Anda di halaman 1dari 3

10292 Chem. Commun.

, 2012, 48, 1029210294 This journal is c The Royal Society of Chemistry 2012
Cite this: Chem. Commun., 2012, 48, 1029210294
A gas phase perspective on the Lewis acidity of metal ions in aqueous solutionw
Xiaojing Chenz and Anthony J. Stace*
Received 13th August 2012, Accepted 5th September 2012
DOI: 10.1039/c2cc35859j
A pattern of instability exhibited by solvated metal dications in
the gas phase shows a remarkable correlation (covering 10 orders
of magnitude) with the ability of the same metal ions to promote
the release of protons in aqueous solution.
When in aqueous solution many metal ions promote acidity via the
hydrolysis reaction:
1
M
n+
+ H
2
O 3 [M(H
2
O)]
n+
3
M(OH)
(n1)+
+ H
+
. Thus, for doubly charged metal cations it
is possible to dene a hydrolysis constant, K
h
, as: [M
+
OH][H
+
]/
[M
2+
], which provides a measure of the acidic strength of water
when M
2+
cations are present in aqueous solution.
1
For a number
of doubly charged metal ions there exists a reasonably good
correlation between their pK
h
(log K
h
) value and their charge to
size ratio (q
2
/r
ion
).
2
Fig. S1w shows an example of this type of plot
for the nine metal ions that are the subject of this work, and it can
be seen that for several of them the correlation is reasonably good,
but that there are at least two very obvious exceptions.
Alternative combinations of the variables q and r have been
used to present the pK
h
data,
36
for example pK
h
vs. q
2
/r
ion
2
and
pK
h
vs. q
2
/[r(MOH
2
)n] have been used, where r(MOH
2
) is the
MOH
2
bond length and n is an assumed primary shell solvation
number. Irrespective of how the data are presented, there are some
very notable exceptions to any of the correlations, with the most
obvious being Sn
2+
and Pb
2+
. The implication from Fig. S1 is
that both ions appear to be more acidic than would be predicted
purely from their size, and therefore additional factors must
inuence the magnitude of pK
h
. For example, Ca
2+
and Sr
2+
have comparable ionic radii to Pb
2+
and yet in aqueous solution
the latter yields a proton concentration that is two orders of
magnitude larger than either of the other two ions. In addition to
the problem highlighted for Sn
2+
and Pb
2+
, Fig. S1w would also
suggest that Mg
2+
is less acidic than it should be for its size.
Although there are no clear explanations as to why Sn
2+
and
Pb
2+
behave dierently, suggestions have included: (i) their soft-
ness;
4
(ii) changes in coordination number, particularly in the case
of Sn
2+
;
5,6
and (iii) covalent contributions to the MOH
2
bond.
5
Many previous attempts to understand and/or quantify hydrolysis
have actually omitted Sn
2+
and Pb
2+
from the discussion. How-
ever, the variant pK
h
vs. q
2
/[r(MOH
2
)n], does give a linear t that
(almost) accommodates both Sn
2+
and Pb
2+
provided n is given a
small enough value (see below).
5
It is acknowledged that, unlike the
other metal ions shown, both Pb
2+
and Sn
2+
have the ability to
formunits in aqueous solution that are more complex than a simple
hydroxide.
5
At a molecular (gas phase) level, the ability of alkaline earth
metal dications to promote proton release fromwater molecules has
been discussed in terms of salt bridge structures that lower barriers
to proton transfer within chains of water molecules.
7
Calculations
of a similar nature, but specic to Sn
2+
and Pb
2+
,
8
show these ions
to have lower energy barriers than is seen in Ca
2+
for the
displacement of coordinated water molecules into geometries that
facilitate the loss of H
3
O
+
(H
2
O)
m
through a salt bridge structure.
8,9
The purpose here is to build on this gas phase view of metal ion
acidity
9
by presenting new experimental data on what could be
considered as the forward step of the hydrolysis reaction, namely:
[M(H
2
O)
n
]
2+
-[M
+
OH](H
2
O)
nm
+ H
+
(H
2
O)
m1
. In the gas
phase this process is frequently referred to as Coulomb ssion and
denotes charge separation in the form of proton loss by an isolated
doubly charged cation in the gas phase. The experimental results
show that the behaviour exhibited by metal ions under these
circumstances provides a newand more fundamental understanding
of factors that inuence the (lack of) pattern seen in Fig. S1w.
Measurements of Coulomb ssion within a series of
[M(H
2
O)
n
]
2+
complexes, have been preformed for the purposes
of identify the degree to which M
2+
ions are unstable when
complexed with water. New results are presented here for the
metal dications Sr
2+
, Ca
2+
, Mg
2+
, Mn
2+
, Cr
2+
and Sn
2+
. These
measurements have been combined with those reported previously
on Cu
2+
, Zn
2+
and Pb
2+
,
1012
to give a data set that spans
10 orders of magnitude in K
h
(Table 1). Details of how the
experiments were performed are contained in the ESIw, and an
example of the very characteristic signature of Coulomb ssion is
given in Fig. S1w for the dication [Mn(H
2
O)
4
]
2+
. Similar observa-
tions were made on each of the metals given above to provide an
unambiguous catalogue of all of the proton transfer pathways they
exhibit. The approach adopted here, where the unimolecular decay
of each ion is monitored, contrasts with some of the previous
experiments where collisional activation has been used to promote
charge transfer.
13,14
For this reason, there are small, but signicant
dierences between the critical sizes for instability presented here
and those quoted elsewhere.
13,14
A complete list of previous critical
size data is presented in Table S1w.
Since no behaviour similar to that given in Fig. S2w is
observed for n 4 4, it is concluded that Mn
2+
requires ve
waters to stabilise the gas phase complex against spontaneous
Department of Physical Chemistry, School of Chemistry, The
University of Nottingham, University Park, Nottingham NG7 2RD,
UK. E-mail: anthony.stace@nottingham.ac.uk
w Electronic supplementary information (ESI) available. See DOI:
10.1039/c2cc35859j
z Present address: Department of Chemistry, University of Basel,
Klingelbergstrasse 80, CH-4056, Basel, Switzerland.
ChemComm
Dynamic Article Links
www.rsc.org/chemcomm COMMUNICATION
P
u
b
l
i
s
h
e
d

o
n

0
6

S
e
p
t
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T
Y

O
F

B
R
I
G
H
T
O
N

o
n

2
7
/
1
0
/
2
0
1
3

1
3
:
5
8
:
2
2
.

View Article Online / Journal Homepage / Table of Contents for this issue
This journal is c The Royal Society of Chemistry 2012 Chem. Commun., 2012, 48, 1029210294 10293
hydrolysis in the gas phase. Previous examples of charge
separation in metal dication complexes have been labelled by the
quantity n
crit
,
13,14
which is the maximum size of complex for which
Coulomb ssion is observed. We have adopted a slightly dierent
quantity n
s
(n
s
= n
crit
+ 1), which is the minimum number of water
molecules required to stabilise a dication complex against sponta-
neous hydrolysis. Table 1 summarises data that have been recorded
for a sample of nine metal cations that span the complete pK
h
range
seen in Fig. S1w. Although the metal dications exhibit considerable
variation in the values determined for n
s
, what is most surprising are
the results obtained for Sn
2+
and Pb
2+
; the latter needs at least 11
water molecules to stabilise the ion against gas phase hydrolysis and
for Sn
2+
an incredible 26 water molecules are required. Both of
these numbers far exceed the conventional picture of a stable
solvated dication surrounded by just a primary shell of water
molecules,
5,15
and the results for Sn
2+
suggest that secondary and
possibly even tertiary shells of water are implicated in the ion
solvation/stabilisation process. An example of Coulomb ssion
recorded for [Sn(H
2
O)
17
]
2+
is shown in Fig. S3w.
Table S2w gives a detailed breakdown of the decay pathways
identied for each of the [M(H
2
O)
n
]
2+
complexes for which
Coulomb ssion has been observed. In each case, it is the metal-
containing fragment ion that has been detected; the signal from the
counter ions is always much weaker because these lighter mass ions
experience a greater degree of instrumental discrimination. What
Table S2w shows is that rstly, fragmentation can be very asym-
metric with the metal-containing fragment carrying away a major
fraction of the water molecules and secondly, the size of the non-
metallic fragment appears to be more specic to the value for n than
to the nature of the metal ion. Thus, n = 34 ions lose H
3
O
+
, n =
56 lose H
3
O
+
(H
2
O) etc., leading eventually to the larger Sn
2+
complexes which predominantly converge to give H
3
O
+
(H
2
O)
6
as a
fragment. The only exception to the above trend is the loss of
H
3
O
+
H
2
O by [Pb(H
2
O)
4
]
2+
, which nicely matches an earlier
calculation of the lowest energy reaction pathway for this ion.
8
The occurrence of H
3
O
+
and H
3
O
+
(H
2
O)
3
as fragments may be
driven by the particular stabilities of these ions. What is possibly
signicant for the larger non-metallic fragments is the fact that
equilibrium measurements on protonated water clusters show that
the cumulative binding energy of approximately 6 molecule is
sucient to match the hydration enthalpy of a proton in bulk
water,
17
which may explain the high degree of asymmetry seen in
the fragmentation patterns of the larger [Sn(H
2
O)
n
]
2+
complexes.
The data in Table 1 suggest a link between pK
h
and the critical size
at which hydrated gas phase dications undergo spontaneous charge
separation. This view is supported by Fig. 1, which shows a plot of
pK
h
against n
s
from where it can be seen that results recorded for
Sn
2+
and Pb
2+
and indeed those for almost all of the ions con-
sidered, exhibit a correlation that is approximately linear. However,
it is not obvious why the data should exhibit a linear dependence on
n
s
and as an alternative, Fig. 1 also shows a t to the equation:
pK
h
= 14 0.7n
s
+ 0.01n
s
2
, which again contains many of the
data points. Within the literature there appear to be uncertainties in
the exact values assigned to pK
h
for some of the metal dications,
and that is reected in the error bars assigned to each value. Sn
2+
in
particular has values given for pK
h
that fall in the range 24.
18,19
Although the complexes [Ca(H
2
O)
2
]
2+
and [Sr(H
2
O)
2
]
2+
can be
identied in our experiments, neither shows any evidence of
Coulomb ssion; therefore, n
s
should be assigned a value of zero.
However, it is quite possible that this absence is due to ion geometry
as the rst step in proton transfer is the promotion of a water
molecule fromthe primary solvation shell to a hydrogen bonded site
on an adjacent water molecule.
7,8
Since closed shell [M(H
2
O)
2
]
2+
complexes, suchas those of Ca
2+
andSr
2+
, will most probably have
the water molecules sitedas far apart as is possible, inthe absence of a
perturbation, the displacement of one molecule inthe directionof the
other is unlikely to happen. In addition, calculations by Beyer et al.
have shown that the barriers to such a promotion in [M(H
2
O)
2
]
2+
complexes are 78 and 58 kJ mol
1
for Ca
2+
and Sr
2+
, respectively.
7
However, both Spears and Fehsenfeld
20
and Feil et al.
21
report the
appearance of CaOH
+
following the collision of either Ca
2+
or
Ca
2+
H
2
O with a water molecule. Likewise, Carl et al.
22
have
provided evidence for the formation of SrOH
+
following the
collisional activation of both [Sr(H
2
O)
2
]
2+
and [Sr(H
2
O)
3
]
2+
with
xenon. Experiments by Carl et al.
23
on the collisional activation of
large Ca
2+
water complexes showed noevidence of protontransfer.
In view of the range of values for n
s
arising from the dierent
experiments on Sr
2+
and Ca
2+
(see also Table S1w), error bars have
been assigned to this number for both ions; however, whichever
values are adopted for Sr
2+
and Ca
2+
there is not a signicant shift
inthe overall patternof Fig. 1. All other values assignedton
s
are as a
consequence of unimolecular (metastable) decay, which due to the
eect of a competitive shift can be shown to follow the reaction
pathway with the lowest critical energy.
24,25
As presented in Fig. 1, the results show evidence of a remarkably
good relationship between n
s
and pK
h
; one which spans 10 orders of
Table 1 Summary of the relevant data used to discuss the behaviour
of metal dications
Metal pK
h
a
Min. stable size/n
s
Ionic radius
c
/pm
Sr
2+
13.113.3 03 116
Ca
2+
12.612.9 03 100
Mg
2+
11.411.8 4 72
Mn
2+
10.110.7 5 82
Zn
2+
8.79.6 8 74
Cr
2+
10.1
b
7 82
Pb
2+
7.29.4 11 94
Cu
2+
7.38.0 8 62
Sn
2+
1.73.4 26 93
a
Compiled from ref. 1, 3, 4, 6 and 15.
b
Considerable uncertainty; this
is an average of a wide range of values.
c
Taken from ref. 16.
Fig. 1 Plot of the acidity constant, pK
h
for each of the metal dications
shown against the minimum number of water molecules, n
s
, required to
stabilise each [M(H
2
O)
n
]
2+
complex against Coulomb ssion. The solid and
dashed lines are linear and quadratic ts to the data points, respectively.
P
u
b
l
i
s
h
e
d

o
n

0
6

S
e
p
t
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T
Y

O
F

B
R
I
G
H
T
O
N

o
n

2
7
/
1
0
/
2
0
1
3

1
3
:
5
8
:
2
2
.

View Article Online
10294 Chem. Commun., 2012, 48, 1029210294 This journal is c The Royal Society of Chemistry 2012
magnitude in K
h
and can quite clearly accommodate not only the
(apparent) excessive acidities of both Sn
2+
and Pb
2+
, but also the
lack of acidity on the part of Mg
2+
. It is interesting to note that the
lines shown in Fig. 1 as ts to the data points approach pK
h
= 14
for n
s
= 0. The implication being that a metal cation exhibiting the
same acidity as pure water (autoprotolysis constant pK
w
= 14),
would either not require any additional water molecules to stabilise
it against hydrolysis or would not actually hydrolyse. It should be
noted that some of the earlier methods used to interpret pK
h
data
do extrapolate to pK
h
B 14,
3
but make no connection to the
behaviour of pure water. Any dierence between the result for pure
water and aqueous metal ions would suggest that, in part, the water
gains acidic strength from the chemical stability of the M
2+
OH

bond and how the resultant structure is solvated.


26
Since both
systems have the H
3
O
+
unit in common, some contrast must come
fromOH

, which in pure water is only stabilised (solvated) through


hydrogen bonding. The most obvious departure in Fig. 1 is Cu
2+
,
but for this ion there are sound geometric reasons why this might be
the case.
4
Gas phase measurements by Bush et al.
27
on the
behaviour of lanthanide metal trications complexed with water also
showed a correlation similar to that given in Fig. 1, but one that
covered just 1.5 units of pK
h
.
Previous attempts to categorise pK
h
data on metal dications,
have divided the metals into four groups:
1,5
Group A most
resistant to hydrolysis (Mg
2+
, Ca
2+
and Sr
2+
); Group B less
resistant for their size and charge (Mn
2+
, Cu
2+
and Zn
2+
); Group
C low resistance to hydrolysis, soft post-transition metals with
lled d shells (Pb
2+
); and Group D anomalously low resistance
to hydrolysis (Sn
2+
). The data in Fig. 1 are consistent with these
groupings. However, two further connections can be made with
established sequences that have been used with considerable success
to classify trends in the behaviour of metal cations. First, the
overall trend in n
s
matches the hard-soft acidbase (HSAB) assign-
ment of metal ions, i.e. soft Pb
2+
(and possibly Sn
2+
) 4borderline
Cu
2+
and Zn
2+
4hard Mg
2+
, Ca
2+
and Mn
2+
.
2830
Secondly,
there is also a more limited link between n
s
and the IrvingWilliams
series, which refers to the relative stabilities of complexes formed by
metal ions. In this instance, n
s
and the IrvingWilliams series follow
the order: Mg
2+
and Ca
2+
o Mn
2+
o Cu
2+
E Zn
2+
.
30,31
Neither of these connections can be made by linking pK
h
, for
example, to the q
2
/r
ion
data given in Fig. S1w.
From their earlier gas phase measurements on the instability of
hydrated metal dications (Table S1w), Kebarle and co-workers
proposed a correlation with the second ionisation energy of the
metal concerned;
13
a suggestion supported by Shvatsburg and Siu
following an investigation into the mass spectra of metal dication
complexes.
14
Calculations by Beyer et al.
7
showed there to be only
an indirect connection between the occurrence of proton transfer
and ionisation energy, and that a more reliable correlation existed
with the radius of a metal ion.
7
A limited range of values for n
s
, for
example those of the alkaline earth metals, do show a trend with
both ionic radius and ionisation energy, but that does not extend
across the complete data set. Compare, for example, the results for
Mg
2+
, Mn
2+
and Pb
2+
, all of which have second ionisation
energies of B15 eV (Table S1w); and the new data presented here
for Sn
2+
fails completely to match any of these correlations.
Measurements presented here on the instability of isolated
metal dication complexes, [M(H
2
O)
n
]
2+
, exhibit an extraordinarily
good correlation with the ability of the same M
2+
ions to release
protons when in aqueous solution. At one extreme, the results show
that the stabilisation of Sn
2+
requires the presence of sucient
molecules to occupy at least two solvation shells. At the other end
of the spectrum, fewer water molecules than is found in a single
solvation shell will stabilise the alkaline earth metals. These
observations suggest that the Lewis acidity of these metal ions is
determined, in part, by the requirement that ions remain full
solvated; something that is readily achieved for Mg
2+
, but
for Sn
2+
thermal motion could easily displace one or more of the
+20 water molecules that appear necessary to suppress the loss of a
proton.
XC and AJS would like to thank the University of Notting-
ham for nancial support.
References
1 C. F. Baes, Jr. and R. E. Mesmer, The Hydrolysis of Cations,
J. Wiley, New York, 1976.
2 D. F. Shriver and P. W. Atkins, Inorganic Chemistry, Oxford
University Press: Oxford, 1999.
3 See for example, P. L. Brown, R. N. Sylva and J. Ellis, J. Chem.
Soc., Dalton Trans., 1985, 723.
4 J. E. Huheey, E. A. Keiter and R. L. Keiter, Inorganic Chemistry,
Harper Collins, New York, 1993, p. 327.
5 D. T. Richens, The Chemistry of Aqua Ions, J. Wiley, Chichester, 1997.
6 D. W. Barnum, Inorg. Chem., 1983, 22, 2297.
7 M. Beyer, E. R. Williams and V. E. Bondybey, J. Am. Chem. Soc.,
1999, 121, 1565.
8 H. Cox and A. J. Stace, J. Am. Chem. Soc., 2004, 126, 3939.
9 A. J. Stace, J. Phys. Chem. A, 2002, 106, 7993.
10 B. J. Duncombe, K. Duale, A. Buchanan-Smith and A. J. Stace,
J. Phys. Chem. A, 2007, 111, 5158.
11 K. McQuinn, F. Hof, J. S. McIndoe, X. Chen, G. Wu and
A. J. Stace, Chem. Commun., 2009, 4088.
12 X. Chen, G. Wu, B. Wu, B. J. Duncombe and A. J. Stace, J. Phys.
Chem. B, 2008, 112, 15525.
13 A. T. Blades, P. Jayaweera, M. G. Ikonomou and P. Kebarle,
J. Chem. Phys., 1990, 92, 5900; A. T. Blades, P. Jayaweera,
M. G. Ikonomou and P. Kebarle, Int. J. Mass Spectrom. Ion
Processes, 1990, 102, 251; M. Peschke, A. T. Blades and
P. Kebarle, Int. J. Mass Spectrom., 1999, 185187, 685.
14 A. A. Shvartsburg and K. W. M. Siu, J. Am. Chem. Soc., 2001,
123, 10071.
15 J. Burgess, J. Metal Ions in Solution,Ellis Horwood, Chichester, 1978.
16 R. D. Shannon and C. T. Prewitt, Acta Crystallogr., Sect. B:
Struct. Crystallogr. Cryst. Chem., 1969, 25, 925.
17 P. Kebarle, Annu. Rev. Phys. Chem., 1977, 74, 1466.
18 M. Pettine, F. J. Millero and G. Macchl, Anal. Chem., 1981,
53, 3041.
19 F. Se by, M. Potin-Gautier, O. F. X. Giaut and E. Donard,
Geochim. Cosmochim. Acta, 2001, 65, 3053.
20 K. G. Spears and F. C. Fehsenfeld, J. Chem. Phys., 1972, 56, 5698.
21 S. Feil, G. K. Koyanagi and D. K. Bohme, Int. J. Mass Spectrom.,
2009, 280, 38.
22 D. R. Carl, B. K. Chatterjee and P. B. Armentrout, J. Chem. Phys.,
2010, 132, 044303.
23 D. R. Carl, R. M. Moision and P. B. Armentrout, Int. J. Mass
Spectrom., 2007, 265, 308.
24 A. J. Stace and A. K. Shukla, J. Am. Chem. Soc., 1982, 104, 5314.
25 A. J. Stace and C. Moore, J. Am. Chem. Soc., 1983, 105, 181.
26 M. Trachtman, G. D. Markham, J. P. Glusker, P. George and
C. W. Bock, Inorg. Chem., 2001, 40, 4230.
27 M. F. Bush, R. J. Saykally and E. R. Williams, J. Am. Chem. Soc.,
2008, 130, 9122.
28 R. G. Pearson, J. Am. Chem. Soc., 1963, 85, 3533.
29 R. G. Pearson, Coord. Chem. Rev., 1990, 100, 403.
30 See also S. J. Lippard and J. M. Berg, Principles of Bioinorganic
Chemistry, University Science Books, California, 1994; J. J. R.
Frau sto and R. J. P. Williams, The Biological Chemistry of the
Elements, Oxford University Press, New York, 2001.
31 H. Irving and R. J. P. Williams, J. Chem. Soc., 1953, 3192.
P
u
b
l
i
s
h
e
d

o
n

0
6

S
e
p
t
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T
Y

O
F

B
R
I
G
H
T
O
N

o
n

2
7
/
1
0
/
2
0
1
3

1
3
:
5
8
:
2
2
.

View Article Online

Anda mungkin juga menyukai