Download 1. Interaction of x

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Circular dichroism wikipedia , lookup

Radiation wikipedia , lookup

Density of states wikipedia , lookup

Nuclear physics wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

DESY wikipedia , lookup

Metastable inner-shell molecular state wikipedia , lookup

X-ray photoelectron spectroscopy wikipedia , lookup

Transcript
X-ray spectroscopy part 1
INTRODUCTION ..................................................................................................................................... 1
SYNCHROTRON RADIATION ................................................................................................................. 4
1.
INTERACTION OF X-RAYS WITH MATTER .................................................................................. 6
1.1.
X-RAY ABSORPTION SPECTROSCOPY ............................................................................................. 8
1.2.
BASICS OF XAFS SPECTROSCOPY............................................................................................11
1.3.
XANES .......................................................................................................................................15
1.4.
XANES SPECTRAL SHAPE ANALYSIS.........................................................................................17
1.5.
X-RAY ATTENUATION LENGTHS ...................................................................................................18
1.6.
DETECTION TECHNIQUES ..............................................................................................................19
Transmission detection .........................................................................................................................19
Fluorescence Yield ................................................................................................................................20
Electron yield ........................................................................................................................................22
Ion Yield ................................................................................................................................................24
Introduction
X-rays are defined as the electromagnetic radiation with a wavelength between
Ultraviolet radiation and gamma radiation. The border between Ultraviolet radiation
and X-rays is not clearly defined but usually set around 10 -8meter, i.e. 10 nm. The
upper limit of x-rays is usually set at 10-12 meter, i.e. 1 pm, although radiation with
an energy above 10-11 meter is also called gamma radiation, in particular if it is
created with a radioactive element. X-rays are sub-divided into soft x-rays, between
10 nm and 100 pm and hard x-rays between 100 pm and 1 pm. Soft x-rays are
called 'soft' because they do not penetrate air and as such are relatively safe to work
with. We will discuss the penetration length of x-rays below. Because of historical
reasons, x-rays are often given in Angstroms, where 1 Å equals 10 -10 meter or 100
pm.
The bottom two axes in the figure give the frequency and the energy of the
electromagnetic radiation, using the inverse proportionality between frequency  and
wavelength , i.e. =c/, where c is the speed of light 3·108 m/s. This yields x-ray
frequencies between 1016 and 1021 Hertz. It is more common to indicate x-rays with
their energies E, using E=h, with the energy given in electron volts (eV) or kilo
electron volts (keV), where they range between 100 eV and 106 eV.
1
X-ray spectroscopy part 1
Calculate the energy in electron volt for a x-ray with 1 Å wavelength.
Use: E= h= hc/, with c=3·108 m/s, h=6.6·10-34 and 1 eV = 1.6·10-19 J
E (1 Å)= 6.6·10-34  3·108 / 10-10 = 2·10-15 J /1.6·10-19 = 1.2·104 eV=12.4 keV
It can be checked in the figure that
energy slightly above 104 eV.
a
wavelength of 10-10meter corresponds to
The wavelength of x-rays are of the same size as molecules, which range from 100
pm for small molecule such as water to 10 nm for macromolecules such as enzymes.
In particular the distance of 1 Å is important in matter, because it corresponds to a
typical interatomic distance. This makes diffraction with x-rays of approximately 1 Å
an important technique to determine the interatomic distances and, in general, the
structure of matter.
X-rays are traditionally produced using x-ray tubes. In an x-ray tube, electrons are
accelerated between the anode and the cathode. When they hit the cathode with
high energy, they are decelerated through inelastic collisions with atoms in the target
material. If the energy of the electrons is high enough this produces a continuous
spectrum of x-rays. In addition, the impinging high-energy electrons excite core
electrons from the cathode material. Core holes decay, in part, via radiative
processes in which a shallow core electron fills a deep core hole, thereby emitting xrays. These x-rays are characteristic for the cathode material and are confined to
only a few energies. This process will be discussed in detail below.
2
X-ray spectroscopy part 1
X-ray tubes are used in many x-ray applications; for example (almost) all medical xray images use x-ray tubes. In addition, laboratory x-ray sources used in x-ray
photoemission and x-ray diffraction make use of x-ray tubes. In x-ray spectroscopy
however the role of x-ray tubes has been declining over the last few decades. The
reason is evident from the figure: X-ray tubes produce a maximum of 1010 'photons
per second', whereas x-rays produced with synchrotrons reached up to 10 19 photons
per second. These photon counts are usually given with the 'brightness'. The
brightness is normalised to an area of 1 mm 2 and further normalised to the x-ray
beam divergence and the energy (range). It is clear that with 10 9 times more
photons per second, one can carry out much faster experiments. In addition, it allows
a range of new experiments that have very low x-ray cross-sections. It should be
noted that the x-rays produced with synchrotrons have very high intensity, but still
they cannot be considered as x-ray lasers, because in general they have insufficient
coherence.
3
X-ray spectroscopy part 1
Synchrotron Radiation
Experimental: Generating x-rays with Synchrotrons
The Intensity of x-ray sources
X-ray tubes:
Wilhelm Conrad Röntgen discovered x-rays in 1895 whilst working with cathode-ray
tubes. Using the principle of fast electrons hitting a metallic target, a first substantial
gain in brilliance was not obtained until the introduction of rotating anode sources
(1960).
Synchrotron Radiation Facilities:
The progress of high energy physics, with the construction of powerful particle
accelerators gave birth to what we now call First generation synchrotron sources
(1970). Using the deflection of high energy electrons by a magnetic field for the
production of x-rays proved so promising that a number of dedicated Second
generation sources were built (1980). Relying on the combination of needle thin
electron beams and Insertion Devices, Third generation synchrotron sources (1995)
are now emitting synchrotron x-ray beams that are a trillion (1012) times more
brilliant than those produced by x-ray tubes.
(Based on the ESRF www-page: http://www.esrf.fr/ )
Initially synchrotron radiation was seen as an annoying side product of charged
particle accelerators. When charged particles are accelerated they start to produce
radiation: the so-called synchrotron radiation. Only after the theory about EXAFS was
established dedicated synchrotrons for radiation studies (like EXAFS) were built.
In a synchrotron charged particles are generated and accelerated to the desired
energy before they are stored in a ring: the so-called storage ring. Without the
presence of an external magnetic field the electrons will travel in a straight line but
when a magnetic field is applied they will bend caused by the Lorentz force. This
force causes an acceleration of the particles and therefore radiation is formed. Of
course this radiation will cause the kinetic energy of the electrons to reduce and
therefore the radius of the circular orbit will decrease. To prevent that after a short
period the electrons will have lost their energy, the energy loss is compensated by a
Radio Frequency cavity. It is important to mention here that not only x-rays are
produced during bending of the electron but also a much wider spectrum of radiation
is formed (depending on e.g. the magnetic field strength, the beam current, the
energy of the beam etc). Synchrotron radiation emerges at the bending magnets
where the orbit of the electrons is changed. Between the magnets the electrons
follow a straight path without the formation of radiation. Insertion devices can be
placed in the electron beam in order to produce radiation. Examples of such devices
4
X-ray spectroscopy part 1
are wigglers and undulators. Basically these devices ‘wiggle’ the electrons in the
straight path in order to produce radiation.
Synchrotrons: Radiation from a fast moving particle source appears to the observer
in the laboratory as being all emitted in the general direction of motion of the
particle. Think of tomatoes being thrown in all directions from a fast moving truck. In
synchrotron radiation sources (storage rings) highly relativistic electrons are stored
to travel along a circular path for many hours. Radiation is caused by transverse
acceleration due to magnetic forces in bending magnets (forming the circular path) or
periodic acceleration in special insertion device magnets like undulators and wigglers.
Bending magnet: Radiation is emitted tangentially to the orbit well collimated in the
nondeflecting, or mostly vertical plane. Because the geometry of the storage ring is
determined by bending magnets, it is not possible to freely choose the field strength
and the critical photon energy is therefore fixed.
Undulator: The electron beam is periodically deflected by weak magnetic fields. The
spectral resolution of the radiation is proportional to the number of undulator periods
and varying the magnetic field can shift its wavelength. Most radiation is emitted
within a very small angle.
Wiggler: Increasing the magnetic field strength causes the pure sinusoidal transverse
motion of electrons in an undulator to become distorted due to relativistic effects. For
very strong fields many harmonics are generated which eventually merge into a
continuous spectrum from IR to hard x-rays. The spectral intensity varies little over a
broad wavelength range and drops off exponentially at photon energies higher than
the critical photon energy. Compared to bending magnet radiation, wiggler radiation
is enhanced by the number of magnet poles.
(Based on the SSRL www-page: http://www-ssrl.slac.stanford.edu/welcome.html)
A XAFS spectrum originates from the fact that the probability of an electron to be
ejected from a core level is dependent on the energy of the incoming beam. For this
reason the energy of the x-rays is varied during an experiment. Of course this
requires a monochromatic beam. However the radiation generated by a synchrotron
is polychromatic. Therefore the desired wavelength has to be filtered from the
polychromatic beam. For this purpose a monochromator is installed in the beamline
(see sheet: XAFS beamline). The incoming beam is ‘reflected’ on a crystal due to
Bragg reflections. The wavelength of the reflected beam is given by n.  2.d.sin  .
-ray propagation direction and the surface of
the crystal, the energy of the x-ray beam can be varied. In principle one crystal is
enough to obtain the desired wavelength of the beam. In practice a double crystal
monochromator is used in order to obtain an outgoing beam that is parallel to the
incoming beam (sheet XAFS beamline). After the beam has passed the
monochromator, its intensity (I0) has to be measured. This can be done by an
ionization chamber. In the next step, the x-ray interacts with the sample of interest.
5
X-ray spectroscopy part 1
To measure the intensity after the sample a second ionization chamber can be used.
This is the transmission mode of x-ray absorption.
1.
Interaction of x-rays with matter
If an x-ray hits a solid material, the material starts to emit electrons. This is
essentially the photoelectric effect that also occurs with ultraviolet radiation. In the
case when a metal surface is exposed to ultraviolet or x-ray radiation, the following
experimental observations can be made:
a. No electrons are ejected unless the
radiation has a frequency above a
certain threshold value characteristic
of the metal.
b. In case of x-rays, there is always
some electron emission, but one
observes intensity jumps in the
emitted electrons at certain energies.
c. Electrons are ejected immediately,
however low the intensity of the
radiation.
d. The kinetic energy of the ejected
electrons increases linearly with the
frequency of the incident radiation.
These observations follow naturally from the quantum nature of electromagnetic
radiation; in fact the photoelectric effect was used by Einstein to (re)define
electromagnetism.
The characteristics of the photoelectric effect
can be explained in terms of photons. If the
incident radiation has frequency , it consists of
a stream of photons of energy h. These
particles collide with the electrons in the metal.
If the energy of a photon is less than the energy
required removing an electron from the metal,
then an electron will not be ejected, regardless
of the intensity of the radiation. The energy
required to remove an electron from the surface
of a metal is called the work function of the
metal and denoted . However, if the energy of
the photon is greater than , then an electron
is ejected with a kinetic energy, E=½mv2, equal to the difference between the energy
of the incoming photon and the work function. In case of x-rays, the explanation is
analogous, with the addition that a core electron is excited instead of a valence
6
X-ray spectroscopy part 1
electron. The binding energy of the core
energy is given as:
1
2
electron must be added
mv 2  
 EB
and the kinetic

In fact, this equation defines the binding energy of a core level and x-ray
photoemission experiments are used to measure core level binding energies. The
binding energy EB is closely related to the ionisation energy that is defined as EB+.
Example: In an x-ray photoemission (XPS) experiment, NiO is excited with al K
radiation with an energy of 1487 eV. One observes peaks in the XPS spectrum at
respectively 607 eV and 637 eV. Assume that the work function of NiO is equal to 2
eV; calculate the binding energies of the core levels associated with the observed
peaks.
Solution: rewrite the equation to: EB =
h - EK - 
1487- 607-2=878 eV
1487-632-2= 853 eV
Explanation: These core level energies are both related to the 2p core level of the
nickel atoms in NiO. The 2p core level is split into two sub-peaks due to the 2p
spin-orbit coupling that will be explained later.
Self-Test; calculate the ionisation energy of a rubidium atom, given that radiation of
wavelength 58.4 nm produces electrons with a speed of 2450 km/s. The mass of an
electron is 9.1 10-31kg.
The photoelectric effect is not the only
interaction of x-rays with matter. Another
important interaction is the elastic scattering
of x-rays due to electrons, which is the origin
of x-ray diffraction experiments. Elastic
scattering is also known as Thompson
scattering. In a x-ray diffraction experiment
one hits a sample with a x-ray and measures
the diffracted x-rays. The diffraction angle
can be correlated with the interatomic
distances using Bragg's law.
In this course we focus on spectroscopy and x-ray diffraction will not further be
treated. With respect to spectroscopy, it is important to note that the amount of x-
7
X-ray spectroscopy part 1
rays involved in x-ray diffraction is almost constant with respect to the photon
energy. As such, it only gives an (almost constant) background upon which the
photoelectric effect is superimposed.
100
Intensity (log)
Photonuclear
Photoelectric
10
Thompson
Electron-
Compton
positron
1
100
1k
10k
Energy (eV)
100k
1M
This figure gives all contributions to the x-ray absorption process for a manganese
atom (in a gas or in a solid). The intensity axis is given on a logarithmic scale and it
can be seen that the photoelectric effect dominates up to approximately 50 keV. At
higher energies the photoelectric effect dies out and in addition to Thompson
scattering, Compton scattering becomes important. Compton scattering refers to
inelastic scattering. Around 1 MeV an additional x-ray process takes place, i.e. the
photonuclear absorption, which refers to the absorption of a x-ray in a nuclear
transition. This process is well known under the name of Mössbauer effect. At even
higher energies the photon energy is high enough to cause the creation of electronpositron pairs, which is the inverse process of the electron-positron annihilation as
used in positron emission tomography (PET) scans.
In the following we will consider x-ray energies up to 100 KeV and assume that only
Thompson scattering and the photoelectric effect take place.
1.1. X-ray absorption spectroscopy
A straightforward experiment is to measure the transmission of an x-ray through a
sample. A photon entering solid can be scattered or annihilated in the photo-electric
effect. All other photons will be transmitted and will leave the sample. In x-ray
absorption spectroscopy (XAS) the absorption of x-rays by a sample is measured.
8
X-ray spectroscopy part 1
The intensity of the beam before the sample I0 and the intensity of the transmitted
beam I are measured.
Consider an infinitely thin layer, of thickness, dl, of the absorbing material. The
intensity I0 of the incident beam is reduced by dI on passing through dl. dI is
proportional to dl and dI is also proportional to the total intensity I, i.e.:
dI    I  dl
 is a proportionality constant called the linear absorption coefficient. It incorporates
the combined effects of all photoelectric () and scattering () processes. One can
rewrite the equation to:
1
dI     dl
I
Integration over a finite thickness from 0 to x yields:
ln I
Ix
I0
x
0
   l
ln I x  ln I 0
 x
The linear absorption coefficient  is dependent on the x-ray energy (indicated with
its frequency ) and rewriting the equation yields:
I x ( )  I 0 ( ) e  ( ) x
alternatively, the linear absorption coefficient  can be given as:
 ( )  1x  ln
I0
It
The figure below indicates the intensity drop through sample of thickness x equal to
2/. This reduces the intensity after the
sample to e-2·I0.
The next task is to determine the absorption
cross section  as a function of the energy .
We assume that the scattering is small and
constant and limit ourselves to the
photoelectric effect. An x-ray acts on
charged particles such as electrons. As an xray passes an electron, its electric field
pushes the electron first in one direction,
then in the opposite direction, in other words
the field oscillates in both direction and
strength. In most descriptions of x-ray
absorption one uses the vector field A to describe the electromagnetic wave. The
vector field A is given in the form of a plane wave of electromagnetic radiation:
A  eˆq A0 cos(kx  t ) 
9
1
2
eˆq A0 e i ( kx t )
X-ray spectroscopy part 1
The cosine plane wave function contains the wave vector k times the displacement x
and the frequency  times the time t. The cosine function is rewritten to an
exponential function, for which only the absorbing e-it term has been retained. êq is a
unit vector for a polarisation q. (These polarisation effects will not be treated
explicitly below.) The Hamiltonian describing the interaction of x-rays with electrons
can be approximated in perturbation theory and its first term can be written as:
H1 
e
mc
p A
The first term of the interaction Hamiltonian H1 describes the action of the vector field
A on the momentum operator p of an electron. The electric field E is collinear to the
vector field, in other words this term can also be understood as the electric field E
acting on the electron moments. The proportionality factor contains the electron
charge e, its mass m and the speed of light c.
The Golden Rule states that the transition probability W between a system in its
initial state i and final state f is given by:
W fi 
2
 f  i
2

E
f
 Ei 
The initial and final state wave functions are built from an electron part and a photon
part. The photon part of the wave function takes care of the annihilation of a photon
in the x-ray absorption process. In the text below, it will not be included explicitly.
The delta function takes care of the energy conservation and a transition takes place
if the energy of the final state equals the energy of the initial state plus the x-ray
energy. The squared matrix element gives the transition rate.
The transition operator T1 describes one-photon transitions such as x-ray absorption.
It is in first order equal to the first term of the interaction Hamiltonian, i.e. T1=H1.
The transition rate W is found by calculating the matrix elements of the transition
operator T1. Two-photon phenomena, for example resonant x-ray scattering, are
described with the transition operator in second order. The interaction Hamiltonian is
found by inserting the vector field into H1:
T1  

q
e
m
2
V k
a eˆ
k
kq
 p e ikr

k
This equation can be rewritten using a Taylor expansion of e ikr=1+ikr+…. Limiting the
equation to the first two terms 1+ikr, the transition operator is:
T1  
q
a
e
k m
2
V k
eˆ
kq
 p   i(eˆkq  p)( k  r )

k
T1 contains the electric dipole transition (êkq·p). The electric quadrupole transition
originates from the second term. The value of kr can be calculated from the edge
energy in eV and the atomic number Z.
k r 
 ( edge)
(eV ) / 80Z
In case of the 1s core levels (called K edges) from carbon (Z=6, edge = 284 eV) to
zinc (Z=30, edge = 9659 eV), the value of kr lies between 0.03 and 0.04. The
transition probability is equal to the matrix element squared, hence the electric
10
X-ray spectroscopy part 1
quadrupole dipole transition is smaller by approximately 1.510-3 than the dipole
transition and can be neglected, which defines the well-known dipole approximation.
The transition operator in the dipole approximation is given by:
T1  
q
a
e
k m
2
V k
eˆ
kq

 p
k
Omitting the summation over k and using the commutation law between the position
operator r and the atomic Hamiltonian (p=m/i[r,H] ) one obtains the familiar form of
the x-ray absorption transition operator:
T1   e
q
2
eˆq  r
V


eˆq  r
q
1.2. Basics of XAFS Spectroscopy
Including this operator into the Fermi golden rule gives:
W fi  
 f eˆq  r  i
2
E
f
 Ei  
q
This equation will form the basis for the rest of the lectures. It is a rather theoretical
and formal description. What happens in practice? If an assembly of atoms is
exposed to x-rays it will absorb part of the incoming photons. At a certain energy
(depending on the atoms present) a sharp rise in the absorption will be observed
(Figure 1.2). This sharp rise in absorption is called the absorption edge.
Figure: The x-ray
absorption
cross
sections
of
manganese
and
nickel. Visible are
the L2,3 edges at
respectively
680
and 830 eV and
the K edges at
respectively 6500
eV and 8500 eV.
The energy of the absorption edge is determined by the binding energy of a core
level. Exactly at the edge the photon energy is equal to the binding energy, or more
11
X-ray spectroscopy part 1
precisely the edge identifies transitions from the ground state to the lowest empty
state. The figure shows the x-ray absorption spectra of manganese and nickel. The
L2,3 edges relate to a 2p core level and the K edge relates to a 1s core level binding
energy. Many complicating aspects do play a role here and we will come back to this
issue later.
In the case of solids, when other atoms surround the absorbing atom, a typical XAFS
spectrum is shown below.
I n t en s it y
1.4
( a .u )
Figure: The L3 XAFS spectrum of
platinum metal. The edge jump is
seen at 11564 eV and above one
observes a decaying background
modulated by oscillations.
1.2
1
0.8
0.6
0.4
0.2
0
11000
11500
12000
12500
13000
En erg y
( eV )
Instead of a smooth background we see here changes in the intensity of the
absorption. Oscillations as function of the energy of the incoming x-rays are visible.
The question, which rises immediately, is what causes these oscillations?
An answer can be found by assuming that the electron excitation process is a one
electron process. This makes it possible to rewrite the initial state wave function as a
core wave function and the final state wave function as a free electron wave function
(). One implicitly assumes that all other electrons do not participate in the x-ray
induced transition. We will come back to this approximation below.
 f eˆq  r  i
2
  i c eˆq  r  i
Figure: The schematic
density of states of an
oxide. The 1s core
electron at 530 eV
binding
energy
is
excited to an empty
state: the oxygen p-
12
2
 
X-ray spectroscopy part 1
projected density of states.
The squared matrix element M2 is in many cases a number that is only little varying
with energy and one can often assume that it is a constant. The delta function of the
Golden Rule implies that one effectively observes the density of empty states ().
I XAS ~
M 2
The x-ray absorption selection rules determine that the dipole matrix element M is
non-zero if the orbital quantum number of the final state differs by 1 from the one of
the initial state (L =  1, i.e. s  p, p  s or d , etc.) and the spin is conserved (S
= 0). The quadrupole transitions imply final states that differ by 2 from the initial
state (L =  2, i.e. s  d, p  f or (L = 0, i.e. s  s, p  p). They are some
hundred times weaker than the dipole transitions and can be neglected in most
cases. It will be shown below that they are visible though as pre-edge structures in
the K edges of 3d-metals and the L2,3 edges of the rare earths.
In the dipole approximation, the shape of the absorption spectrum should look like
the partial density of the (L =  1) empty states projected on the absorbing site,
convoluted with a Lorentzian. This Lorentzian broadening is due to the finite lifetime
of the core-hole, leading to an uncertainty in its energy according to Heisenberg's
principle. A more accurate approximation can be obtained if the unperturbed density
of states is replaced by the density of states in presence of the core-hole. This
approximation gives a relatively adequate simulation of the XAS spectral shape, at
least in those cases where the interaction between the electrons in the final state is
relatively weak. This is often the case for 1s  4p transitions (K-edges) of the 3d
metals.
13
X-ray spectroscopy part 1
Web-exercise: A very useful web-page can be found at the Electronic Structures
Database: http://cst-www.nrl.navy.mil/es-access.html
This database includes a large set of common systems, in particular the pure metals,
some alloys, simple oxides, etc, from which the band structure and the density of
states curves can be viewed and downloaded.
Example: Search for the density of states and the band structure of Cu (in the fcc
structure).
Exercise: Search for the density of states and the band structure of TiO (in the NaCl
structure).
14
X-ray spectroscopy part 1
1.3. XANES
The single electron approximation as stated in section 1.3 gives an adequate
simulation of the XANES spectral shape if the interactions between the electrons in
the final state are relatively weak. This is the case for all K edges and to a very good
degree also for all other edges with binding energies above 3 keV (with a few
exceptions). We will assume this is correct and use the equivalence of I XAS and the
density of states as a starting point.
Figure: Schematic band picture of (a) a TiO6- cluster and (b) a FeO6- cluster. Ti4+ in
TiO2 has an empty 3d band and Fe3+ in Fe2O3 contains a half filled 3d band.
A usual picture to visualise the electronic structure of a transition metal compound,
such as oxides and halides, is to describe the chemical bonding mainly as a bonding
between the metal 4sp states and the ligand p states, forming a bonding
combination, the valence band and empty antibonding combinations. The 3d states
also contribute to the chemical bonding with the valence band that causes them to be
antibonding in nature. This bonding, taking place in a (distorted) cubic crystal-line
surrounding, causes the 3d states to be split into the so-called t2g and eg manifolds.
15
X-ray spectroscopy part 1
This situation is visualised for a TiIVO6 cluster in the figure above. For comparison the
oxygen 1s X-ray absorption spectrum is given. On the right this picture is modified
for a partly filled 3d band in a FeIIIO6 cluster. The 3d band is split by the crystal field
splitting and a large exchange splitting. This qualitative description of the electronics
structure of transition metal compounds can be worked out quantitatively within DFT.
The picture of the left is a
DFT
calculation
of
the
density of states of TiO2. The
total DOS is projected to
oxygen (O) and titanium (Ti)
respectively. The picture
below shows the broadened
oxygen DOS in comparison
with
the
experimental
spectrum. In line with the
dipole selection rule only the
oxygen p-projected DOS is
used.
We assume that the density of states of complex systems is calculated using the DFT
approximation. Programs to calculate the X-ray absorption spectral shape include
FEFF, GNXAS and Wien2k. These and other software packages can be found at
http://www.esrf.fr/computing/scientific/exafs/intro.html.
16
X-ray spectroscopy part 1
1.4. XANES spectral shape analysis
We have seen that the oxygen K edge x-ray absorption spectrum of TiO2 can be
simulated with the oxygen p-projected DOS as calculated by DFT. (Actually in case of
the oxygen K edge of TiO2, the core hole effect is small and can be neglected). The
similarity between x-ray absorption and the projected DOS allows a qualitative
analysis of the spectral shapes without the actual inclusion of calculations.
Above we give the oxygen K edges of
a series of transition metal oxides, ranging from Sc2O3 to CuO. Sc2O3 contains ScIII
and is formally a 3d0 compound, i.e. it has all ten 3d states empty. CuO is a 3d 10
compound and has all ten 3d-states occupied. One can distinguish two regions in the
spectrum: the dashed region is related to the empty 3d-band, actually to the oxygen
p-projected DOS inside this 3d-band. The structure at 540 eV is related to the empty
4s and 4p states of copper. The dashed region is split by the cubic crystal field
splitting and, in first approximation, one observes two peaks that are respectively
related to the t2g (xy, xz, yz) and eg (z2, x2-y2) orbitals.
17
X-ray spectroscopy part 1
1.5. X-ray attenuation lengths
Figure: The attenuation
length of x-rays through
carbon (solid), iron
(dotted) and lead (dashdotted).
The figure gives the attenuation lengths of x-rays from 300 eV to 10 keV through C,
Fe and Pb. X-rays with energies less than 1 keV have an attenuation length of less
than ~1 m, implying that for soft x-rays the sample has to be thin. The figure below
shows the transmission through 1m and 10m iron and it can be seen that soft xrays (<1 keV) have little to no transmission through 1m iron. At the iron K edge at 8
KeV the transmission through 10m iron drops from ~0.65% to less than 0.10%. The
extra absorption of the iron 1s core level takes away most of the x-rays in the first
10m. In addition soft x-rays have a large absorption cross section with air, hence
the experiments have to be performed in vacuum.
Figure: Transmission
through air and bulk
iron. Given are
respectively 10 cm
air (solid), 10 m air
(dashed), 1m iron
(dash-dotted) and
10m iron (dotted).
Questions:
18
X-ray spectroscopy part 1
-
Why is lead the best element to be used to protect against the transmission of xrays?
A synchrotron that yields x-rays up to 2 KeV is not protected with lead shielding.
Why is this not necessary?
Webpage:
- Go to the webpage of the Center for X-ray Optics (CXRO) in Berkeley
(http://www-cxro.lbl.gov/) and follow the link to X-ray tools.
- Check the attenuation lengths of copper and lead between 100 eV and the
maximum energy possible. Modify each plot also to a logarithmic scale.
- Check the transmission of x-rays for Be and quartz (SiO2) between 100 eV and
the maximum energy possible.
Questions:
- In Ultra-High-Vaccuum equipment one uses berylium windows. Why Be?
- To attenuate the synchrotron beams at energies above 3 KeV (for example for
radiation sensitive experiments) one uses aluminum foils. Why is Al used and not
for example Cu or Pb?
1.6. Detection techniques
Transmission detection
After the beam has passed the monochromator, its intensity (I0) is measured in an
ionisation chamber, or using a thin metallic foil or grid absorbing part of the photons.
The x-ray interacts with the sample of interest and the intensity after the sample is
measured with a second ionisation chamber or a photodiode. This is the transmission
mode of x-ray absorption. Transmission experiments are standard for hard x-rays.
An important limitation of transmission detection arises from the requirement for a
homogeneous sample. Variations in the thickness or pinholes are reasons for the socalled thickness effect that can significantly affect the spectral shape by introducing
non-linear responses. This is important in particular for the EXAFS analysis. The
combination of the short attenuation length with the thickness effect makes
transmission experiments not suitable for x-ray absorption below 1 keV. For the soft
x-ray range other detection modes must be used.
A special technique to measure transmission experiments is to use an energy
dispersive monochromator. The monochromator crystal is bent and a range of
energies is sent through the sample simultaneously. Each energy has its particular
angle of incidence and after the transmission through the sample the complete x-ray
absorption spectrum can be measured momentary with an array detector. Energy
dispersive experiments suffer from similar limitations as normal transmission
experiments with respect to attenuation length and thickness effect. A major
19
X-ray spectroscopy part 1
advantage of energy dispersive x-ray absorption is that the complete XAS spectrum
is obtained momentarily. It is clear that this yields interesting options for time
resolved measurements. Real time (non repeated) experiments can be probed with
typically 1 s time resolution. A typical energy dispersive XAS beamline, is beamline
ID24 at ERSF in Grenoble. See their website:
http://www.esrf.fr/UsersAndScience/Experiments/XASMS/ID24/
Fluorescence Yield
The decay of the core hole gives rise to an avalanche of electrons, photons and ions
escaping from the surface of the substrate. By measuring any of these decay
products, it is possible to measure samples of arbitrary thickness. In this section the
different methods are discussed, specifically with regard to the conditions under
which a specific yield measurement represents the x-ray absorption cross section and
the related question of the probing depth of the specific yield method.
An important prerequisite for the use of decay channels is that the channels are
linearly proportional to the absorption cross section. In general this linear
proportionality holds, but there are cases where for example the ratio between
radiative a non-radiative decay varies significantly over a relatively short energy
range. The reason behind this phenomenon is a varying fluorescence decay
depending on the symmetry of the final state in the x-ray absorption process. It turns
out that in cases where multiplet effects are important (discussed in chapter 3), for
example the 2p x-ray absorption of 3d metals, the fluorescence decay varies by
almost a factor of two over an energy range of approximately 15 eV, i.e. from the
first to the last peak in the edge region. Because for soft x-rays Auger decay
dominates over fluorescence, this effect is only visible in fluorescence yield detection.
The fluorescent decay of the core hole can be used as the basis for the absorption
measurement. The amount of fluorescent decay increases with energy and a
comparison with the amount of Auger decay shows that Auger decay dominates for
all core levels below 1 keV. In case of the 3d-metals, the K edges show strong
fluorescence and all other edges mainly Auger decay. The photon created in the
fluorescent decay has a mean free path of the same order of magnitude as the
incoming x-ray, which excludes any surface effect. On the other hand it means that
there will be saturation effects if the sample is not dilute.
20
X-ray spectroscopy part 1
For dilute materials the background absorption B dominates the absorption of the
specific edge and the measured intensity is proportional to the absorption coefficient.
For less dilute materials the spectral shape is modified and the highest peaks will
appear compressed with respect to the lower peaks, an effect known as selfabsorption or saturation.
Figure: The effect of
saturation due to selfabsorption on the
platinum L2 spectrum
measured in
fluorescence yield
assuming a relative
background
absorption of 1.5. The
transmission
spectrum is given with
the solid line. The
saturated (and renormalised)
fluorescence spectrum
with the dashed line.
The figure shows the
main effect of saturation due to self-absorption: a reduction of the peak heights and
depths and as such a blurring of the spectrum and as such of the spectral
information. Using the inverse formula one can in principle reconstruct the original
absorption spectrum from the saturated one. An uncertainty in this procedure is the
exact value of B and in addition this data treatment increases the noise.
Partial Fluorescence Yield
Recently it became possible to use fluorescence detectors with approximately 1 eV
resolution to tune to a particular fluorescence channel. This could be denoted as
21
X-ray spectroscopy part 1
partial fluorescence yield. The technique is also known as selective x-ray absorption
because one can select for example a particular valence and measure the x-ray
absorption spectrum of that valence only. Other possibilities are the selectivity to the
spin orientation and the nature of the neighbouring atoms. Partial fluorescence yield
effectively removes the lifetime broadening of the intermediate state. This effect can
be used to measure x-ray absorption spectra with unprecedented spectral resolution.
Electron yield
With the total electron yield method one detects all electrons that emerge from the
sample surface, independent of their energy. One can use a number of detection
devices: (i) one can detect the current that flows to the sample with a pico-ampere
meter, and (ii) one can use a channeltron to amplify the emitted electrons to a
detectable signal. In addition it can be shown that if the measurements are carried
out in a gaseous atmosphere with a pressure of approximately 10 mbar, the detected
ionised gas molecules yield a signal that is under certain conditions proportional to
the absorption cross section. The interaction of electrons with solids is much larger
than the interaction of x-rays which implies that the electrons that escape from the
surface must originate close to the surface.
The probing depth of total electron yield lies in the range of approximately 3 to 10
nm, depending also on the material studied. A quantitative study on the oxygen K
edge of Ta2O5 determined an electron escape depth of 4 nm. Studies of rare earth
overlayers on Ni revealed very short escape depths of the order of 1 nm. The
extremely large x-ray absorption cross sections of the rare earth M4,5 edges (related
to 3d to 4f transitions) makes that the x-ray penetration depth falls in the same
range as the electron escape depth. This implies that both the electron escape depth
e and the (angular dependent) x-ray penetration depth p()sin affect the
measured x-ray absorption cross section (Ie), according to:
I e ( ) 
e
 p ( ) sin 
Angular dependent measurements on Dy/Ni(110) revealed an x-ray penetration
depth of only 12 nm at the M5 resonance of Dy. In most other cases, the x-ray
penetration depth will be significantly longer in line with the x-ray transmission plots
as given above.
Instead of just counting the escaped electrons, one can detect their respective
energies and as such measure the partial electron yield. The figure shows a
schematic spectrum of the electron decay energies of the nickel K edge.
22
X-ray spectroscopy part 1
Figure: Schematic electron
emission spectrum for a Ni K
edge experiment. The peak at
low energy goes up to 80.0 on
the scale of the figure. The
primary Auger channel is the
1s2p2p channel at 6.5 keV.
The main secondary channel is
the 2p3d3d Auger channel.
This figure shows the primary 1s2p2p Auger electrons with kinetic energies of
approximately 6500 eV. These primary electrons can loose part of their energy in
inelastic scattering phenomena after which they have energies anywhere between
6500 eV and zero. In addition secondary Auger channels are visible as separate
peaks, for example the 2p3d3d channel at approximately 850 eV. One can view this
process as follows: The 1s x-ray absorption process has created a 1s core hole at
~8200 eV. The main Auger channel is the 1s2p2p channel that leaves two core holes
in the 2p level, each with a binding energy of 850 eV. This implies that the kinetic
energy of the Auger electron is 8200 - 850 - 850 = 6500 eV (neglecting correlation
effects). However the system is now stuck with two core holes in the 2p core level.
Both these holes can decay again via 'cascade' Auger processes, for example 2p3d3d.
The 2p3d3d Auger electron has an energy of 850 - 0 -0 = 850 eV and the system is
left with only valence holes that can decay usually via phonons and recombination
processes
Each such secondary peak brings with it its inelastic tail. The majority of the emitted
electrons from a sample are true secondary electrons with emission energies of less
than 100 eV. These electrons have lost most of their energy (due to inelastic
scattering processes) in their route to the surface of the sample. The figure shows an
actual Auger decay spectrum for NiO. This figure shows that apart from the main
channels such as 2p3d3d, essentially all Auger decay channels do contribute. The
figure also includes the relatively weak 2p3s3s, 2p3s3p and 2p3p3p Auger decay
channels.
Figure: Auger decay spectrum for NiO
excited with x-rays of 853.4 eV. The
structures visible are respectively
related to the 2p3p3p, 2p3s3p and
2p3s3s Auger decay channels. The
dark areas at the bottom relate to the
23
X-ray spectroscopy part 1
multiplet calculations of these channels (see chapter 3).
If one detects an Auger decay channel with high resolution one can perform similar
experiments as with high-resolution fluorescence detection and do selective x-ray
absorption experiments.
Ion Yield
If the absorption process takes place in the bulk and the core hole decays via an
Auger process a positively charged ion is formed, but due to further decay, inelastic
electron scattering and screening processes the original charge neutral situation will
be restored after some time, typically in the femtosecond range. If however the
absorption process takes place at the surface, the possibility exists that the atom
which absorbs the x-ray is ionised by Auger decay and escapes from the surface
before relaxation processes can bring it back to a bound state. If the escaping ions
are analysed as a function of x-ray energy the signal is in general proportional to the
absorption cross section. Because only atoms from the top-layer are able to escape,
ion yield is extremely surface sensitive. The mere possibility of obtaining a
measurable signal from ion yield means that the surface is irreversibly distorted, but
as there are of the order of 1015 surface atoms per cm2 this does not necessarily
mean that a statistically relevant proportion of the surface has to be affected for a
measurable signal of, say, 107 ions.
24