Vous êtes sur la page 1sur 7

I

Hard and Soft Acids and Bases,


Ralph G. Pearson
Northwestern University
Evanston, Illinois 60201
I HSAB, Part I
I Fundamental principles
According to G. N. Lewis a base is an
atom, molecule, or ion which has at least one pair of
valence electrons which are not already being shared in a
covalent bond. An acid is similarly a unit in which at
least one atom has a vacant orbital in which a pair of
electrons can be accommodated. The typical acid-
base reaction is
The species A:B may be called a coordination com-
pound, an adduct, or an acid-base complex. I n fact,
a wide variety of acid-base complexes exist under dif-
ferent names. The species A is usually called a Lewis
acid, or a generalized acid, to avoid confusion with
Brfinsted acids. A Lewis base, B, is identical with a
Brfinsted base.
Sidgwick suggested the terms electron acceptor, in
place of Lewis acid, and electron donor, in place of
base. These terms are widely used, particularly in
Europe. Also certain types of weak generalized acid-
base interactions are almost always discussed under the
heading of donor-acceptor complexes. The disad-
vantage is that a different term, electron donor, is
used to describe substances which are generally and
conveniently called bases. It is also true that special
names are sometimes used for special categories. The
use of the term ligand in place of base when A in eqn.
(1) is a metal ion is firmly established. Also in speak-
ing of rates of reaction it is usual to call A an electro-
phile and to call B a nucleophile.'
Probably the most important class of chemical reac-
tion is the generalized acid-base reaction of eqn. (1).
The easiest way to appreciate this is to consider the
different kinds of acid-base complexes, A:R, that may
be formed. For example, all metal atoms or ions are
Lewis acids. They are usually found coordinated to
several bases or ligands simultaneously since they are
polyvalent. The combination m+y be electrically
charged, in which case we have a complex ion formed.
Also the combination may be electrically neutral in
which case a normal inorganic molecule such as SnCli is
formed.
Most cations are T,ewis acids and most anions are
bases. Hence salts are automatically acid-base com-
plexes. MgClz in the solid state consists of the acid
Mg2+ coordinated to six neighboring C1- ligands. In di-
lute solution the magnesium ion is coordinated to the
EDITOR'S NOTE: This is a two part article. The second
part, which discusses the theorys underlying the HSAB principle,
will appear in the October issue.
HAYEK, E., Osters. Chem-Zenl., 63, 109 (1962), has a general
discussion of the problem of acid-base terminology.
base water forming t,he solvated ion. The chloride ion
also forms an acid-base complex, via hydrogen bonding
in which water molecules are 1,ewis acids.
Inorganic compounds, as solids, liquids, gases, or in
solution, are examples of acid-base complexes. The
same t.hing can be said for organic compounds. The
met,hod here is to ment,ally dissect, t,he organic molecule
into two fragments, one of which is a 1,ewis acid and
the other a base. For example, ethyl alcohol can be
thought of as composed of t,he ethyl carbonium ion,
C2Hs+, and the hydroxide ion, OH-. All carbonium
ions are Lewis acids and the hydroxide ion is the base.
Similarly, ethyl acetate, can be thought of as a com-
plex of the acylium ion, CH3CO+, which is an acid,
and t.he base ethoxide ion, CZH50-. Even a hydrocar-
bon can be broken down (conceptually) into an acid
such as H+ and a carbanion, which is a base. Thus
methane can be considered to be H+ and CHI- com-
bined. It is equally t,rue that CHI can be viewed as a
combination of CH3+ and H-. Such ambiguity is, in
fact, universal among both organic and inorganic
molecules. While at first confusing, it turns out to be
absolutely necessary to explain the variety of reactions
undergone by these molecules. That is, methane some-
times reacts as if it were splitting into CH8+ and H-,
and sometimes as if it were splitting into CHa- and Hf .
In addition, there are reactions in which it behaves as
CH3 ' and H ' . These are redox, or free radical reac-
tions, however, and do not concern us here.
I n the case of ethyl acetate, the molecule can be
viewed as an acid-base complex as explained above.
It is also true that it is a Lewis acid and a Lewis base.
Ethyl acetate acts as a Lewis base when it forms com-
plexes through one of its oxygen atoms to the proton,
or other Lewis acids. I t acts as a Lewis acid when it
adds bases, such as the hydroxide ion, to its carbonyl
group. Such acid-base processes are important in the
acid and base catalyzed hydrolyses of esters.
I t should be noted that a certain group of atoms is
often designated as an acid or base even if it has no
stable existence. A carbonium ion such as CH3+ is
considered to be a Lewis acid because its structure
shows that it can accept a pair of electrons from a base.
The breaking down of a molecule, such as CH3C1, into
a methyl carbonium ion and a chloride ion is a purely
conceptual process and has nothing to do with the stabil-
ity of CH3f. The point is that most reactions of
methyl chloride can be classified as being exchanges of
the chloride ion by other bases, or of the methyl cat-
ion by other Lewis acids. Just as the proton does not
exist free under ordinary circumstances, so it is likely
that CH3+ does not ordinarily exist as a free species.
Volume 45, ~ umb; r 9, September 1968 / 581
This brings up the point that equ. (1) is oversimpli-
fied.
What actually occurs in most cases is the ex-
change reaction
A: B' + A' : BS A: B+ A' : B' (2)
In solution A' and B' are often solvent molecules.
Thus, as already mentioned in the case of ions dis-
solved in water, most solute-solvent interactions can
be classified as generalized acid-base reactions. A polar
molecule will always have an electron rich, or basic
site, and an electron poor, or acid site.
Other kinds of acid-base complexes are the so-called
charge transfer complexes which are responsible for the
colors produced when many substances are mixed.
An example is iodine and the intense brown color it
gives in solvents which are bases, such as water, alco-
hols, or ethers. Many charge transfer complexes are
formed between Lewis acids which are unsaturated
molecules with electron-withdrawing substituents, such
as tetracyanoethylene, and unsaturated molecules with
electron donating substituents, snch as toluene. These
systems are called r-acids and r-bases, respectively.
Finally, free atoms and radicals containing electro-
negative elements act as Lewis acids and form com-
plexes with a variety of bases. These complexes of
free radicals cannot be isolated but they have a very
great effect on the reactivity of the radicals.
It is apparent that it is possible, in principle at least,
to view the greater part of chemistry as examples of
interaction of generalized acids and bases. This in
turn means that any rules that can be developed con-
cerning the stability of complexes A. B in reaction (1)
will have very wide application and can he useful in
many areas. Recently a rule has heen suggested,
"The Principle of Hard and Soft Acids and Bases," or
HSAB principle, which does seem to have value in un-
derstanding a wide variety of chemical phenomena (1).
Acid and Base Strength
The concept of acid or hase strength comes in at this
point. In a qualitative way, what is meant by general-
ized acid or hase strength is simply that a strong acid,
A, and a strong base, B, will form a stable complex,
A:B. A weaker acid and hase will form a less stable
complex. Operationally we may define generalized
acid and base strength by competition experiments.
Consider the acid-base substitution reactions
A1 + A: B - Af : B + A (3)
B ' + A: B - A: B ' + B ( 4)
If the reactions go as indicated, it means that A' is a
stronger Lewis acid than A, and that B' is a stronger
base than B. If it were possible to put all Lewis acids
into an order of decreasing strength, and the same for
all bases, then i t would he possible to predict the stabil-
ities of all possible acid-base complexes. That is, we
could predict what chemical reactions would occur nn-
der various conditions, what compounds would be
stable, etc.
We might then expect the equilibrium constant of
eqn. (1) to be given by an eqnation such as
log K = SASB ( 5)
where SA is a strength factor for the acid and SB is a
strength factor for the base. Equation (5) is not the
only equation that might result, but it wonld be the
simplest one that would correctly predict the direction
of displacement reactions such as (3) and (4). Of
course, SA and Se would he functions of the environ-
ment and the temperature. The "intrinsic" strengths
wonld presumably refer to gas phase reactions. There
would then be strong solvent corrections, but even so
an eqnation such as eqn. (5) would be most useful
since a series of SA and SB values would need to be de-
termined once and for all in water at 25"C, and the
stabilities of possible complexes would be known in
that medium. SA and SB values for 100 acids and 100
bases wonld predict the stabilities of 10,000 complexes,
for e~ampl e. ~
Unfortunately it is not possible to write down any
universal order of acid or base strength (B). If a series
- ~.
of different bases, B', are used in testing against a fixed
reference A:B, then the order of base strength that one
gets is very much a function of the nature of the refer-
ence acid, A. When A is the chromic ion, one gets a
different order from the case when A is the cuprous
ion. In fact, the order is different for cupric ion and
for cuprons ion so that even a change in oxidation state
of the reference acid can have an effect on relative base
strengths.
We are forced to the conclusion that there is no
straight-forward way of evaluating base strengths or
acid strengths even if we were to agree that only gas
phase data for reaction (1) be used. Of course, a useful
scale of hase strengths towards the special Lewis acid,
the proton, in aqueous solution does exist. It must
always be remembered that this scale is not valid for all
Lewis acids. It can be useful in a general sense, how-
ever, in telling us that some bases snch as Clop-, are
very weak, and other bases, such as H-, are very strong.
Such a rough ordering will usually be valid if two hases
of widely different strengths are being compared.
A comparable scale to order Lewis acids in aqueous
solution has not been devised, partly because many
Lewis acids are decomposed by water. For many
others, a scale in which the hydroxide ion is a reference
base could be used. In aqueous solution we could write
both the reactions
H+(aq) + B( a d f BHt(aq) (6)
OH-(aq) + A(aq) + AOH-(aq) (7)
Any arbitrary value of SA could be assigned to the pro-
ton. Let us pick 9.0, for example. This would lead
t o values of Ss ranging from about 5 for a strong base
snch as CHa- to -1 for a weak base such as I-. The
value of SB for the hydroxide ion wonld then be (15.74/
9) = 1.75. The number 15.74 is the log10 of the equi-
librium constant for eqn. (6) when B is the hydroxide ion
(55.5/1.0 X 10-14).
This value for SB of OH- then leads to values of SA
of 5.9 and 8.6 for the aqueous Hg2+ and CH3+, respec-
tively. These figures come simply from the acid
ionization constants of Hg(H20)22+ and CH30H2+ ap-
plied to reaction (7). Again, such numbers are useful
'One might ask at this point why standard free energies of
formation are not used as the fundamental properties. The
answer is that these involve prior knowledge of the stability of
A:B. We are seeking a method that predicts the properties of
A: B from a knowledge of A and B only, in a given medium.
582 / Journal of Chemicol Education
in telling us that H+ and CH3+ are probably stronger
acids than Hg2+. However, they cannot be combined
with the values of Ss previously obtained to accurately
predict the stabilities of CH3HgH20+, HgI(H20) +,
CHJ, or C2HB in aqueous solution. They may how-
ever, give some rough idea of stabilities. Further-
more, there will be some Lewis acids sufficiently like
the proton so that a Br@nsted relationship exists be-
tween the equilibrium constants, KA, for the reaction
A M + B(aq) = A:B(aq) (8)
and the equilibrium constant for eqn. (6), K.,
KA = GKP (9)
I n such cases eqn. (5) will be valid since eqn. (9) is
si m~l v another form of the eauation urith a eaual to
s*js;+.
Usually the simple eqn. (5) will not be adequate, and
it would be logical to replace it with a more complex
equation involving more parameters. That is, in-
stead of having only one parameter, S, characteristic of
each acid and each base, it will be necessary to have at
least two. Such an equation (not the only one possible)
would be
log K = S& + o*on
(10)
Now a* and us are parameters for each acid and each
base which measure some different characteristic from
that of strength. We will call them "softness" pa-
rameters for reasons that will become clearer later on.
Equations of the form of eqn. (10) have often been
used to represent rate or equilibrium data. Such four-
parameter equations are t,he next necessary stage after
it is found that two-parameter equations (linear free
energy relationships) fail. For example, Drago and
Wayland (5) have suggested an empirical equation
-AH = EAEB + C*CB
(11)
which accurately reproduces heats of reaction for cer-
tain complex forming reactions in nonpolar solvents.
These are reactions between neutral bases and weak
Lewis acids such as Ip or phenol, which forms hydro-
gen bonds.
The E parameters in eqn. ( l l ) , vary with the polarity
of the acids and bases in such a way as to suggest that
their product represents electrostatic bonding. Simi-
larly, the C parameters vary in such a way as to sug-
gest that their product represents covalent bonding be-
tween A and B. The heat of reaction, AH, is closely
related to the free energy changes AGO, and hence to
the log of the equilibrium constant.
Another four-parameter equation is the celebrated
Edwards equation (4)
logdK/Ko) = nE. + BH (12)
E, is a redox factor defined by E, = Eo + 2.60, where
E" is the standard oxidation potential for the process.
2 B - = B , + 2e- (13)
H is a proton basicity factor defined by H = 1.74 +
pK.. Both definitions are arranged so that H = 0
and En = 0 for water at 25'C. KO is therefore the
constant when the base, B, is water. Equation (14)
may conveniently be called the oxibase scale (5). It
quite successfully correlates a large amount of rate and
equilibrium data.
Table 1. A Comparison of cr and 6 Values for the Edwards
Equation (4)
Lewis Acid a B
Data frornyingst and McDaniel(6) except for I*, Edwards (4).
The way in which a and B in eqn. (12) depend upon
the Lewis acid substrate is quite revealing. Table 1
shows a compilation of values of a and (3for the forma-
tion of a number of metal complexes with various bases
(6). It can be seen that p is large for Lewis acids with
a high positive charge and small size, and small for
Lewis acids of low charge and large size. I n other
words p varies just exactly as SA is expected to vary in
eqn. (5). The term that it is multiplied by, H, is
simply another way of expressing Sn. We can ac-
cordingly identify t,he product BH with SJ, . Then
aE, must be identified with oAon which means that on
is large for bases t.hat are easily oxidized, such as I-,
and small, or negative, for bases that are hard to
oxidize, such as F-. We can also see t,hat a, or a*, is
large for Lewis acids of large size, low positive charge,
and containing unshared electrons in p or d orbitals in
the valence shell, such as Ag+. Also oA is small for
Lewis acids of the opposite characteristics, such as
R4g2+.
While the Edwards equation is of the form of eqn.
(5), it is not the only equation that might be used,
even in aqueous solution. For example, a11 entirely
empirical equation, such as eqn. ( l l ) , might be de-
veloped. In fact, one can see that if eqn. (5), or eqn.
(lo), or eqn. (11) were valid, that equations of the form
log(K/Ko) = y log(K1/Ko') + 6 log(K"/KoX) (14)
should generally exist, where K, K', and K" are any
series of related equilibrium constants. The constants
K' might be taken as the H values of the Edwards
equation, and K" might be values for a typical Lewis
acid with the opposite properties to the proton in Table
1, such as Hg2+.
An even better standard is the methylmercury(1)
cation, CHaHg+, for which a large amount of equilib-
rium data in aqueous solution has been accumulated
by Schwarzenbach.
CH,Hgt (aq) + B(aq) e CHJkBC(aq) (15)
Like the proton, CHaHg+ has the advantage of having
a coordination number of one, which simplifies the
equilibria involved.
Volume 45, Number 9, September 1968 / 583
Table 2.
Equilibrium Constant i n H z 0 at 25'C for the
Reaction CHaHg(H@)+ + BH+ CHaHgB+ + HaO+
B log K , B log K,,
F-
CI -
Br-
I-
OH-
HP04*-
HP08=-
sz-
HOCHGHIS
SCN-
See MUS~RAVE and KELLER, Inorganic Chem., 4,1793 (1965).
Estimated by assuming th+t K., for CH8HgCl + C H S
(CHa)~Hg + HCI is the same in the gas hese and 111 water.
Remaining data from G. Schwarsenbsch d'~. Schellenberg (7).
With two reference acids of different properties we
can test various bases to see if they prefer to bind to H+
or CH2Hg+. Table 2 shows some data (7) for the
equilibrium constant for the exchange reaction
BH+ + CH,Hg(HnO)+ CH8HgB+ + HsOt (16)
The important feature which we note is that bases in
which the donor atom is N, 0, or F prefer to coordinate
to the proton. Bases in which the donor atom is P, S,
I, Br, C1, or C prefer to coordinate to mercury.
The donor atoms in the first group are those which
are of high electronegativity, of low polarizability, and
hard to oxidize. Let us call bases containing these
donors "hard" bases, to emphasize the fact that they
hold on to their electrons tightly. The donor atoms of
the other bases are of low electronegativity, of high
polarizahility, and easy to oxidize. Let us call bases
containing these donor atoms "soft" bases, a term
which graphically describes the looseness with which
they hold their valence electrons.
The Principle of Hard and Soft Acids and Bases
We can now classify every conceivable base into one
or the other of three categories, hard, soft, or borderline.
Table 3 shows such a classification. The borderline
category takes into account such factors as the presence
of unsaturation in some nitrogen donors, which should
loosen up the valence electrons. I t also recognizes
that chloride ion is less soft than bromide ion, which in
turn is less soft than iodide ion. An important featufe
Table 3. Clossification of Bases
Hard Softs
H1O, OH-, F- RS, RSH, RS-
CH1CO2; PO&'-, S O P I; SCN-, S10r2-
CI-, Coal-, Clod-, NO,- RsP, &As, (ROhP
ROH, RO; R,O CN-, HNC, CO
NH,, RNHn, N2Hd CnHa, C&
T I - . R-
Borderline
CsHsNH2, CrHrN, N1-, Br; Nos-, SO?, Nx
-
6 The symbol It stands for an alkyl or group.
is that Table 3 could be constructed in two independent
ways: by considering the properties of the donor atom
(easily oxidized, polarizable; etc.), or by estimating
the equilibrium constant of reaction (16). The re-
sults would be the same, with a few exceptions. For
example, C1- has an unusually high affinity for mer-
cury ion, as do Br- and I-, and the values of K,, in
Table 2 are somewhat anomalous for this reason. Per-
haps C1- should be considered a borderline base.
We can now proceed to an equivalent classification
of Lewis acids into three categories, including border-
line cases. We simply ask the question whether the
acid is like the proton in preferring the hard bases of
Table 3, or like CH,Hg+ in preferring the soft bases.
Operationally this is best done by using the criteria of
Schwarzenbach (8) and of Ahrland, Chatt, and Davies
(9). For complexes with different donor atoms, the
following sequences of stabilities are found
(hard) 0 >> S > Se > Te
F > C l > R v > l
If we consider any Lewis acid, such as Cu+, NO+, or
I%, we simply examine the literature to see if a complex
such as CuI or CuF is more stable, or if Cu(PR&+ is
more &able than Cu(NHJZ+.
The term stable is ambiguous, as ordinarily used, and
a strict definition would refer to the equilibrium con-
stants for reactions in water such as
CuF(aq) + 1-(aq) CuI(aq) + F- (17)
Cu(PRdt(aq) + NHdaq) = Cu(NHd+(aq) + PRdaq) (18)
Oftentimes the data is incomplete and a variety of in-
terpolations must be made to draw a conclusion (1).
Nevertheless, it is usually possible to conclude that a
Lewis acid prefers either the hard bases of Table 3,
or the soft bases.
When this is done for a large number of Lewis acids,
the results are as shown in Table 4. The entries in the
Table 4. Classification of Lewis Acids
Hard
lVl"""
BeMe., Bh , B(OR)r
A1(CH1)l, AICls, AlHa
RPOpt, ItOPO2
RSOst, ROSOg+, SO,
I,+ I$+ Cl'f C"6+
nc'o+, 'co,, kc+
HX (hydrogen bonding
molecules)
Soft
GeCb, Gab, InCL
IM+, RSet, RTe'
L, Brz, ICN, etc.
trinitrabenzene, etc.
chloranil, quinones, etc.
tetraeyanaithylene, etc.
0, CI, Br, I, N, RO' , lt02'
M" (metal atoms)
bulk metals
CHS, carbenes
Borderline
FeP+ Cox+ Nis+ CN~+, %a+, Pba+, Sn2+ SbJ+, Biz+
R$+, ~r:: B( ~ H&, SO*, NO+, ltu2+, b+, RaCt,
CsH5+, Ga&
584 / Journol of Chemical Education
left hand column are class (a) acids, and those on the
right are class (b) acids. However, instead of following
this method of naming, which has historical precedence,
a different system of naming has some advantages.
If we examine the class (a) Lewis acids, we find that
the acceptor atoms are small in size, of high positive
charge, and do not contain unshared pairs of electrons
in their valence shell (not all of these properties need
he possessed by any one acid). Now these are all
properties which lead to high electronegativity and
low polarizability. It seems appropriate to call such
acids "hard." The class (b) Lewis acids, generally
speaking, have acceptor atoms large in size, of low posi-
tive charge, and containing unshared pairs of electrons
(p or d electrons) in their valence shell. These proper-
ties lead to high polarizability and low electronega-
tivity. Again it seems reasonable to call these 1,ewis
acids "soft."
A comparison of Tahles 1 and 4 shows that Lewis
acids with large m and f l values are all in the soft group
and those with small a and large p values are in the
hard group. Thus we can say that soft acids form
stahle complexes with hases that are highly polarizable
and are good reducing agents, and not necessarily good
bases towards the proton. Hard acids, of which the
proton itself is typical, will usually form stable com-
plexes with bases that are good bases towards the pro-
ton. Polarizability, or the reducing properties of the
base, play a minor role.
If we arrange the donor atoms of the most common
bases in an order of increasing electronegativity, we will
have As, P < C, Se, S, I < Br < N, Cl < O < F. Soft
Lewis acids will form more stable complexes with left
hand members of this series, and hard Lewis acids will
form more stable complexes with the right hand mem-
bers of the series. For example, a rare earth ion, La3+,
will form complexes only with N, 0, and F donors, or
with hard bases.
If one accepts the system of naming used in Tables 3
and 4, a very simple rule can now be stated concerning
the stability of acid-base complexes. The rule is that:
Hard acids prefer to bind to hard bases and soft acids pre-
fer to bind to soft bases.
The rule is a concise statement which sums up the
experimental information used to compile Tables 3
and 4. It is a statement of fact and is not to be re-
garded as a theory or as a hypothesis. Such generalized
statements covering many facts were often called laws.
However, modern scientific practice is to reserve the
name "law" only for those generalizations which are
capable of rather precise mathematical formulation.
For this reason it seems best to call the above rule
"The Principle of Hard and Soft Acids and Bases."
Eventually, it may become possible to make qnantita-
tive predictions based on equations similar to eqn.
(10).
Unfortunately eqn. (10) cannot be expected to he
exact, or even nearly exact.
It is too simple to repre-
sent the complexity of changes that occur when elec-
tron-donating groups combine with electron-accepting
groups. It is certainly much better than eqn. (5)
since it has more parameters in it.
It is as good as
eqns. (11) and (12), since it is simply the general pro-
totype of any four parameter equation. Yet eqns. (11)
and (12), while very good over a limited range, cannot
reproduce data over a very wide range of Lewis acids
and bases. Of course, in any case, different values of
all the parameters of these equations would be needed,
if one changed the solvent or the temperature.
I n the case of the HSAB principle we have a simple,
but imprecise, law with a very wide range of appli-
cability. In spite of the lack of precision the rule does
appear to have considerable utility. It can be used in
prediction; ~er haps more important, it can be ex-
tremely helpful in correlating the vast amount of
chemical information which we already have at hand.
Estimations of Strength and of Softness
What has been suggested in the previous section is
that two properties of an acid and a base are needed to
make an estimate of the stability of the complex which
they might form. One property is what we may call
the intrinsic strength (SA or Sa in eqn. (10)); the other
is the hardness or softness (FA or uB in eqn. (10)).
While various arbitrary scales of strength and softness
can be devised, such as pK. for strength and pKoHIHg+
for softness, it seems best to leave them undefined
operationally at present, and to use qualitative defini-
tions based on the properties of the acids and bases.
That is, we need to have methods of estimating the
strength and the softness of an acid or base which
depend on a knowledge of their chemical compositions
and electronic structures.
In fact we have long had such rules for estimating
Lewis acid or base strength. We know that for cations
increased charge and decreased radius make for strong
acids. For anions increased charge and decreased
radius also increase base strength. Thus 0%- is a
stronger base than OH-, and stronger than SeZ-.
The ions A13+, A1C12+, AlCL+, and AICla will have
steadily decreasing intrinsic strengths. Neutral acids
and bases will have strengths proportional to the local
dipoles at the acceptor or donor atom sites. More
remote substituents also have rather predictable ef-
fects on acid or base strength in terms of electron with-
drawing properties that they may have, or the local
dipoles that they create (10).
Of course, in chemistry we rarely are concerned with
the properties of a bare ion, such as A13+. Instead we
deal with such ions in various environments. This
alters the nature and strength of the Lewis acid in-
volved. We are more likely to need to know that the
Lewi sa~i dsAl (H~O)~~+, Al(H~0)&1'+, andA1(HZ0)3C12+
decrease steadily in acid strength. The coordination
number (number of base molecules or ligands attached
to the central acceptor,molecule) is reduced by one in
these examples to show the unit which is the Lewis
acid. I n Table 4 are listed a number of ions. These
are always meant to be the aquated ions less one mole-
cule of water, Ni(HZO)?++, B( HZO) ~+, etc. Such aquo
ions are very much weaker acids than the rather hypo-
thetical bare ions would be.=
' Solvation effects on base strengths are also very important.
For example, CHa- is a much stronger base than OH-in aqueous
solution. In the gas phase, however, the base strengths are
virtually identical, that is, the intrinsic strengths of CH8- and
OH- are the same. Strong solvation of OH- by water accounts
for the difference in solution. JOLLY, W. L., J. CREM. EDUC.,
44, 304 (19671, has 8. discussion of salvation effects.
Volume 45, Number 9, September 1968 / 585
Just as we can make reasonable guesses about in-
trinsic acid and base strength, so we can make estimates
of hardness and softness. This was, in fact, done for
hases in constructing Table 3 by a simple examination
of the nature of the donor atom (electronegative and
nonpolarizable like F, or polarizable and not electro-
negative like I). We can also assign increasing softness
within related series without much ambiguity. Thus
SbR3, AsRs, PR3, and NRs should be of decreasing
softness, as are CH3-, NH2-, OH-, and F-. The ef-
fect of oxidation state on a given donor element is
predictable; sulfur(1V) in SOZ2- should be a harder
base than sulfur(-11) in St-.
It is somewhat suprising that there does not seem to
he much difference in hardness between H20, OH,- and
02-. All three are very hard bases by any criteria
and any difference between them is masked by other
effects. We would have expected the polarizability
(and hence the softness) to increase in the order H20
<OH-<02-. Also it is not easy to decide if there is
any difference in hardness between various oxygen
donors such as CH&OO-, Son2-, POa-, etc.
For the Lewis acids, the important properties that
determine softness are size, charge or oxidation state,
electronic structure, and the other attached groups.
Both metals and nonmetals can be acceptor atoms in
acid-base complexes. For elements of variable valence
there is usually a smooth increase in hardness as the
oxidation state increases. Thus Ni(0) (in Ni(CO)a, for
example) is soft, Ni(I1) is borderline, and Ni(1V) is
hard. The sulfur atom of the sulfenyl group RS+ is a
soft Lewis acid whereas the sulfur atom of the sulfonyl
group RSOz+ is hard. The formal oxidation state
changes from plus two to plus six. Other examples
can be found in Table 4.
Exceptions do occur at the end of the transition
series. It is certain that Tl(II1) is softer than Tl(1)
and it is likely that Hg(I1) is softer than Hg(1). The
evidence is incomplete, but it is possible that Ph(1V) is
softer than Pb(I1). These cases all involve the inert
pair of electrons in the 5s or 6s orbitals. It seems that
the presence of electrons in these particular orbitals
decreases softness by a shielding effect on the outer d
electrons.
The importance of the d electrons for metal ions is
very great. As Ahrland has pointed out (11) no really
good class (b), or soft, acceptor among the metals ex-
sists which does not have at least a half-filled outer d
shell. This accounts for another anomaly. I n going
across a transition series, e.g., from Ca to Zn, the ioniza-
tion potentials of the atoms increase because of in-
creasing nuclear charge. One would intevret this as
meaning that the elements become more electronega-
tive, that is, harder, as one goes across from Ca to Zn
(12). I n fact, chemically the elements become softer.
This is a consequence of the increasing number of d
electrons, a factor which outweighs increasing electro-
negativity.
Fortunately, for the representative elements, and for
the nonmetals in particular, this complication does not
arise and softness seems to be a predictable function of
oxidation state. Of equal importance is the nature of
the other groups attached to the acceptor atom. We
see in Table 4 that BF, is a hard acid, and BH3 is a
soft acid. Experimentally it is found that BF3.0R2
is more stable than BF8.SR2, whereas for BH3 the re-
verse is true. Borine will even form a carbonyl,
BH3C0. Formation of complexes with carbon mon-
oxide or olefins is a good test for soft behavior.
I n both BF3 and BH3 the boron is formally in a plus
three oxidation state, yet quite different behavior is
noted. The presence of hard fluoride ions in BF3 makes
it easy to add other hard bases. The presence of soft
hydride ions in BHJ makes it easy to add other soft
hases. This important effect was particularly com-
mented on by Jgrgensen who coined the name "sym-
biosis" to dcscrihe it (19). Soft bases tend to group
together on a given central atom and hard ligands tend
to group together ("birds of a feather flock together").
The mutual stabilizing effect is called symbio~is. ~
The explanation for symbiosis is rather easily seen.
The hard F- ligands form a complex which is largely
ionic. Hence the boron atom in BF3 is nearly B3+ and
is hard. The soft H- ligands donate negative charge
extensively to the central boron, by covalency or by
simple polarization. As a result the boron atom in
BH3 is almost neutral and naturally becomes soft.
The conclusion that it is the actual charge on the
central atom, rather than the formal charge, which
determines softness seems perfectly logical. While it
complicates the assignment of hardness or softness in
some cases, it helps explain many otherwise puzzling
phenomena (14). For example, the existence of ions
such as Ass4=- and MO(SCN)~- can be rationalized in
spite of the high formal oxidation state of the central
atom. I n the latter case, the thiocyanate ion is be-
lieved to be S-bonded rather than N-bonded.
The mode of bonding of the thiocyanate ion is often
used as a test of hard or soft behavior. The sulfur end
is assumed to be much softer than the nitrogen end,
and hence to prefer soft Lewis acids. An interesting
test of this assumption has been made (15). A study
was made of the complexing of both an alkyl thio-
cyanate, RSCN, and an alkyl isothiocyanate, RNCS,
with the soft Lewis acid, iodine, and the hard Lewis
acid, phenol. The thiocyanate, RSCN, with a free
nitrogen end to act as a donor, formed a more stable
complex with phenol (hydrogen-bonding) than with 1 2
(charge-transfer). With the isothiocyanate, exactly
the opposite results were found. However, the dif-
ferences were very small and the criterion must be used
with caution.
Note that saying that the group MO(SCN)~ is a soft
Lewis acid is not the same as savine that Mo(V) is a -
soft acid. In fact, MOO(NCS)~;- Grns out to be N-
bonded (16). Similarly, Rh(NH&'+ is a borderline
case whereas Rh(SCN)? is a soft acid (17). As
J#rgensen (18) point,s out, it is likely that most elements
in a very high oxidation state could form complexes
containing a maximum number of soft groups, if it
were not for spontaneous oxidation-reduction (MnS,-,
"ymbiosis is counter-halanced by other factors. Otherwise
all eqnilibria of the type
MX, + MY. S 2MX,Y,
would lie t,o the left.. In fact, they can go in either direction.
The other factor is probably that of intrinsic strength which will
favor the mixed species if X and Y differ markedly in base
strength.
586 / Journal of Chemical Education
for instance). However, mixed species such as R'InOz-
Sz- should be unstable in any case because they lack
symbiotic stabilization.
Liferatwe Cited
(1) PEARSON, R. G., J. Am. Chem. Soc., 85, 3533 (1963);
Science, 151, 172 (1066); Chmislrli i n Britain, 3, 103
(1967).
(2) LEWIS, G. N., "Vdenre and the Stnlctnre of At,om md
Molecules," The Chemical Catalog Co., New York,
N. Y., 1923; LEWIS, G. N., J . Franklin Institute, 226,
293 (1938).
(7) S c n w a ~ z ~ ~ n a c ~ , G., AND SHELLENRERG, M., Helv. Chim.
Acta, 48, 28 (196.5).
(8) SCHIYRZENRICH, G., Ezperentia Suppl., 5, 162 (1956).
(9) AHRLAND, S., CHATT, J., AND DAVIES, N. R., Quart. Revs.,
(London1 12.265 (1958I.
(10) BE'LL, R. P., ';The ~ r n & in Chemistry," Cornell Univer-
sity Press, Ithilca, 1959, Chapter 7.
(11) AHELAND, S., Structureand Bonding, 1, 207, (1966).
(12) Williams, R. J . P., and tlale, J. D., Structure and Bonding,
1, 249 (1966).
(13) J&GENSEN, C. K., Inorg. Chem., 3,1201 (1964).
(14) BASOLO, F., AND BURMRISTER, J., Ino~g. Chem., 3, 1587
11964).
, ~,
(3) DRAGO; R. s., .AND W~ Y~ AND, B. R., J . Am. Chem. Soe., (1.5) WAYLAND, B. B., AND (GOLD, R. H., Inorg. Chem., 5, 154
87, 3571 (1965). (1966).
(4) EDWARDS, J . O., J. Am. Chem. Soc., 76, 1540 (1954). (16) MITCHELL, P. C. H., Quwt. Rev. (London), 20, 103 (1966).
(5) Dnvrs, R. E., J. Am. Chem. Soc, 87, 3010 (1965). (17) SCHMIDTKE, H. H., J. Am. Chem. Soc., 87, 2522 (1965).
(6) YINGST, A,, AND MCDANIEL, 11. H., Ino~ganic Chm. , 6,1067 (18) JORGENSEN, C. K., St ruct u~e~ and Bonding, 1, 234 (1966).
(1967).
Vol ume 45, Number 9, September 1968 / 587

Vous aimerez peut-être aussi