Thermodynamics and Kinetics of Adsorption Schmoeckwitz
Thermodynamics and Kinetics of Adsorption Schmoeckwitz
Thermodynamics and Kinetics of Adsorption Schmoeckwitz
subtitle:
What can we learn about adsorption systems from thermodynamic and kinetic measurements?
IMPRS-Lecture Series 2012: Experimental and Theoretical Methods in Surface Science
Klaus Christmann
Institut fr Chemie und Biochemie, Freie Universitt Berlin
Organization
1. Introduction; terms and definitions
2. Thermodynamics of adsorption
Coverage dependence of the adsorption energy: lateral interactions and a-posteriori heterogeneity
Strength of interaction
van-der-Waals
ionic
covalent
Strong (quantum-chemical
origin)
> 50 kJ/mole
metallic
Strong (quantum-chemical
origin), E > 50 kJ/mole
e
e
t
Consequence:
In the surface region, the interaction forces
responsible for the bulk properties become
asymmetric, since the outermost bonds are
unsaturated, resulting in a peculiar chemical
reactivity of the surface. In our context the most
important consequence is that the surface becomes
capable of adsorbing foreign atoms or molecules.
Consequences: Clean surfaces try to rearrange their atoms positions in order to minimize the surface free energy.
One distinguishes:
multilayer relaxation, i.e., in the surface region the distances perpendicular to the surface differ from those of the bulk
reconstruction, i.e., the surface atoms take on new periodic positions parallel to the surface (formation of superstructures)
Both phenomena can (and usually will) be affected by adsorption.
Free energy as a
function of the
surface
configuration.
This
configuration is
determined by
the adsorbate
coverage. At a
critical value of
this coverage,
the surface flips
to a more stable
equilibrium
configuration.
pconvex
p
V
2
= exp
p0
RT rp , Kelvin
Kelvin equation
Information on the mechanism of adsorption may be obtained from measurements of adsorption isotherms or from
thermal desorption spectra as will be discussed later.
early barrier
late barrier
The dissociation of
molecular hydrogen has
been studied theoretically in
great detail, for example on
Pd(100) surfaces by A. Gro
et al.. Quite important is the
so-called dynamical
steering where the H2
molecule is guided to the
optimum dissociation sites.
Consider the differential change of Free Enthalpy (Gibbs Energy), dG, of a thermodynamic
system during any change of state:
G
G
G
G
+
+
+
dG =
dP
dT
dA
dni
A P ,T
P T , A
T P , A
ni T ,P , A
dG = V dP S dT + dA + i dni
i
= + V;
P T , A
G
= S
T P , A
G
=
A P ,T
with P = pressure, T = temperature, A = surface area, = surface tension, V = system volume, = chemical potential
and S = entropy. The third term becomes decisive, when the surface area is large in relation to the bulk volume (high
degree of dispersion). Example: Raney-Nickel as a catalyst is often utilized in hydrogenation reactions.
The surface tension [N/m] is entirely equivalent to the surface energy [Nm/m2], which is the driving
force for all surface phenomena.
dn
i P ,T , other mole numbers
i =
In the (dynamical) phase equilibrium, the chemical potentials ad and gas are equal and remain
equal (persisting or ongoing equilibrium):
ad = gas
and
ad + OF
dad+OF = -sad+OFdT + vad+OFdP + d
p ,T
Rearranging and considering that the term ad +OF d gets zero for constant coverage (d=0) yields
p ,T
the well-known Clausius - Clapeyron
equation for the ongoing phase equilibrium
between gas phase and adsorbate phase(s):
s g sad +OF
S
dP
=
=
dT
V
Vg Vad +OF
Setting
Vad + OF << Vg
s g sad + OF
1 dP
=
P dT
RT
RT
P
h
finally
T
hg had + OF d ln P
1 dP
qst
=
=
=
2
P dT
RT 2
dT RT
qst is called differential isosteric heat of adsorption and represents the energy difference between
the state of the system before and after the adsorption of a differential amount of particles on that
surface. It can be obtained from equilibrium, i.e., adsorption isotherm measurements. These
considerations apply particularly well for homogeneous single crystal surfaces, for which we shall
provide examples later. Nevertheless, qst is often coverage-dependent due to lateral particle particle
interactions.
Inhomogeneous, porous surfaces that one encounters in practical heterogeneous catalysis require a
modified view and more complicated thermodynamic treatment, since they can provide a whole variety
of energetically and geometrically different adsorption sites.
2. Thermodynamics of adsorption:
Concept of the adsorption isotherm
The adsorption isotherm displays information of how much material will adsorb for a
given set of state variables (pressure P and temperature T when using gases as
adsorptives, concentration c and temperature T when dealing with adsorption from
solutions). The amount of adsorbed material can be introduced as surface coverage ,
i.e., the fraction of occupied/available sites, or by the total volume vm taken up by the
sum of particles adsorbed in the first monolayer.
The adsorption isotherm is one of the most characteristic quantities for an adsorption
process. In practice, the uptake of adsorptive of a given catalyst material is decisive,
since a high uptake usually means a good catalyst efficiency. The respective uptake is
often called sorption capacity and characterised by the sorption (or adsorption)
isotherm. The form of the sorption isotherm provides a lot of first-glance information
about the chemical and physical properties of the catalyst material and about how the
adsorption process proceeds. In principle, various kinds of isotherms can be
distinguished, depending on the nature of the catalyst and the kind of interaction.
2. Thermodynamics of adsorption:
The concept of the adsorption isotherm
Langmuir isotherms (type I) are usually observed during adsorption of chemically active gases on
metal surfaces, but also with non-polar gases (methane, nitrogen N2, ethane) on zeolites, whereas
the adsorption of water vapour on -Al2O3 leads to isotherms of type II with condensation. Using an
unpolar surface like charcoal and polar molecules as adsorptive (methanol or H2O) the situation is
characterised by initially rather repulsive interactions leading to a reduced uptake, while the
increasing presence of adsorbate molecules facilitate the ongoing adsorption leading to isotherms of
type III. If porous adsorbents are used, additional capillary condensation effects lead to isotherms of
type IV and V.
Monolayer adsorption on a
homogeneous surface at equilibrium
pressure p/p0 . The heat of adsorption
of the first monolayer is much
stronger than the heat of adsorption
of the second (and all following)
layers (typical chemisorption case)
Multilayer adsorption/condensation
on a homogeneous surface at
equilibrium pressure p/p0 . The heat
of adsorption of the first (blue) layer
is comparable to the heat of
condensation of the subsequent
(red) layers. Often observed during
physisorption.
A way of getting to the adsorption isotherm is to consider the adsorption - desorption equilibrium.
The adsorption process between gas phase molecules, A, vacant surface
sites, S, and occupied surface sites, SA, can be described by the equation
S + A SA
assuming that there are a fixed number of surface sites present on the
surface. An equilibrium constant, K, can be written:
K=
[SA]
[A] [S ]
If we introduce the symbol for the fraction of surface sites occupied (0 < < 1) and consider that [SA] is proportional to
the surface coverage of adsorbed molecules, i.e., proportional to , whilst [S] is proportional to the number of vacant
sites, 1 - , and [A] is proportional to the collision number, i.e., the gas pressure P, we arrive at
an expression:
K P
K=
=
or, rearranged for ,
(1 )P
1+ K P
which has been derived as early as 1916 by Irving Langmuir. Note that K is only a constant if the enthalpy of adsorption
is independent of coverage. As with all equilibrium constants, K and, hence, the position of adsorption - desorption
equilibrium, will depend on (i) the relative stabilities of the adsorbed and gas phase species involved, (ii) on the
temperature of the system, and (iii) on the pressure of the gas above the surface. Factors (ii) and (iii) exert opposite
effects on the concentration of adsorbed species - that is to say that the surface coverage may be increased by raising
the gas pressure but will be reduced if - at constant pressure - the surface temperature is raised.
A somewhat more revealing derivation of the Langmuir adsorption isotherm is based on the fact that in
adsorption - desorption equilibrium the rates of adsorption and desorption are equal. While the rate of
adsorption is determined by the collision frequency of the gas phase particles with the surface
according to:
Rad =
particles
P
(1)
S 0 = k ad
P S0
2
2 mkT
m s
(S0 stands for the uncovered fraction of the surface ( (1 - )), P for the gas pressure), the desorption
rate can easily be formulated as being proportional to the fraction of the surface covered by adsorbed
particles (here denoted as S1, i.e., ), a frequency factor , and an exponential term containing the
depth of the potential well, i.e., the adsorption energy (Arrhenius equation):
particles
E
Rdes = S (1) exp des = S (1) k des (T )
2
kT
m s
Equating Rad and Rdes and introducing the coverage by noting that S1 and S0 (1 - ) yields
immediately the Langmuir expression derived before.
s0
Edes
12
(2 mkT ) exp ( )
K (T ) =
kT
N max
with s0 representing the initial sticking coefficient, the frequency factor of the desorption reaction
and Nmax the maximum number of adsorbed particles: For non-activated adsorption, the
desorption energy Edes is equal to the energy of adsorption, Ead. In the derivation given above we
have assumed first-order processes both for the adsorption and the desorption reaction. For
second-order processes, e.g. adsorption/desorption reactions involving rate-limiting dissociation
and recombination steps, a similar expression can be derived
( p )T =
K ' (T ) P
1 + K ' (T )P
From the initial slope of a log - log plot of a Langmuir adsorption isotherm (K(T)P << 1), the order of
adsorption can easily be determined (slope 1 = 1st order adsorption; slope 0,5 = 2nd order adsorption).
If we plot the surface coverage against the concentration or the pressure of particles of the adsorbing phase, we arrive at
typical saturation curves: Near the monolayer capacity all surface sites are occupied, and higher pressures cannot
increase the coverage or surface concentration beyond = 1 anymore.
the rate of adsorption into the ith layer (clean surface = index o)
(i )
(i )
Rad
= k ad
P S(i )
(i )
(i )
Rdes
= k des
S(i ) e
E( i )
kT
E( 1 )
kT
(eq. I )
(1)
(1)
Rad
= k ad
P S0
(1)
(1)
Rdes
= k des
S (1) e
( 2)
( 2)
Rad
= k ad
S (1) P
( 2)
( 2)
Rdes
= k des
S( 2) e
E(1 )
kT
E( 2 )
kT
In equilibrium, however, the surface area S1 does NOT change, and all four processes taken together
must not change the overall concentration c. Hence, processes I and IV increase S1, while II and III
decrease it:
E( 2 )
E(1 )
k PS ( 0 ) + k
(1)
ad
( 2)
des
S( 2)e
kT
=k
(1)
des
S (1) e
kT
( 2)
+ k ad
S (1) P
(eq. (II))
( 2)
des
S( 2) e
E( 2 )
kT
( 2)
= k ad
S (1) P (eq. III )
Following this scheme, one realizes that we can continue these considerations for all following layers
and obtain a general expression of the kind:
( i +1)
k des
S (i +1) e
E( i +1)
kT
( i +1)
= k ad
S ( i ) P (eq. IV )
for all surface parts Si, with i = 1, 2, 3.... The total volume adsorbed is obtained when we multiply the
surface part Si of each layer with the number of layers below and additionally multiply this with the
specific volume per surface part S, V0
Vtotal = V0
i S
i =1
Vtotal
=
= i =1
Vmono
i Si
S
i =1
In order to carry out these summations we have to make some (plausible) assumptions: First, all
adsorption energies of the second and all higher layers will be set equal = EL, and second, we assume
that all the ratios of adsorption/desorption constants are equal to a constant g:
( 2)
( 3)
(n)
k des
k des
k des
= (3) = ... = ( n ) = g
( 2)
k ad k ad
k ad
as expected for a liquid-like behaviour. Note that the very first adsorbed layer is treated differently,
because its rate constants as well as its adsorption energy is largely different from those of the higher
layers. By defining
(1)
x=
k
P
E
E1
exp L with L 2 and y = ad
exp
(1)
k
RT
g
RT
des
and introducing the constant c = y/x we can express all surface parts Si by S0 according to:
Si = x Si - 1 = xi - 1 S1 = c xi S0
If we define the ratio of the total volume adsorbed / the volume of the monolayer as coverage
V
= total =
Vmono
c S0 i x
i =1
S 0 1 + c x i
i =1
c i x
i =1
1 + c xi
i =1
xi =
i =1
x
1 x
d i
d x
x
i
x
x
x
x
x
x
x
x
=
+
+
+
=
+
+
+
=
x
=
x
2
3
...
1
2
3
...
=
i
dx i =1
dx 1 x (1 x )2
i =1
i =1
1
Vtotal
cx
1
cx
cx
=
=
=
=
2
2
cx
1
x
+
cx
Vmono (1 x ) (1 + 1 x ) (1 x ) ( 1 x ) (1 x ) (1 x + cx )
This expression can be rearranged to yield:
1
Vtot
c 1 1
1
x
+
=
x
c Vmono
1 x Vmono c
The unknown quantity x can be substituted by the fact that the volume Vtotal must become infinity for p =
p0. Since Vmono is always finite, x has to become 1 for this case. Otherwise, X = p/p0 . Since p as well as
p0 are measurable quantities, we have thus eliminated the unknown quantity x and finally arrive at the
famous BET isotherm in the common form:
1 p
1
c 1 1
p
=
+
Vtotal p0 p c Vmono
c Vmono p0
Intensities of adsorbate-specific electronic excitations (Auger, UV- and X-ray hotoelectron spectroscopy)
Intensities of adsorbate-specific vibrational modes (IR, HREELS)
Intensities of adsorbate-induced LEED superstructures (not always unambiguous)
Changes of the surface work function induced by the adsorbate [Tracy & Palmberg, JCP 51 (1969) 4852]
Changes of the surface conductivity (thin films!)
Changes of the paramagnetic resonance
In each case it is necessary to calibrate the respective quantity against the actually adsorbed
amount, e.g., by thermal desorption spectroscopy or (in case of solids) by a quartz microbalance
etc.). It is helpful, if the quantity measured is linearly correlated with the adsorbed amount.
In the following it will be exploited that the adsorbate-induced work function change
is often a quite convenient indirect monitor of the adsorbed amount, since it measures
the sum of the surface dipole moments of the adsorbed particles. The physical basis is
the validity of the Helmholtz equation
= 0
N ad
0 A
The adsorption of the noble gas Xe decreases the W.F. of Ni(100) by ca. 340 meV. Stepwise increase of PXe at a
fixed temperature (93,6 K) shifts the adsorption - desorption equilibrium towards higher Xe coverages, which can
be followed by the corresponding W.F. changes. Stopping the xenon pressure (arrow) leads to complete removal of
the Xe atoms by thermal desorption. Next the corresponding pairs of (~ ) and P are plotted as a 93,2 K
isotherm [K. Chr. & J.E. Demuth, Surf. Sci. 120 (1982) 291].
Note: The proportionality between and has been checked independently by TDS and UPS (not shown here).
One immediately realises that the adsorption - desorption equilibrium leading to a given Xe coverage (e.g., = 0,5)
can be achieved by different pressures and temperatures. This leads us to the important quantity hidden in the
adsorption isotherm, namely, the differential heat of adsorption: The strength of the adsorbate bond determines
how high the temperature must be to remove an adsorbed particle from the surface and to what extent one has to
increase the pressure in order to push this particle back to the surface.
At this point we recall the Clausius-Clapeyron equation describing the adsorption - desorption equilibrium:
hg had + OF d ln P
qst
1 dP
=
=
=
2
2
P dT
RT
dT RT
hg had +OF d ln P
q
1 dP
= st
=
=
2
1
P dT
RT
R
d ( T )
Plotting the ln P vs. the inverse absolute temperature should yield, for each coverage , straight lines with slope
- qst/R (left figure). In this way one obtains the heat of adsorption for all coverages of interest (right figure). This
important graph carries the information about the energetic homogeneity of the surface in question (see later).
Example: Hydrogen on a Nickel(110) surface [K.Chr. et al., J.Chem. Phys. 60 (1974) 4528]
Adsorption isotherms
Adsorption isosteres
In calorimetric
measurements certain
gas portions are
subsequently admitted
to the system while the
heat released is
monitored by a T
increase. Accordingly,
the heat portions sum
up to the integral heat of
adsorption. If the
surface is subjected to a
large gas portion
leading to saturation,
the whole integral heat
of adsorption is
liberated.
Measured heat of Ag atom adsorption on oxide surfaces, versus the diameter of the Ag particle to which it adsorbs, for four different
surfaces: two 4-nm-thick CeO2(111) films with different extents of surface reduction (CeO2-x , x = 0.1 and 0.2), a 1-nm-thick CeO 1.9(111)
film, and a 4-nm-thick MgO(100) film. The first three films were grown on Pt(111), and the fourth was grown on Mo(100). (Right)
Structural models for perfect CeO2(111) and MgO(100); black lines indicate their unit cells. After Farmer JA, Campbell CT. 2010. Ceria
maintains smaller metal catalyst particles by strong metal-support binding. Science 329:93336
Surface
Ru(0001)(1x1)-H
Ru
Cu
Ni
Pd
Pt
Co, Ni, Ru, Rh, Pd
Mo, W, Re
Ni, Pd
Ru, Pd
Rh, Re
E0 from...to [kJ/Mol]
2...5
10...20
10...30
15...25
20...40
60...80
30...100
125...175
100...150
200...250
200...300
copper Cu
Ru, W, Re
250...330
On a given surface there can inherently exist adsorption sites with different
local geometry providing different chemical coordination; good examples
being surfaces with regular steps. Usually, the strength of adsorbate
substrate interaction increases with the coordination number. Example:
H/Pd.
Real-space model
of a real surface
with various defects
(steps, kinks, holes
etc.)
Coverage dependence
of Ead for Pd(111)/H: full
circles: smooth (111)
surface; open circles:
periodically stepped
Pd(111) surface (after
Conrad et al.)
Consider the rate of an ordinary chemical reaction of type A B. The general definition of the
reaction rate is change per time, i.e. number of molecules per time, or concentration change of a
certain species per unit time interval:
Rate R =
moles
dN A
dN P
particles
d [ A]
d [ P]
]
=+
in [ 3 ] or R =
=+
in [
dt
dt
dm s
dt
dt
Volume s
dN S , A
dN S , P
particles
d
moles
d [ A]
d [ P]
=+
=+
]=
in [
in [ 2 ] or R =
Rate R =
unit area s
dt
dm s
dt
dt
dt
dt
Often, rates are expressed in terms of change of coverage =
N ad
N ad , max
In an adsorption experiment, an initially bare surface is exposed to a certain gas pressure P, and the
rate of collision of the gas particles with the unit surface is given by kinetic theory:
Collision frequency
dN
=
A dt
P
particles
2
2 mkT m s
The decisive process now is the so-called trapping or sticking: Will the impinging particle stay on the
surface or will it be reflected? The respective probability is called (initial) sticking probability s0 and
varies between zero and one. As the surface is gradually covered, s decreases simply because the
number of empty adsorption sites becomes smaller.
If an adsorbed particle statistically occupies
a single site, s() = 1 - ;
if it dissociates, two sites are blocked by one collision event, and accordingly
s() = (1- )(1- ) = (1 - )2 2nd order adsorption.
*
Sometimes, the adsorption reaction requires a certain activation energy Ead . Only particles having at
least this energy will be able to stick.
(t ) = 1 e
k ad t
, with
k ad =
s0
N max
P
2 mkT
which is a typical saturation function. The larger the initial sticking coefficient, the more rapidly the
surface becomes covered. Metal single crystal surfaces (Ni) contain ca. 1019 adsorption sites/m2.
A rough estimate neglecting the coverage dependence of the sticking function (constant unity
sticking probability) yields that a surface would be completely covered in one second, if one
maintains a pressure of ~10-6 mbar.
The initial sticking probability s0 is an interesting quantity; it contains all dynamical and steric effects
and is governed by the ability or effectiveness of a given particle to dissipate its kinetic energy to the
heat bath of the surface (phonon excitation; electron hole pair excitation). The sticking of diatomic
molecules, hydrogen in particular, is of special interest, since the sticking at finite temperatures includes
the dissociation reaction (activated and non-activated dissociation paths).
An extensive consideration of the quantum-dynamical background of sticking can be found in the book
by A. Gro (Theoretical Surface Science, Springer Berlin 2003)
Complications can arise when the colliding particle is trapped for some time (typically microseconds) in
a weak potential in which it can freely move across the surface and search for an empty adsorption site.
This weakly bound state is called a precursor. Major consequence of a precursor state: The sticking
coverage function f() is no longer linear, but has a convex shape: At not too large coverages the
sticking remains high, but as the diffusion length in the precursor state becomes shorter than the mean
diameter of the already covered area (island), the particles can no longer be accommodated and are
finally reflected back into the gas phase before they can find an empty site.
pd"
K = '
pch + pd'
s ( ) = s0
1
1 + 1 K
The removal of particles from the adsorbed state into the gas phase is called desorption. It can be achieved by
thermal energy (thermal desorption, temperature-programmed desorption), electron impact (EID, DIET), ion impact,
resonant photon irradiation etc. Here, only thermal desorption is considered. Again, desorption is understood as a
normal chemical reaction and described by the respective kinetic formalism: Aad Agas with rate constant kdes.
particles
dN ad
= k des N adx
2
dt
m s
= Nad/Nmax
we obtain:
d
x 1
= k des x N max
dt
As in any chemical reaction with an activation barrier, kdes can be expressed as a product of a pre-exponential factor,
des , and an exponential term containing the activation energy :
Edes
k des = exp
kT
Inserting this in the first equation yields the well-known Polanyi-Wigner equation which is the basis for a determination
of both energetic and kinetic quantities from a thermal desorption spectrum:
Rate of desorption: R = R (, T)
Edes
d
kT
x
x 1
= x N max e
dt
R = rate of desorption: R = R (, T) =
Edes
d
kT
x
x 1
= x N max e
dt
The desorption rate of Cu from Re(0001) as a function of two variables T and [R. Wagner + K.C. Surf. Sci. 469 (2000) 55]
D. Schlatterbeck et al,
Surf.Sci. 418 (1998) 240