Empson, C. J. Detecting delocalization. barbed arrow may be used as a component of an equilibrium symbol
New J. Chem. 2008, 32, 1359−1367. (b) De Kock, R. L.; Bosma, W. B. or an electron in a molecular orbital diagram. In addition to straight
The three-center, two-electron chemical bond. J. Chem. Educ. 1988, arrows, curved arrows are also used. For example, a curved arrow is
65, 194−197. used to indicate the motion of a pair of electrons in a reaction
(10) Three-center−four-electron interactions are also known. See, mechanism, whereas a curved, half-barbed (fishhook) arrow is used to
for example, Green, M. L. H; Parkin, G. The classification and indicate the motion of a single electron in radical reactions. In each
representation of main group element compounds that feature three- case, the meaning of an arrow must be interpreted in the context of its
center four-electron interactions. Dalton Trans. 2016, 45, 18784− use, but it must be emphasized that the straight half-arrow used to
18795. represent a 3c−2e interaction has a distinctly different meaning from
(11) The source of this quote attributed to Rundle is the opening that of a curved half-arrow. See (a) Alvarez, S. Chemistry: A panoply
comment in Wade’s 1971 book on Electron Def icient Compounds (ref of arrows. Angew. Chem., Int. Ed. 2012, 51, 590−600. (b) Lakshminar-
5a). Interestingly, a very similar comment was attributed to Mulliken ayanan, A. Arrows in chemistry. Resonance 2010, 15, 51−62.
by Burdett in his 1997 book (Burdett, J. K. Chemical Bonds: A Dialog; (23) Bau, R.; Teller, R. G.; Kirtley, S. W.; Koetzle, T. F. Structures of
Wiley, 1997, p 103): “Robert Mulliken said at one time, referring to transition-metal hydride complexes. Acc. Chem. Res. 1979, 12, 176−
the term ‘electron deficient’ molecules, that there was nothing 183.
deficient about the molecules, just the theory.” Neither Wade nor (24) In addition, even if the interaction with the outer atom is
Burdett give the sources for their quotes, and so it is difficult to significant, the half-arrow representation is sometimes preferable
ascertain the originator. However, in private correspondence with because it minimizes clutter in the drawing.
G.P. (January 2011), Wade commented that he believes he first heard (25) Wulfsberg, G. Inorganic Chemistry; University Science Books:
the comment from Rundle at a lecture in England during the period Sausalito, CA, 2000.
1954−1959. (26) (a) Parkin, G. Metal−metal bonding in bridging hydride and
(12) Another problem with the term “electron deficient” is that it is alkyl Compounds. Struct. Bonding 2010, 136, 113−145. (b) Baik, M.-
also used to refer to molecules in which one of the atoms is H.; Friesner, R. A.; Parkin, G. Theoretical investigation of the metal-
hypovalent (i.e., it has less than an octet configuration for a main metal interaction in dimolybdenum complexes with bridging hydride
group element or less than an 18-electron configuration for a and methyl ligands. Polyhedron 2004, 23, 2879−2900.
transition metal). For example, monomeric BH3 is considered to be (27) Sherwood, D. E.; Hall, M. B. Electronic structure of metal
electron deficient (hypovalent) because the boron has only a sextet clusters. 2. Photoelectron spectrum and molecular orbital calculations
configuration. of decacarbonyldihydridotriosmium. Inorg. Chem. 1982, 21, 3458−
(13) (a) Miessler, G. L.; Fischer, P. J.; Tarr, D. A. Inorganic 3464.
Chemistry, 5th ed.; Pearson: Boston, MA, 2014. (b) Housecroft, C. E.; (28) (a) Collman, J. P.; Hegedus, L. S.; Norton, J. R.; Finke, R. G.
Sharpe, A. G. Inorganic Chemistry, 5th ed.; Pearson: Boston, MA, Principles and Applications of Organotransition Metal Chemistry;
2018. University Science Books: Mill Valley, CA, 1987. (b) Owen, S. M.
(14) This statement assumes that the molecule is neutral. Obviously, Electron counting in clusters: A view of the concepts. Polyhedron
if the molecule were to be charged, then one of the partners that 1988, 7, 253−283. (c) Housecroft, C. E.; Sharpe, A. G. Inorganic
provides one of the electrons must necessarily also be charged. Chemistry; Prentice Hall: New York, NY, 2001. (d) Bennett, M. J.;
(15) Haaland, A. Covalent versus dative bonds to main group Graham, W. A. G.; Hoyano, J. K.; Hutcheon, W. L. Synthesis and X-
metals, a useful distinction. Angew. Chem., Int. Ed. Engl. 1989, 28, ray structure of di-μ-hydrido-octacarbonyldirhenium. J. Am. Chem.
992−1007. Soc. 1972, 94, 6232−6233.
(16) Parkin, G. Valence, oxidation number and formal charge: Three (29) Besora, M.; Lledós, A. Coordination modes and hydride
related but fundamentally different concepts. J. Chem. Educ. 2006, 83, exchange dynamics in transition metal tetrahydroborate complexes.
791−799. Struct. Bonding 2008, 130, 149−202.
(17) It should be noted that dative bonds are often drawn without (30) (a) Brookhart, M.; Green, M. L. H.; Parkin, G. Agostic
using an arrow or including the formal charges. Although such interactions in transition metal compounds. Proc. Natl. Acad. Sci. U. S.
representations indicate connectivity and minimize clutter in complex A. 2007, 104, 6908−6914. (b) Brookhart, M.; Green, M. L. H.; Wong,
structures, they do not provide a good indication of the electronic L. L. Carbon hydrogen transition-metal bonds. Prog. Inorg. Chem.
nature of the molecule. 2007, 36, 1−124. (c) Brookhart, M.; Green, M. L. H. Carbon-
(18) (a) Green, M. L. H.; Parkin, G. Application of the Covalent hydrogen-transition metal bonds. J. Organomet. Chem. 1983, 250,
Bond Classification method for the teaching of inorganic chemistry. J. 395−408.
Chem. Educ. 2014, 91, 807−816. (b) Parkin, G. Classification of (31) (a) Kubas, G. J. Metal Dihydrogen and σ-Bond Complexes:
organotransition metal compounds. In Comprehensive Organometallic Structure, Theory, and Reactivity; Kluwer Academic/Plenum Publish-
Chemistry III; Crabtree, R. H., Mingos, D. M. P., Eds.; Elsevier: ers: New York, NY, 2001. (b) Kubas, G. Chem. Rev. 2007, 107, 4152−
Oxford, 2007; Vol. 1, pp 1−57. (c) Green, M. L. H. A new approach 4205.
to the formal classification of covalent compounds. J. Organomet. (32) Upmacis, R. K.; Gadd, G. E.; Poliakoff, M.; Simpson, M. B.;
Chem. 1995, 500, 127−148. (d) Covalent Bond Classification. http:// Turner, J. J.; Whyman, R.; Simpson, A. F. The photochemical
www.covalentbondclass.org/ (accessed Sept 19, 2019). synthesis of [Cr(CO)5(H2)] in solution: I.r. evidence for coordinated
(19) (a) Jensen, W. B. The place of zinc, cadmium, and mercury in molecular dihydrogen. J. Chem. Soc., Chem. Commun. 1985, 27−30.
the periodic table. J. Chem. Educ. 2003, 80, 952−961. (b) Constable, (33) (a) Weller, A. S.; Chadwick, F. M.; McKay, A. I. Transition
E. C. Evolution and understanding of the d-block elements in the metal alkane-sigma complexes: Synthesis, characterization, and
periodic table. Dalton Trans. 2019, 48, 9408−9421. reactivity. Adv. Organomet. Chem. 2016, 66, 223−276. (b) Hall, C.;

2474 DOI: 10.1021/acs.jchemed.9b00750


J. Chem. Educ. 2019, 96, 2467−2475
Journal of Chemical Education Article

Perutz, R. N. Transition metal alkane complexes. Chem. Rev. 1996, 96,


3125−3146. (c) Crabtree, R. H. Aspects of methane chemistry. Chem.
Rev. 1995, 95, 987−1007.
(34) Corey, J. Y. Reactions of hydrosilanes with transition metal
complexes and characterization of the products. Chem. Rev. 2011, 111,
863−1071.

2475 DOI: 10.1021/acs.jchemed.9b00750


J. Chem. Educ. 2019, 96, 2467−2475

You might also like