Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

X-Ray Fluorescence Spectroscopy

Related terms:

Microscopy, Heavy Metal, Analytical Method, Elemental Analysis, Energy-Disper-


sive X-Ray Spectroscopy, X-Ray Photoelectron Spectroscopy, Auger Electron Spec-
troscopy, Nanoparticles, Impurity

View all Topics

Sputter Processing
Andrew H. Simon, in Handbook of Thin Film Deposition (Fourth Edition), 2018

X-ray fluorescence
X-ray fluorescence (XRF) is a well-established technique for materials analysis that
has been adapted for in-line semiconductor industry use [77]. The sample under
analysis is illuminated by X-rays or gamma rays, which results in the excitation of
core-level electrons to excited states. The radiative decay of these electrons from
the excited states back to their respective ground states results in the emission of
fluorescent or “secondary” X-rays that are characteristic of the energy levels of each
atomic species and thus serve as a spectroscopic fingerprint for each element present
in the sample. Product-wafer spot sizes can be as small as ~100 nm or less.

Since the XRF signal intensity for each atomic species correlates directly to the
number of atoms present, the XRF signal can be used as a direct measurement of
the thickness of metals and alloys. The primary strength of XRF as a measurement
technique lies in its ability to assess thin-film thicknesses and alloy concentrations
independent of any numerical modeling techniques. The high-frequency trans-
parency of metals means that samples of several microns thickness can be measured.
At the opposite extreme, XRF can, in theory, be used to measure arbitrarily thin layers
of <10 nm thickness. The main challenge in measuring very thin layers is the long
acquisition time needed for the XRF detector to acquire a statistically significant
number of fluorescence counts and the care needed to deconvolve any spectral
overlap coming from substrates, underlayers, etc.
There are some limitations to XRF. For rigorous quantitative results, the XRF signal
should be calibrated against known thickness standards. XRF also is less useful for
measurement of elements with low atomic numbers, typically Z<11, due to weak
fluorescence from these species. The X-ray transparency of the films in question
means that XRF generally cannot be used for depth profiling: it can measure the
thicknesses of stacked films but typically cannot tell which one is on top of the
other. Finally, care needs to be taken in selecting which spectral lines to sample since
strong spectral signals from substrates or underlayers can potentially overlap with
the thin-film signals being measured, leading to error in the estimation of the signal
strength of the latter.

> Read full chapter

Ultratrace Impurity Analysis of Wafer


Surfaces
Steven M. Hues, Luke Lovejoy, in Handbook of Silicon Wafer Cleaning Technology
(Third Edition), 2018

13.2.3 VPD/TXRF
TXRF analysis alone, without coupling to VPD, is a surface-sensitive technique capa-
ble of detecting medium and high-Z elements (sulfur to uranium) at very low levels;
detection limits generally are in the 1010 atoms/cm2 range using a tungsten rotating
anode X-ray source. More detail on the TXRF technique is provided in Section 13.2.2.
Although this technique alone has proven successful in the determination of surface
metallic contamination on the wafer surface, the combination of VPD with TXRF
has been shown to improve the detection limit of TXRF by preconcentration of the
surface contaminants using VPD. The detection limit improvement can be estimated
from the ratio of the total wafer area to the analysis area; improvement of the
detection limit of two orders of magnitude may be achieved for a 200-mm wafer.
There are some disadvantages in using VPD with TXRF. VPD alone is a destructive
technique. TXRF without VPD preconcentration is a nondestructive technique where
the wafers are exposed only to soft X-rays that do not disturb the wafer surface,
whereas VPD preconcentration prior to TXRF completely decomposes the surface
oxide and concentrates the contaminants into one small area. Another disadvantage
of using VPD prior to TXRF is that the VPD may add contamination to the wafer
and thus produce data that are not representative of the actual contamination.
Further, using VPD assumes the contamination on the wafer surface is distributed
homogeneously on the wafer.
When performing TXRF analysis, the chemical and physical state of the wafer surface
impacts the intensity of the X-rays fluoresced; wafer roughness and the physical
structural of the metallic contamination on the wafer surface are the main factors
contributing to the intensity of the fluorescent X-rays. A classic technique used in
TXRF analysis is termed a rocking scan, which is obtained by measuring the fluores-
cent X-ray intensity as a function of primary X-ray incident angle. This measurement
can elucidate whether the contamination on the wafer surface is bulk, films, or
particles. With bulk contamination, the signal intensity increases constantly with
the sharpest increases at angles greater than the angle needed for total reflection
of the primary X-ray off the surface of the wafer. With a contaminating film, the
signal intensity increases as the angle is increased but finds a sharp drop in signal
intensity at angles higher than the angle needed for total reflection of the primary
X-ray off the surface. Finally, particulate contamination exhibits the same behavior as
uniform film contamination, except it shows higher intensity at the reflection angle
and continues to fall after the reflection angle is passed. When a wafer is prepared
by the VPD technique and the sample droplet is dried on the wafer surface, TXRF
analysis is affected by the physical state of the dried droplet on the wafer surface
and the glancing angle of the incident X-rays. Experiments [27] have shown that the
dried SCS residue is indeed particlelike, thus lower glancing-incident X-ray angles
can be utilized to produce higher intensity fluorescence from the contaminants on
the wafer surface.

> Read full chapter

Physicochemical characterization of
nanomaterials: polymorph, composi-
tion, wettability, and thermal stability
Ahmed Barhoum, ... Guy Van Assche, in Emerging Applications of Nanoparticles and
Architecture Nanostructures, 2018

3 Elemental composition
X-ray fluorescence spectrometry (XRF) is a relatively effective quantitative technique
to determine the elemental composition of any material, with trace element abun-
dances at the ppm level [21]. XRF is often used to verify the XRD data and vice versa.
The sample is excited by a primary high-energy X-ray. Characteristic X-rays of energy
lower than the incident X-rays are emitted from the sample, so-called secondary
X-rays or X-ray fluorescence (Fig. 9.7). The fluorescent X-rays are used to determine
the abundances of elements, while the intensity of these rays is proportional to the
abundance of the elements present in the sample [21]. X-rays can be characterized by
two main types of detectors. The first is wavelength dispersive detection (WDXRF),
in which the characteristic fluorescence X-rays diffracted from an element of interest
are selectively measured. The second is energy dispersive detection (EDXRF) which
detects a broad range of elements simultaneously [22]. The detectors can be adapted
to measure a broad range of elements simultaneously (i.e., EDXRF) or only the X-rays
produced by fluorescence of atoms of specific elements (i.e., WDXRF) [23]. However,
only WDXRF provides high elemental specificity and high sensitivity.

Figure 9.7. Possible interactions between X-rays and a sample for various X-Ray
techniques.(From H. Chen, M.M. Rogalski, J.N. Anker, Advances in functional X-ray
imaging techniques and contrast agents, Phys. Chem. Chem. Phys., 14(39) (2012)
13469–13486 [57]).

> Read full chapter

Defect and failure analysis techniques


for encapsulated microelectronics
Haleh Ardebili, ... Michael G. Pecht, in Encapsulation Technologies for Electronic
Applications (Second Edition), 2019

8.6 X-ray fluorescence spectroscopy


XRF spectroscopy is a fast, accurate, and nondestructive technique used to identify
and detect material composition. It is compatible with solid, liquid, and powdered
samples, and requires no or minimal sample preparation. There is no need to place
the sample in vacuum chamber as with energy-dispersive spectroscopy systems. XRF
spectrometer types can vary from light handheld devices to tabletop machines.

A movable X-Y stage and variable Z-axis source allows great flexibility with sample
size: from a single surface-mount chip to a handheld portable electronic device. A
video camera in sync with the emitter allows continuous monitoring of the position
on the sample, making it possible to locate an area of interest prior to analysis.

XRF spectrometers can operate in two main modes: material analysis mode and
thickness measurement mode [1]. The material analysis mode of the XRF spec-
trometer is capable of wide range of material analysis from aluminum (Z = 13) to
uranium (Z = 92). Material composition ranging from 0.1% to 100% can be detected
accurately. A typical XRF spectrometer system may consist of four collimators with
programmable, motorized controls. Analysis of small regions of interest on the
specimen can be achieved with the minimum XRF collimator size of 100 μm.

The XRF spectrometer thickness measurement mode provides the ability to measure
the depth of known layers of material. Each layer can be either a single element or
an alloy. The depth of penetration in XRF spectrometry varies based on the materials
used, but over 50 μm is possible.

> Read full chapter

Powder Characterization and Testing


Oleg D. Neikov, Nikolay A. Yefimov, in Handbook of Non-Ferrous Metal Powders
(Second Edition), 2019

X-Ray Fluorescence Spectrometry


X-ray fluorescence spectroscopy (XRF) is based on the excitation of atoms of the
material under study by an X-ray beam, resulting in the secondary fluorescent
emission. X-ray fluorescence intensity directly depends on the concentration of
every element in the specimens. Spectra analysis allows determining the specimen's
elemental composition.

Information about the source and energy of the fluorescent signal is contained in
Table 1.6. The X-ray tube with Rh-anode is a widespread device in this technique
field. However, in certain cases, other materials may be preferable (e.g., Mo, Cr,
Au, etc.). Generally, XRF is used to determine the concentrations of elements from
beryllium (Z = 4) to uranium (Z = 92) in the range from ppm fractions to 100% with
accuracy of 10− 7–10− 6 g. Usually, the sample weight is 0.1 g.

Two types of X-ray spectrometers—the energy-dispersive spectrometer (EDS) and its


modification wavelength-dispersive spectrometer (WDS)—are used by XRF analysis
for decomposition of the fluorescent radiation in the spectra. These spectrometers
are equipped with the scanning electron beam instruments (Table 1.6) and EPMA
apparatus. Additional information about these detectors is contained above, in
“Electron Probe X-Ray Microanalysis” section.

In WDS spectrometers, the selection of the different wavelengths corresponding to


the investigated elements is used for the spectrum decomposition. For this purpose,
the crystal monochromators (analyzers) with the crystal planes parallel to the surface
and with the interlayer distance d are used. When the radiation with wavelength
is directed on the crystal at the angle , the diffraction occurs only if the distances
passed by the photons reflected from the adjacent crystal planes differ on an integer
number n of wavelength . With the change of the angle by rotation of the crystal
relative to the radiation current, the diffraction will be occur sequentially for different
wavelengths, according to Wulf-Bragg’s law:

The angular position ( ) of the crystal is determined depending on the wavelength,


which is necessary to separate from the spectrum to analyze the desired element.

Single crystals such as germanium (Ge111) and lithium fluoride (LiF200/220/440)


are ideal analyzers (crystal monochromators) for the radiation of many elements.
Multilayer synthetic coatings are used to increase the sensitivity by analysis of the
light elements.

In WDS, for relatively long wavelengths (light elements), the gas-filled proportional
detectors (flowing or sealed) are used. Their effect is based on the gas ionization by
radiation and the measurement of electrical impulses passed through the ionized
gas. For shorter wavelengths (heavy elements), the scintillation detectors are used.
Their operations are based on the registration by photo-detector the light flashes
arising when X-rays are collided with scintillator (NaI/Tl) and measurement of
electrical impulses formed by a photomultiplier.

In sequential action WDS (with a scanning channel), the measurements are carried
out by successive selection of each of the characteristic X-ray lines of any number of
elements by means of the mobile crystal monochromator and precision goniometer.
In simultaneous action WDS (with fixed channels), the intensity of the characteristic
radiation of elements is measured simultaneously by using multiple turned fixed
channels located around the sample. In fact, each channel is a separate spectrometer
with a crystal monochromator and a detector turned to accept a certain wavelength
of one element.

In EDS the whole energy range of secondary (characteristic) radiation from the
specimen is registered simultaneously. The resulting spectrum corresponds to a
dependence of the intensity on the element's radiation energy. This is achieved by the
element-wise decomposition of the energy current of characteristic X-ray emissions
from the excited specimen.
In EDS spectrometers, the semiconductor solid-state detectors (Si or Ge) based on
the inner ionization of the semiconductor are used. During the analysis by the Peltier
effect or by liquid nitrogen, the cooling of the detector is mandatory. Voltage up to
several kV is applied to the semiconductor crystal, and it provides the gathering of
all the charges formed in the detector's inner volume. The electric pulse arises, and
it is further amplified and recorded by the counting electronics.

The spectrum processing by the mathematical processing programs enables provid-


ing the quantitative and qualitative (semiquantitative) analysis. In modern XRF spec-
trometers, the amplifiers and pulse analyzers allow us to accomplish the elemental
composition measurement of a specimen in less than 2 s with satisfactory statistical
error.

> Read full chapter

Application of ionic liquids for


rare-earth recovery from waste electric
materials
Fukiko Kubota, Masahiro Goto, in Waste Electrical and Electronic Equipment Recy-
cling, 2018

12.5.1 Leaching from phosphor powder scraps


Based on the XRF analysis results, waste phosphor powders of fluorescent lamps
contain rare-earth elements such as Y, Eu, La, Ce, Pr, and Tb at ~30% of the total
weight, and other elements are mainly Ca and Sr, and a variety of metals such as Al,
Ti, V, Fe, and Zn as shown in Fig. 12.3B. Leachates from the waste phosphor powders
were prepared by a two-stage leaching method [8]. In the rare-earth metals that
are contained in the phosphor powder, Y and Eu exist as the oxide Ln2O3, which is
readily soluble in acidic solution, whereas the other rare-earth metals are contained
as complex crystals of their phosphates, which are not easy to dissolve, unlike Y and
Eu [5,8]. The leaching efficiency is affected by several factors such as the type of
acid used, leaching time, and temperature. In the first stage, the leaching efficiency
reached a maximum with mixing for 6 h at 100°C, in the case using 5 M sulfuric acid
as a leaching solution, as shown in Fig. 12.13. All Y and more than 95% of Eu were
leached. However, all the light rare earths La, Ce, and Pr and the middle rare earth
Tb remained in the residue without being dissolved.
Figure 12.13. Acid-leaching method and typical leaching behavior. Phosphor pow-
der: 1 g, leaching solution: 20 mL, 5 M H2SO4, leaching time 6 h.

In the second stage, leaching was repeated from the residue on the first leaching
by using sulfuric or nitric acid in a similar way as in the first stage of leaching.
Sulfuric acid (5 M) could leach only ~50% La and Ce. Strong sulfuric acid (18 M)
was required to leach more than 90% of the metal ions as shown in Fig. 12.14A.
Removal of more than 91% La, 86% Ce, and 98% Tb was obtained by using 5 M
nitric acid. Only ~25% Pr was dissolved in 18 M sulfuric and 5 M nitric acids. The
use of a low acid concentration for leaching is preferable because a high alkaline
dosage is not required for controlling the leachate pH, which is subjected to the
extraction operation. Thus, in the two-stage leaching processes, two leachates are
obtained. One leachate is 5 M sulfuric acid that contains Y and Eu and the other is
5 M nitric acid solution that contains La, Ce, Pr, and Tb. In each leaching process,
metal impurities are also dissolved in the leachate. The leaching efficiency through
the two-stage leaching is shown in Fig. 12.14B.
Figure 12.14. Leaching efficiency from fluorescent phosphor powder. (A) Effect of
acid and concentration on second-stage leaching. (B) Total leaching degree of first
stage by 5 M H2SO4 and second stage by 5 M HNO3. Open part by first stage and
closed by second stage.

For the phosphor powder in a CRT tube, the rare-earth metals, Y and Eu, could be
leached quantitatively with 3–5 M nitric acid at room temperature, where a large
amount of Zn was dissolved simultaneously.

In Section 12.5.2, we show the application of novel IL extraction systems that we have
developed, for the recovery of rare-earth metals from real leachates.

> Read full chapter

Thin film fabrication using nanoscale


flat substrates
Takashi Nishida, in Nanoscale Ferroelectric-Multiferroic Materials for Energy Har-
vesting Applications, 2019
6.4 Evaluation of PbTiO3 thin films by X-ray fluorescence spec-
troscopy and X-ray diffraction
PTO thin films deposited on the atomically flat substrates shown in Fig. 6.2 were
evaluated by XRF and XRD. On the basis of the X-ray fluorescence spectroscopy (XRF)
data, the Pb:Ti ratio in each of the films was found to be 1.2:1, indicating that they
were Pb-rich by 20 at.%. The thickness of each film was 100 nm. The XRD patterns
for the films are shown Fig. 6.3. These films were epitaxially oriented, and the results
of a scan are presented in Fig. 6.4.

Fig. 6.3. X-ray diffraction patterns generated by PTO thin films on atomically-flat
substrates prepared at various temperatures [3].

Fig. 6.4. The results of a scan of a PbTiO3 thin film on an atomically flat substrate
together with data for -Al2O3 (annealing temperature 1000°C) for comparison
purposes [3].

An epitaxial PTO thin film oriented solely in the (111) direction was deposited
only on the substrate annealed at 1000°C. In contrast, the films on the other two
substrates generated either a PbO (111) or pyrochlore peak in addition to the PTO
(111) peak. These other phases could possibly have resulted from the substrate
surface conditions. For this reason, it appears that single steps and flat terraces might
contribute to the deposition of an epitaxial PTO thin film.
In addition, because the number of peaks in the scan for the PTO was twice that
for an -Al2O3 reference (Fig. 6.4), the epitaxial PTO film evidently had two different
in-plane orientations.

> Read full chapter

X-ray Microanalysis and Electron Ener-


gy Loss Spectroscopy (EELS)
Gurram Giridhar, ... Gudimalla AppaRao, in Microscopy Methods in Nanomaterials
Characterization, 2017

8.1.2.2.1 Wavelength Dispersive X-ray


Wavelength dispersive X-ray (WDX) is based on measurement of the wavelengths
and intensities of X-ray spectral lines, which are emitted by secondary excitation,
Wavelength-dispersive X-ray secondary-emission spectrometry, or X-ray fluores-
cence spectrometry (XRFS), is useful for qualitative and quantitative analysis of
chemical elements. Depending on the concentration, the emission of secondary
spectral lines having characteristics of elements in a specimen takes place by X-ray
excitation. The spectral lines are dispersed specifically by crystal diffraction prior to
detection.

In wavelength-dispersive spectrometers, several X-ray lines emitted by the specimen


are dispersed specifically by crystal diffraction, prior to detection, on the basis of
their wavelengths, and the detector receives only one wavelength at a time. This
method is very much applicable to all chemical elements down to atomic number
4 (beryllium), but standard commercial instruments are limited to atomic number
9 (fluorine). The sample might be in any form, including solid, briquette, fusion
product, powder, film, liquid, slurry, and fabricated forms such as rod and wire, and
parts of any form and size are acceptable for the analysis.

> Read full chapter

Assessment methods of bone-to-bio-


materials regeneration
Vincent M.J.I. Cuijpers, ... John A. Jansen, in Dental Implants and Bone Grafts, 2020
11.4.3 X-ray fluorescence contrast
Although the term “fluorescence” is well known in light microscopy, it can also
be applied to X-rays. The term fluorescence then describes the phenomenon in
which the absorption of X-radiation of a specific energy results in the re-emission
of X-radiation of a different energy.

Thus X-ray fluorescence (XRF) is the emission of characteristic secondary X-rays


from a material that has been excited by being bombarded with high-energy X-rays.
In this way, obtained X-ray fluorescence can be used for functional imaging or to
provide molecular information. As the secondary X-rays also allow for nondestructive
chemical mapping, this permits the co-registration of 3D micro-morphology and
3D chemical composition. XRF has been used to study moderately X-ray-transparent
(soft) tissues [69], but the use of XRF to study bone tissue has also been initiated.
In a study on osteoporosis in a rat model, the strontium (Sr) distribution received
from a daily dose of Sr-containing drugs was evaluated and co-registered with a
micro-CT absorption contrast image (Fig. 11.6). Despite such new developments
in X-ray imaging, spatial resolution was very much restricted compared to more
common imaging modalities such as optical imaging.

Fig. 11.6. Micro-X-ray fluorescence (micro-XRF) imaging in tissue engineering appli-


cations. 3D volume rendered micro-CT image (A) and a fused micro-CT–micro-XRF
image (B) of a vertebra of an osteoporotic rat which received a daily dose of stron-
tium-containing drugs. The front cover quarter is virtually removed to show the in-
ternal microstructure and concentration distribution. The strontium concentration
is displayed color-coded in a blue-red scale whereby the highest uptake is indicated
in red.Image courtesy of Bruker MicroCT, Belgium.

> Read full chapter


Overview of Surface Cleanliness Mea-
surement Methods
In Developments in Surface Contamination and Cleaning, Volume 12, 2019

The applications of the characterization methods discussed in this volume are broad-
ly summarized for the key types of contaminants. Sampling of the contaminants for
analysis is performed by the methods discussed in Chapter 3.

• Particles•Counting and sizing: visual examination, microscopy, optical particle


counters, and size classification techniques•Elemental analysis: electron mi-
croscopy with analytical detectors•Bulk analysis by spectroscopic techniques,
mass spectrometry, and X-ray diffraction and fluorescence•Imaging: optical
and electron microscopies, surface analytical techniques
• Films•Thickness: ellipsometry, X-ray techniques, surface analytical techniques-
•Composition: carbon coulometry, surface analytical techniques, mass spec-
trometry•Surface deposition: surface acoustic wave, quartz crystal microbal-
ance, gravimetric method, optically stimulated electron emission, ellipsome-
try, MESERAN
• Ionic•Composition: ion chromatography

• Metals•Distribution (mapping): surface potential difference, X-ray fluores-


cence, surface analytical techniques•Elemental analysis: electron microscopy
with analytical detectors, surface analytical techniques, neutron activation
analysis, deep-level transient spectroscopy, Hall effect•Bulk analysis: mass
spectrometry, spectroscopy, X-ray diffraction and fluorescence
• Organic•Distribution (mapping): spectroscopic techniques•Composition:
spectroscopic techniques, mass spectrometry, chromatography
• Biological contaminants•Counting and sizing: microscopy•Identification
[1296–1304]: visual examination with dyes (e.g. Coomassie®, Amido black
10B), ATP (adenosine triphosphate) bioluminescence, microscopy, spec-
troscopy with fluorescent dyes (e.g. ruthenium-based stain SYPRO Ruby,
silver stains, DAPI (4 , 6-diamidino-2-phenylindole dihydrochloride), dansyl
polymxin, FTIR and Raman spectroscopy, MALDI-TOF-MS, radioactive tracers

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like