Radioanalytical Chemistry Guide
Radioanalytical Chemistry Guide
Radioanalytical Chemistry
Library of Congress Control Number: 2006925171
C 2007 Springer Science+Business Media, LLC
All rights reserved. This work may not be translated or copied in whole or in part without the written
permission of the publisher (Springer Science+Business Media, LLC, 233 Spring Street, New York,
NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use
in connection with any form of information storage and retrieval, electronic adaptation, computer
software, or by similar or dissimilar methodology now known or hereafter developed is forbidden.
The use in this publication of trade names, trademarks, service marks, and similar terms, even if they
are not identified as such, is not to be taken as an expression of opinion as to whether or not they are
subject to proprietary rights.
9 8 7 6 5 4 3 2 1
springer.com
Contents
Acknowledgments..................................................................... vii
1 Introduction ........................................................................ 1
Bernd Kahn
v
vi Contents
Glossary................................................................................. 421
The following members of the Editorial Advisory Board (EAB) contributed enor-
mously to this textbook by suggesting format, reviewing successive drafts, and
some even writing chapters (as indicated in chapter headings). It was a great plea-
sure to work with all contributors, and they have my heartfelt thanks for their
participation.
r Moses Attrep, Jr., LANL-retired
r Darleane Hoffman, LBL
r Kenneth Inn, NIST
r John Keller, ORNL
r Harry Miley, PNNL
r Stan Morton, RESL-IDO-retired
r Glenn Murphy, UGA
r Dick Perkins, PNNL
r Charles Porter, EPA-retired
r Jake Sedlet, ANL
r John Wacker, PNNL
Isabel Fisenne and A. L. Boni were members of the EAB for several years and
then resigned.
My deepest sympathy goes to the family of Jake Sedlet, who died while this
work was in progress.
I appreciate the efforts of others who contributed by writing a chapter:
r Craig Aalseth, PNNL
r Gregory Eiden, PNNL
r Pam Greenlaw, EML
r Jeffrey Lahr, GT
r Scott Lehn, PNNL
r Keith McCroan, EPA
r Robert Rosson, GT
r Paul Schlumper, GT
r Linda Selvig, Boise ID
vii
viii Acknowledgments
r Liz Thompson, GT
r Arthur Wickman, GT
Each chapter is credited to the main author or authors, but many portions of
their original manuscripts were moved to other chapters for better text integration.
This effort was supported by NNSA–DOE under grant DE-FG07-01ID14224.
Dan Griggs and Stephen Chase were successive project officers. I am grateful to
them for their active involvement in the preparation of this text, and to Dale Perry
and Terry Creque of NNSA for participating in the EAB meetings.
I thank the Georgia Institute of Technology, my professional home when the
book was written, and my coworkers there. Liz Thompson was the editorial writer
who untiringly worked on every aspect of the text to prepare it for publication with
great insight and competence. Robert Rosson and Jeff Lahr are radiochemists who
were most helpful in working with me and writing specific chapters. Jean Gunter
was the skilled coordinator of all EAB meetings, who made attendance a pleasure.
I wish to express my deepest respect and thanks to the two radiochemists who
taught me, William S. Lyon at ORNL and Charles D. Coryell at MIT.
This book is dedicated with love to my wife, Gail, and my parents, Alice and
Eric Kahn.
1
Introduction
BERND KAHN
1.1. Background
Radioanalytical chemistry is devoted to analyzing samples for their radionu-
clide content. For this purpose, the strategies of identifying and purifying the
radioelements of interest by chemical methods, and of identifying and measur-
ing the disintegration rate (“activity”) of radionuclides by nuclear methods, are
combined. Radioanalytical chemistry can be considered to be a specialty in the
subdiscipline of nuclear and radiochemistry.
This textbook was written to teach radioanalytical chemistry in the classroom
and support its application in the laboratory. Its emphasis is on the practical as-
pects of the specialty, notably setting up the laboratory, training its staff, and
operating it reliably. The information presented herein, outlined in Section 1.4,
is the accumulated product of a century of nuclear chemistry and radiochemistry
practice.
Radioanalytical chemistry was first developed by Mme. M. Curie, with contribu-
tions by many other distinguished researchers, notably E. Rutherford and F. Soddy.
These pioneers performed chemical separations and radiation measurements on
terrestrial radioactive substances during the 20 years following 1897 and in the
process created the very concept of radionuclides. Their investigations defined
the three major radiation types, confirmed the emission of these radiations by the
nucleus and the associated atomic transformations, established the periodic table
between bismuth and uranium, and demonstrated the distinction between stable
and radioactive isotopes.
Thereafter, cosmic rays were observed and explained, and many cosmic-ray-
produced radionuclides were identified. The number of known and character-
ized radionuclides increased dramatically with the development and application
of nuclear-particle accelerators in the 1930s and nuclear-fission reactors in the
1940s.
Since Mme. Curie’s time, applications of radioanalytical chemistry have pro-
liferated. Modern practitioners of nuclear and radiochemistry have applied chem-
ical and nuclear procedures to elucidate nuclear properties and reactions, used
1
2 Bernd Kahn
r Operate the radioanalytical chemistry laboratory safely and in accord with reg-
ulations, with assured quality and cost-effectiveness;
r Plan and perform identification and measurement of radionuclides through a
combination of chemical separation and radiation detection methods;
r Select sample size, sample processing, radiation detection instruments with pe-
ripherals, and measurement period to match detection sensitivity specifications;
r Identify radionuclides by their chemical behavior and radiation, and resolve
ambiguities by expanding or revising separations and measurements;
r Calculate radionuclide concentration values and their uncertainty with consid-
eration of the radionuclide decay scheme, sample processing losses, radiation
detection efficiency, and interfering radiation.
2.1. Introduction
As noted in Chapter 1, a background in analytical chemistry is essential to the
practice of radioanalytical chemistry. The parallels between the two fields will
become evident as analytical and radioanalytical chemistry principles are covered
in Chapters 3 and 4, and culminate in the discussion of applied radioanalytical
chemistry in Chapter 6. However, the theoretical underpinnings of the two disci-
plines are markedly different.
The primary distinction between analytical chemistry and radioanalytical chem-
istry is the nature of the transformations being examined. The analytical chemist
is concerned with chemical transformations, brought on by the interaction of
an atom’s valence electrons with its physical environment. The radioanalytical
chemist, on the other hand, is primarily interested in the nuclear transformation
of a given atom. For practical purposes, the physical environment of the atom has
no effect on the nuclear event. Consequently, many of the instrumental methods of
detection most widely utilized in the normal course of analytical characterization
have little use in the radioanalytical laboratory.
Instead, the radioanalytical chemist focuses on the detection of radiation, the
by-product of a nuclear transformation. The analyst must understand the types
of radiation that may be encountered and the way that each interacts with matter.
With this knowledge, the analyst can adapt the method of detection to the particular
radionuclide of interest. The goal of this chapter is to provide a brief review of
nuclear chemistry as it relates to the principles of radiation detection. Next, an
overview of the operating principles of commonly used detectors is provided as a
basis for understanding the material presented in Chapter 8.
1
344 Kimberly Lane, Los Alamos NM 87544.
2
Los Alamos National Laboratory (retired), Los Alamos, NM 87545.
7
8 Moses Attrep, Jr.
TABLE 2.1. The three common types of radiation emitted from unstable nuclei
Types of radiation Symbol Description
Alpha α A helium-4 nucleus composed of two protons and two neutrons;
mass is approximately 4 Da; charge +2; no spin
Beta β An electron; mass ∼1/1822 Da; charge −1 or +1; spin 1/2
Gamma γ Electromagnetic radiation; no mass; no charge
of one or more forms of radiation. The three basic types of radiation are alpha (α),
beta (β) and gamma (γ ) radiation (see Table 2.1). Three primary modes of nuclear
decay are alpha-particle emission, beta-particle emission, and gamma-ray emis-
sion. Above atomic number 90, spontaneous fission (SF), a natural decay mode in
which the nucleus spontaneously divides into two large fragments, becomes an in-
creasingly important mode of decay. During all of these processes, mass, charge,
and energy must be conserved. Thus, the sum of the mass numbers A and the
atomic numbers Z of the products (or daughters) must equal those of the initial ra-
dionuclide (the parent). Additionally, the mass of the parent must equal the masses
of the daughter and emitted particles plus mass equivalents of the kinetic energy
of the products. The known radioactive nuclides and their radiation are given in
tables and charts of nuclides together with the stable isotopes (Firestone 1996).
The specific mode of decay listed in Table 2.1 depends upon the nature of
the parent’s instability relative to the lower energy states to which it may decay,
and functions to transform the radioactive nucleus into a more stable nucleus.
Alpha emission (α) occurs in the neutron-deficient lanthanides and is the major
mode of decay in many heavy radionuclides of the actinides and transactinides.
Beta radiation, which can take the form of either negatron (β − ) or positron (β + )
emission, results from an imbalance in the neutron-to-proton (N /Z ) ratio of the
parent and occurs in the isotopes of elements throughout the periodic table. Gamma
radiation (γ ) is the result of an excited nucleus de-exciting to lower energy levels.
A fourth mode of decay is SF, where the parent nucleus fissions into two large,
neutron-rich fission fragments (FF). The FF may be stable or may de-excite by
neutron emission to lower mass nuclei and by β − emission to higher atomic number
nuclides until further decay is no longer energetically possible.
Each of these decay process is explained in more detail in the following sections.
All processes described below are exoergic; i.e., they give off energy.
A
ZX → A−4
Z −2 Y + 42 α (2.1)
2. Radiation Detection Principles 9
Because the momentum and energy evolved in the decay must be conserved,
they are distributed between the product nucleus (sometimes called the recoil
nucleus) and the emittedg alpha particle. The unit of energy commonly used in
describing nuclear decay and radiation is the electron volt and its multiples. One
electronvolt equals 1.6 × 10−19 J, which is numerically equal to the electron charge
e in coulombs.
Given the energy difference Q between isotopes X and Y , the kinetic energy of
the alpha particle is QY/(Y + α) (where Y and α are masses) and the kinetic energy
of the recoiling isotope Y is Qα/(Y + α). The lighter alpha particle is quite ener-
getic between 2.5 and 9 MeV, although more commonly between 4 and 6 MeV. In
contrast, the recoil energy of the heavier product nucleus typically is about 0.1 MeV.
A radionuclide may emit several alpha-particle groups of different energies and
intensities, but every alpha particle in a given group is of the same energy. As shown
in the example of 241 Am in Section 9.3.7, the most energetic alpha-particle group
decays to the ground state. Less energetic groups decay to states slightly more
energetic than the ground state and then decay promptly by conversion electron
(CE) and gamma-ray emission (Section 2.2.3) to the ground state.
FIGURE 2.1. Mass–yield curves for some fission products, including 235 U and 239 Pu. The
curve labels refer to the compound nuclei, as in 235 U + 1 n = 236 U. (From Vertes et al. 2003,
p. 222.)
FIGURE 2.2. Radioactive decay. The ordinate is in count rate and the abscissa is the time in
hours. (From Vertes et al. 2003, p. 263.)
as displayed graphically in Fig. 2.2. Note that the two straight lines pertain to two
radionuclides that are measured in the same sample. The curves may be plotted
in terms of the net count rate, RG – RB , shown in Eq. (8.1) as the measured count
rate from which the background count rate has been subtracted, if the counting
efficiency—the number of observed counts per disintegration—remains the same
for all measurements. The curved line in Fig. 2.2 is the total net count rate, the
sum of the two individual net count rates.
Measurements of the slope of the line in Fig. 2.2 or the disintegration rate at
two separate times can be used to calculate the half-life t1/2 , i.e., the time period
during which the radionuclide decays to one-half of a previous value. The half-life
is ln 2/λ, i.e., t1/2 = 0.693/λ.
The disintegration rate can be determined from the measured values of radionu-
clide mass m in grams and the half-life t1/2 in seconds. Equation (2.7) relates the
decay rate to the mass in terms of Avogadro’s number Av of 6.02 × 1023 atoms
per mol and the isotope mass number A in g/mol:
dN 0.693mAv
= (2.7)
dt At1/2
The mass at radionuclide disintegration rates commonly encountered in the lab-
oratory is extremely small: at a typical disintegration rate of 1 disintegration s−1
or Bq, the mass may be about 1 × 10−15 g. Weighable amounts of radionuclides
occur for radionuclides that have half-lives of billions of years. For such long-
lived radionuclides, it has been commonplace to calculate the disintegration rate
from the known half-life and measured mass, or the half-life from the measured
14 Moses Attrep, Jr.
mass and disintegration rate. Recently, masses have been measured by mass spec-
trometer for radionuclides with half-lives of as short as few thousand years (see
Chapter 17).
In some instances, a radionuclide decays not to a stable nuclide but to a second
radionuclide (or even an entire chain of radionuclides) before a stable nuclide is
finally reached. In that case, when the second radionuclide is separated from the
first, it immediately begins to grow into the first radionuclide while decaying in the
separated portion. For two successive radionuclides, the disintegration rate of the
daughter (subscript 2) is related to the disintegration rate at the time of separation,
t = 0, of the parent (subscript 1) by
dN dN λ2 −λ1 t −λ2 t dN
= (e −e )+ (e−λ2 t )
dt 2 dt 1(t=0) (λ2 − λ1 ) dt 2 (t=0)
(2.8)
The equation indicates that for a daughter that has a much larger value of λ than
its parent, i.e., a short-lived daughter of a long-lived parent, the second part of
the first term on the right approaches 1 and the third part approaches (1 − e−λ2 t ).
This is called secular equilibrium; i.e., the daughter reaches the same disintegra-
tion rate as the parent (see Fig. 9.10). The second term on the right describes
the decay of the separated daughter. Two other distinctive cases can occur (see
Section 9.3.1). Transient equilibrium refers to the value of λ2 that is only some-
what larger than that of λ1 , so that the daughter disintegration rate after some
time exceeds at an almost constant ratio the disintegration rate of the parent (see
Fig. 9.11). In the case of no equilibrium, λ2 is less than λ1 ; in this scenario, the
daughter decays at a slower rate than the parent after initial ingrowth, as depicted
in Fig. 9.12.
Calculations for longer decay chains under some conditions can be simplified
by assuming that short-lived daughters and long-lived parents have the same dis-
integration rate. In some complex chains, the Bateman equation (in the same form
as Eq. (2.8), but with terms added to describe further decays) can be used to de-
termine the ingrowth and decay pattern of three or four successive radionuclides
(Evans 1955).
will occur. For a thin target irradiated by a well-defined particle beam, the number
of target atoms is expressed by the term nx, where n is the number of atoms of
target nuclei per cubic centimeters and x is the thickness of the target material in
centimeters. The beam intensity is I , the number of incoming particles per second.
The probability of the reaction is given by the cross section with symbol σ and
units of centimeter square per atom. The cross section is tabulated in units of barns
(b), where 1 b is 10−24 cm2 .
The rate of production, P, in s−1 of atoms by reactions such as that shown in
Eq. (2.9) is
P = nxIσ (2.10)
In the case of neutron activation in a nuclear reactor, the rate of formation of
atoms is given by
P = N Φσ (2.10a)
where Φ is the neutron flux expressed in neutrons s−1 cm−2 and N is the number
of atoms in the sample, i.e., the product mAv /A in Eq. (2.7).
At constant irradiation intensity, the accumulated number of product atoms is
the integral of P over the time of production. In a simple case, the result is Pt.
It is less than this value when the projectile flux applicable to this reaction is
depleted significantly by parallel reactions; the number of target atoms decreases
significantly as a result of the reactions; and the number of product atoms decreases
significantly because of radioactive decay or further nuclear interactions.
If the produced atom is radioactive, the rate of radionuclide production in terms
of the disintegration rate [shown in Eq. (2.4)] is Rλ. The disintegration rate of the
accumulated atoms, balancing the production and decay rates, is then
dN
= N σ (1 − e−λt ) (2.11)
dt
in units of disintegrations per second.
r Some spontaneously fissioning radionuclides, such as 252 Cf, produce 3–4 neu-
trons per fission with an energy spectrum of average energy of about 2–3 MeV.
The neutron generation rate depends on the amount of 252 Cf, which emits ∼(2–
3) × 1012 neutrons s−1 g−1 .
r Accelerators provide a variety of nuclear reactions for production of neutrons.
Cockcroft–Walton accelerators can generate 14.8 MeV neutrons by accelerating
deuterons (2 H) onto a tritium target to produce 108 –1011 neutrons s−1 . Cyclotrons
and linear accelerators can produce high-energy neutrons with a broad spectrum
of energies in spallation reactions that result from the bombardment of heavy
elements by charged particles.
r Nuclear reactors are the most common source of neutrons. Inside the reactor, a
sustained nuclear reaction of fissile material produces fast neutrons. When 235 U
is used as the reactor fuel, 2–3 MeV neutrons are produced, along with other
neutrons at other energy ranges. A neutron moderator slows the fast neutrons to
reduce their energies to the thermal level. This is required to continue the chain
reaction by further absorption of neutrons by surrounding atoms of 235 U. Other
neutron energy ranges are the epithermal between 0.1 and 1 eV and resonances
between 1 eV and 1 keV.
The sum of the atomic masses of the heavy fragment, the light fragment, some
additional free neutrons, and the released energy equals the mass of the fissioning
atom. The Q value for such neutron-induced fission reactions is in excess of
200 MeV per fission.
For a given radionuclide, some fission-fragment pairs are more common than
others. This is easily seen in the mass–yield curves of Fig. 2.1, which show the spec-
trum of atomic masses given off by the fission of several radionuclides. Atom fission
18 Moses Attrep, Jr.
yields range up to 6–7% for an individual fission product at the heavy and light
mass peaks. These peaks represent the mass fragments that are most common when
the parent nucleus splits. The particular shape of the spectrum depends on both the
identity of the fissioning nucleus and the energy of the projectile. Fission product
yields, in the form of tables that give the yield of each fragment nuclide at each
mass number (from approximately 66 to 171), have been compiled over the course
of several decades, and are available online at http://ie.lbl.gov/fission/endf349.pdf
(January 2006); these are also published in a series of reports by England and
Rider (1994).
Fission products are neutron-rich and thus are negatron emitters. At each mass,
a decay series may consist of as many as five radionuclides that decay, one into the
other, until the chain stops at a stable element. In a decay series, the fission yield
may begin at a smaller value for the initial short-lived radionuclides and increase for
subsequent radionuclides to the maximum shown in the figure. Among these fission
products are ones with half-lives from days to years that are readily measured for
monitoring nuclear reactors and nuclear weapon tests.
The fission process releases the large amounts of energy used in nuclear reactors
and weapons and also several neutrons. The number of these neutrons depends on
both the process and the fissioning atom, as indicated in the 235 U decay equations
above. When the number of neutrons exceeds one per fission and the energy of the
neutrons is suitably moderated to induce further fission, a chain reaction is induced
that can be used for energy production.
10
7
Energy (MeV)
6
Aluminum
5
Air
4
Aluminum:
3 d = 2.70 g/cm3
Air:
2 d = 1.29 g/cm3
0
0 2 4 6 8 10 12
Range (mg/cm2)
FIGURE 2.3. Alpha-particle range in air and aluminum. (Data from ICRU 1993, pp. 192
and 213.)
range of 3.4 cm (CSDA range 4.37 mg cm−2 divided by the density of air, 1.29 mg
cm−3 ) in air. Equation (2.13) gives a value of 3.47. The Bethe equation for the
energy loss of charged particles is complex but indicates that the loss is related
to the electron density, i.e., the atom density times the atomic number Z . The
empirically derived Bragg–Kleeman rule suggests that the range in one material
relative to another is proportional to ρA0.5 (Evans 1955), where ρ is the density of
the attenuating material.
10.00
1.00
Energy (MeV)
Polystyrene
Aluminum
0.10
0.01
0.1 1 10 100 1000 10000
Range (mg/cm2)
FIGURE 2.4. Beta-particle range–energy curve (log/log) in aluminum and polystyrene. (Data
from ICRU 1984.)
range. A set of range–energy curves is shown in Fig. 2.4, with range in units of
mg cm−2 in aluminum and polystyrene. Range values in terms of this unit can
be applied to various substances within a restricted range of Z . An equation for
beta-particle range similar to the equation for alpha particles above is
n
= 412 E β (2.14)
2.5
2
Beta-distribution function
1.5
Y-90
Sr-89
S-35
1
0.5
0
0 0.2 0.4 0.6 0.8 1
(MeV)
FIGURE 2.5. Three typical beta-particle spectra. (Data from ICRU 1997, pp. 107–108.)
0.1
R/ R0
0.01
0.001
0 100 200 300 400 500
Absorber thickness (mg/cm2)
210
FIGURE 2.6. Beta-particle attenuation curve: Bi in aluminum absorbers. (Based on
Zumwalt 1950, p. 40.)
2. Radiation Detection Principles 23
Compton
electron
Compton
scattered
photon
The units of μ can be cm2 g−1 , cm2 atom−1 , or cm−1 ; respective units of x are g
cm−2 , atom cm−2 , or cm. The interaction fraction with the medium is I0 − If ,
Compton scattering occurs when a gamma ray interacts with an electron bound
weakly (i.e., an outer orbital electron) to an atom or molecule and transfers a
fraction of its energy to the electron. The resulting weaker gamma ray departs at
an angle, as illustrated in Fig. 2.7. The energy distribution between the electron and
the produced gamma ray depends on the angles at which the two move relative to
the initial gamma ray. Much or all of the energy transferred from the initial gamma
ray to the electron is deposited in the absorbing material in which the electron
slows down.
The electron energy is distributed between zero and a maximum energy that is
somewhat less than the energy of the incoming gamma ray, E γ . The maximum
electron energy is
2E γ /0.51
E max = E γ (2.16)
1 + 2E γ /0.51
For example, if the incoming gamma-ray energy is 1.0 MeV, the maximum
electron energy is 0.8 MeV. The energy of the scattered gamma ray is the en-
ergy of the initial gamma ray minus that of the ejected electron. That gamma ray
2. Radiation Detection Principles 25
TABLE 2.2. Photon interaction and energy absorption coefficients for common materialsa
Mass attenuation (cm2 g−1 ) Mass energy absorption (cm2 g−1 )
Energy
(keV) Water SiO2 CaCO3 Water Air Muscle Bone
50 0.208 0.282 0.485 0.0394 0.0384 0.0409 0.158
100 0.165 0.158 0.181 0.0252 0.0231 0.0252 0.0386
200 0.136 0.135 0.125 0.0300 0.0268 0.0297 0.0302
300 0.118 0.106 0.107 0.0320 0.0288 0.0317 0.0311
500 0.096 0.087 0.087 0.0330 0.0297 0.0327 0.0316
1000 0.0707 0.0636 0.0636 0.0311 0.0280 0.0308 0.0297
2000 0.0494 0.0447 0.0448 0.0260 0.0234 0.0257 0.0248
3000 0.0397 0.0363 0.0366 0.0227 0.0205 0.0225 0.0219
a Data from Shleien (1992).
Sodium iodide
Sodium iodide cm2/g
sr m0 r
ss r
ma r
kr
s
tr
sa r
r
sa
s
r
a
r
s
s
r
Mev
FIGURE 2.8. Gamma-ray interaction coefficients as function of energy in NaI. (From Evans
1955, p. 716.)
and is now calculated by Monte Carlo simulation. This “buildup factor” B (Shleien
1992) exceeds 1.0 and is used to multiply If in Eq. (2.15). The buildup factor
depends on gamma-ray energy, source-shield configuration, and shield thickness.
It usually is reported in terms of the dimensionless relaxation length μ0 x.
The second mechanism by which gamma radiation loses energy is the photo-
electric effect. In this interaction, all of the energy is transferred from the gamma
2. Radiation Detection Principles 27
Lead
Lead cm2/g
ss r m0 r
ma r
kr
tr
sr
ss
sa r
r
r
ss
r
Mev
FIGURE 2.9. Gamma-ray interaction coefficients as function of energy in Pb. (From Evans
1955, p. 717.)
ray to a strongly bound interior orbital electron. When the energy transmitted by
the gamma ray exceeds the electron binding energy E BE , the electron is ejected
from the atom or molecule with energy E e according to Eq. (2.17):
E e = E γ − E BE (2.17)
The photoelectric effect becomes more likely with lower E γ and as electrons
are more tightly bound in their shells (i.e., K > L > M). The missing electron
28 Moses Attrep, Jr.
2.5. Detectors
2.5.1. Gas-Filled Detectors
Gas-filled detector systems collect and record the electrons freed from gaseous
atoms and molecules by the interaction of radiation with these atoms and
molecules. The systems have been classified into the primary categories of ioniza-
tion chambers, proportional counters, and Geiger–Mueller (G-M) counters. The
detectors have a variety of designs, but essentially consist of a chamber, which
serves as the cathode, and a center wire, which serves as the anode. The electri-
cal field is characterized by chamber shape and radius, the wire radius, and the
applied voltage (Knoll 1989). The chamber is filled with a gas and a potential is
applied between the two electrodes. The configuration of the system is shown in
Fig. 2.10.
Ionized electrons in the gas are collected onto the anode. Figure 2.11 shows the
response of the system in terms of the number of electrons or ions produced in
the gas when the applied voltage is increased. The curves show six regions, each
of which has different properties. The higher and lower curves pertain to particles
that deposit more and less energy, respectively, in the detector.
In Region 1, many electrons and ions produced in the gas recombine because
the voltage applied between cathode and anode is not large enough to collect all
electrons. This region is not useful for counting radiation.
Region 2 is the saturation region. The potential difference is sufficient to collect
all freed electrons. The resulting very weak current or pulse over brief periods is
measured with extremely sensitive electrometer devices. A detector working in the
saturation region is called an ionization chamber. Its output is proportional to the
deposited radiation energy. Internal or thin-window ionization chambers are used
as alpha-particle and fission-fragment detectors. External samples are measured
for beta particles and gamma rays with ionization detectors; the latter are larger
and contain counting gas at elevated pressure.
Cathode
Source Anode
Output signal
Window
FIGURE 2.11. Variation of pulse height with applied voltage in a counter. (From Knoll 1989,
p. 162.)
In Region 3, the proportional region, the applied voltage is strong enough that the
electrons freed by the initial radiation are accelerated, so that they, in turn, ionize
additional atoms or molecules (secondary ionization) to free more electrons. This
electron multiplication generates an avalanche toward the anode for each primary
electron that was freed. The applied voltage domain is called the proportional
region because each avalanche is characterized by the same electron multiplication
at a given applied voltage. The output signal is directly proportional to the deposited
energy, although each pulse is many times larger than in the ionization region. As
the applied voltage is increased, the amplification increases uniformly for the entire
range of deposited energy.
A problem in proportional counters operating at these higher voltages is produc-
tion of a secondary avalanche due to molecular excitation, which interferes with
the detection of subsequent pulses. To prevent excitation that causes this secondary
avalanche, a quenching agent is added to the fill gas (see Section 8.5.1). The multi-
plication factor in this region is about 104 , which requires the measurement device
to be far less sensitive than that for the ionization chambers. The proportional
counter system has a preamplifier and a linear amplifier.
In Region 4, the proportionality of the output signal to the deposited energy at
a given applied voltage no longer applies. Amplification of the greater deposited
energy reaches its limit while that for the lesser deposited energy continues to
increase. This region is called the region of limited proportionality and is usually
avoided as a detection region.
2. Radiation Detection Principles 31
FIGURE 2.12. Plateau regions for both alpha and beta particles in a proportional counter.
The third useful region for detection of radiation is Region 5, the G-M region.
This applied voltage is high enough that any deposited energy produces sufficient
secondary electrons to discharge the entire counting gas. The linear amplifier is
no longer needed. One no longer can distinguish between a small and a large
deposition of energy. This discharge must be quenched so that the next pulse can
be detected. Either the applied voltage must be removed briefly or quenching gases
(see Section 8.5.1) must be added.
In Region 6, an electrical discharge occurs between the electrodes. This voltage
region has been used for some purposes but generally is avoided because the
discharge can disable the detector.
Although gas-filled detectors are operated in the second, third, and fifth voltage
regions, it would be a mistake to assume that a particular detector can be used in all
of them simply by changing the applied voltage. Detector components are designed
to be used in a single voltage domain. The operating voltage then is selected at a
plateau region on the basis of the curve of count rate vs. applied voltage, shown in
Fig. 2.12.
In proportional and G-M detectors, this curve begins with zero count rate at a
relatively low voltage, reaches a plateau at intermediate voltage, and then increases
when a continuous electrical discharge occurs. The initial increasing count rate
represents the increasing number of pulses that are sufficiently amplified to pass
the lower discriminator. At the plateau, all pulses are detected. In a proportional
counter, a first plateau is reached for the much larger alpha-particle pulses, and
a second plateau at higher applied voltages is then reached for beta particles as
well as alpha particles. Detectors are operated near the middle of the plateau to
avoid erroneous data due to small fluctuations in applied voltage. In proportional
counters, the applied voltage is reduced to measure alpha particles but not beta
particles.
Radioactive source
n-type semiconductor
Depleted region
p-type semiconductor
Conducting metal
replaces electrons and positive ions. One advantage is the thousandfold greater
electron density in solids for the much greater stopping power that is desirable for
detecting gamma radiation. Another advantage is the 10-fold smaller value of w,
the energy absorption per ionization, which yields better energy resolution. The
solid medium currently is hyperpure germanium or silicon that has been prepared
to function as semiconductor material. A schematic of a reverse bias p-n junction
detector is shown in Fig. 2.13.
A p-type material is one with positive holes and an n-type material is one with
electrons in excess. To create these types of matrices, small amounts of impu-
rities are incorporated into some host materials. For example, if the host mate-
rial is silicon, which has four valence electrons, it may be “doped” with boron,
which has three valence electrons. This creates regions in the matrix where there
are “holes” for electrons to occupy; it becomes a p-type material. The energy
for creating an electron–hole pair is 3.7 eV in silicon and 3.0 eV in germa-
nium (Knoll 1989). Thin wafers of silicon diode surface barrier detectors that
have a very thin layer of gold on the surface are used for alpha-particle spec-
troscopy. Hyperpure germanium detectors, typically a closed-end coaxial, 6 cm in
2. Radiation Detection Principles 33
diameter and 6 cm or more in length, are currently used for gamma-ray spectro-
scopy.
Conversely, if phosphorus or arsenic, each with five outer shell electrons, is
inserted into the silicon crystalline material, then the resulting material will have
an “excess” of electrons in the lattice, producing the n-type material. The placement
together of an n-type and p-type creates a p-n junction.
When a positive charge is applied to the n-type semiconductor and a negative
charge is applied to the p-type material, the positive holes are attracted to the neg-
ative electrode and the electrons are attracted toward the p-n junction. This creates
the depletion layer. Radiation enters the n-type side where the depleted region
serves as the radiation-sensitive volume. There, electron–hole pairs are created
that will be rapidly collected to create the pulse for amplification. Silicon semi-
conductors of this type are generally used for beta-particle and CE spectroscopy.
Surface-barrier detectors for measuring alpha particles are formed from n-type
silicon with an oxidized p-type surface. A very thin gold layer is evaporated onto
this surface to function as one of the electrodes. The sample that emits alpha
particles faces this side.
are dispersed throughout the LS detector, whereas in other detectors the radiation
would be absorbed in the sample.
σw w E w
= = (2.18)
E E w E
36 Moses Attrep, Jr.
FIGURE 2.14. Peak resolution in gamma-ray spectra from Ge and NaI(Tl) detectors. (From
Friedlander et al. 1981, p. 259.)
The resolution of a peak is commonly described by the full width at half max-
imum, FWHM. Since FWHM equals 2.36σ , to describe the resolution in these
terms,
w √
FWHM = 2.36E = 2.36 wE (2.19)
E
At the w value of 3.0 eV for germanium and 3.7 eV for silicon, the equation yields
the FWHM of 4.7 keV for germanium at 1.33 MeV and of 9 keV for silicon at 4
MeV. The resolution measured at these energies is about 3-fold better in germanium
and 1.5-fold worse in silicon.
Two sets of factors that work in opposite directions √ are not considered in
Eq. (2.19). The equation overestimates the FWHM by F, where F is the em-
pirically observed Fano factor. The existence of this factor is attributed to the
circumstance that the generated electrons do not necessarily act independently in
producing the ionization pulse, and so the peak is narrower than when attributed
to random events. On the other hand, detector drift, noise, and incomplete carrier
collection each contributes to widening the FWHM (Knoll 1989).
The counting efficiency of these detectors is calibrated with radionuclide stan-
dards or Monte Carlo simulation (Briesmeister 1990). Typically, the alpha-particle
detector has the same efficiency for thin samples at all commonly encountered
2. Radiation Detection Principles 37
energies, while the gamma-ray detector efficiency has a maximum near 100 keV
for samples of various dimensions.
Commercially available spectrometer systems include a computer for operating
the system, storing data, analyzing data, and providing information output. These
functions are used to set counting times and periods, identify peaks by channel and
energy, store and subtract background, calculate gross and net counts in multiple
peak regions, calculate count rates, and convert count rates to disintegration rates
on the basis of stored decay scheme data. Statistics software programs are available
to estimate output data uncertainty.
transmission. When multiple beta-particle energy groups are detected, only the
more energetic beta particles in the group with the highest maximum energy can
be seen by themselves because the other groups overlap.
The advantage of an LS counter as alpha-particle spectrometer is the minimal
energy loss because the sample is integral to the detector medium. The disad-
vantages are the relatively poor resolution associated with the PMT, some edge
effects where the alpha-particle emitter is near the vial wall, and a higher radiation
background count rate. Typical FWHM is around 500 keV at 5 MeV. About 2-fold
better resolution is achieved by special pulse timing discrimination (McDowell
1992).
3
Analytical Chemistry Principles
JEFFREY LAHR and BERND KAHN
3.1. Introduction
To separate and purify the radionuclide of interest in the sample, the analyst can de-
pend on the similar behavior of the stable element and its radioisotopes. Chemical
reactions involving the radionuclide will proceed with essentially the equilibrium
and rate constants known for the stable element in the same chemical form. Slight
differences result from small differences between the isotopic mass of the radionu-
clide and the atomic mass (i.e., the weighted average of the stable isotopic masses)
of the stable element. Because of this similarity in chemical behavior, many ra-
dioanalytical chemistry procedures were adapted from classical quantitative and
qualitative analysis. For the same reason, new methods published for separating
chemical substances by processes such as precipitation, ion-exchange, solvent ex-
traction, or distillation are adapted for and applied to radionuclides. One exception
occurs when the radionuclides to be separated are two or more isotopes of the same
element. Here, effective separation can be accomplished by mass spectrometer (see
Chapter 17).
This chapter reviews classical separation techniques and their roles in the ra-
dioanalytical chemistry laboratory. Practical application to specific radionuclides
and sample types is discussed in Chapter 6.
Chemical separation of a radionuclide from other radionuclides is intended
to recover most of the radionuclide while removing most of the accompanying
contaminants. Purification specifications usually can be relaxed by selecting a
radiation detector that measures the radionuclide of interest without detecting
some of the contaminant radionuclides, either by discrimination against certain
types of radiations or by spectral energy analysis.
The measure of purification is the decontamination factor, DF, defined as the
concentration of the interfering radionuclide before separation divided by its con-
centration after separation. The required DF depends on the initial concentrations
of the interfering radionuclide and the radionuclide of interest, and the extent of
acceptable contamination when counting the emitted radiation. The DF must be
large if the radionuclide of interest is a small fraction of the total initial radionu-
clide content. The interfering radionuclide that remains in the source that is counted
39
40 Jeffrey Lahr and Bernd Kahn
should contribute not more than a few percent to the count rate of the radionuclide
of interest. Evaluation of the fractional contribution of the radiation from inter-
fering radionuclides must consider the radioactive decay of the radionuclide of
interest and the interfering radionuclides.
The separation method generally is selected from available methods that are de-
scribed as purifying the radionuclide of interest from the identified contaminant ra-
dionuclides. Scavenging steps can be inserted that separate the major contaminant
radionuclides from the radionuclide of interest. The selected method must be tested
to determine all DF values for the contaminant radionuclides or at least to demon-
strate for the sample under consideration that none interfere with the radiation
measurement. A separation step will have to be repeated or additional separation
steps will have to be added when a single step does not achieve the required DF.
Radionuclide analysis methods are published in analytical chemistry and ra-
diochemistry journals, and in methods manuals issued by nuclear facilities such
as government laboratories. For example, the Environmental Measurements Lab-
oratory Procedures manual, HASL-300 (Chieco 1997), is an excellent source.
Standard methods for radionuclide analysis (see Section 6.7) are available, and
should be used whenever appropriate. If conditions differ from those to which
published methods have been applied, radionuclide recovery and decontamination
must be tested and additional process steps may have to be inserted.
If no applicable analytical method is found, then individual separation steps
have to be selected and combined sequentially, each step to remove one or more
contaminants until all are removed to the extent necessary. The first step must match
the sample form, each subsequent purification step must match the preceding step,
and the final step must produce the counting source in its specified form. Each step
must give high recovery of the radionuclide of interest and the required removal
of interfering radionuclides. Suitably designed tracer tests can provide otherwise
unavailable information.
Each separation method also must be evaluated for suitability with the stable
(nonradioactive) dissolved substances in the sample. These substances may inter-
fere in separations by competing or blocking reactions. Separation or scavenging
steps must be introduced to remove these interfering stable substances or reduce
them to acceptable amounts.
The techniques for separating and purifying radionuclides as part of the radio-
analytical chemistry process are discussed in the following sections. Although
separate sections present the different techniques, the analyst is expected to com-
bine separation techniques that produce optimum analytical efficacy.
In Eq. (3.1), Ab+ is the cation with charge b+ and B a− is the anion with charge
a − . The solubility product is the product of the thermodynamic activities of the
component ions in solution at equilibrium:
K sp = [Ab+ ]a [B a− ]b (3.2)
Here, the term “activity” is used differently than in other chapters of this text-
book; it stands for the thermodynamic value that replaces the molar concentration
of a substance. The units of K sp are mol/l to the power (a + b). Values are listed in
handbooks, such as the CRC (Weast 1985) and Lange’s Handbook of Chemistry
(Dean 1999), as well as in general chemistry textbooks and online.
The solubility product may be calculated from the free energy of formation G
of the solids and aqueous forms (see Weast 1985) in Eq. (3.3), with the absolute
temperature T and the gas constant R in consistent units:
−G
ln K sp = (3.3)
RT
If the product of the activities of the two components initially in solution is equal
to or less than the K sp of the product compound, then no precipitate is formed.
When the product of the known initial activities of the two components exceeds the
K sp value, a precipitate is formed. One can then estimate the activity of the residual
ion of interest (say B a− ) left in solution from the known value of K sp .Consider for
simplicity that concentrations are so low that they can be substituted for activity,
and designatethe initial concentrations of the two components by a zero subscript,
K sp K sp
(B a− )b = = (3.4)
b+
(A ) a [(A )0 − [(a/b) (B )0 ] + [(a/b) (B a− )]]a
b+ a−
For example, if the iodide reagent added as carrier for radioactive iodine is
precipitated as AgI by adding silver nitrate to solution, where a = b = 1, the
fraction of iodine remaining in solution after the precipitation, (I 1− )/(I 1− )0 , is
1−
(I ) K sp
= 1− (3.6)
1−
(I )0 (I )0 [(Ag )0 − (I 1− )0 + (I 1− )]
1+
The last term in the denominator of Eq. (3.6) can be ignored when it is very small
compared to the difference between the other two terms within the brackets. The
42 Jeffrey Lahr and Bernd Kahn
3.2.2. Practice
A first separation step for a radionuclide may be coprecipitation or scavenging
removal with reagents suggested by group separations applied in the past for
qualitative and quantitative analysis. One incentive is to scavenge, as soon as
possible, any relatively intense contaminant radionuclides to retain a sample that
can be handled with less radiation exposure and contamination potential. Group
precipitation can reduce the sample volume by carrying radionuclides from a
large initial water sample on a solid that is then dissolved in a relatively small
volume. This strategy can also serve to place the radionuclides into a solution
that no longer contains interfering substances and is more amenable to subsequent
chemical processing. For example, individual rare earth elements then can be
separated from each other on an ion-exchange column with a complexing agent
under closely controlled conditions such as precise pH values.
Carriers that have proved effective for coprecipitation are the hydrous oxides
of the metals, particularly of iron, other transition metals, and aluminum. Their
efficacy is due to their large surface area, gelatinous character, and ability to coag-
ulate. Some distinction among carried ions can be achieved because, as indicated
by Table 3.1, ions are precipitated as hydroxides at various pH thresholds.
the ion concentration on the ion-exchange medium relative to the ion concentration
in solution at equilibrium (both values in units of mol/l or Bq/L:
[Ab+ ]ix
DV = (3.9)
[Ab+ ]aq
The larger the value of DV , the greater is the selectivity of the resin for that
ion. On ion-exchange resins (see below), trivalent ions are bound more strongly
than divalent ions, which are in turn bound more strongly than monovalent ions.
Among monovalent ions, the order of selectivity generally is Cs > Rb > K >
Na > Li (Walton and Rocklin 1990), which is in the order of the ionic radius. The
value of DV can vary with the saturation of the ion-exchange resin by the backing
ion; if the concentration of Ab+ is low, the concentration of B a+ will be relatively
high in both the solution and the resin phase.
The value of DV is determined by shaking or stirring a specified amount of
solution to which the ion of interest, such as the radionuclide, had been added with
a specified amount of ion-exchange medium. When the radionuclide distribution
between solid and liquid phase reaches equilibrium, the concentration of the ra-
dionuclide in each phase is measured. The batch of ion-exchange medium must
be saturated initially with the specified backing ion and the initial solution must
contain the same ion at a specified concentration. For example, if the radionuclide
is a cationic radionuclide such as 42 K+ and the system for comparison is a sodium
salt, then the ion-exchange medium must be in the sodium form and the solution
that contains 42 K+ must be at a specified sodium backing-ion concentration. The
concentration of nonradioactive potassium ion must also be specified. Any other
radionuclides from which the radionuclide of interest is to be separated must be
equilibrated under identical conditions of known volume ratios, ionic concentra-
tion, temperature, and equilibration period.
A column experiment can also yield the value of DV . The radionuclide-bearing
solution is added at one end of an ion-exchange medium column and is then eluted
and collected in incremental volumes that are measured for radionuclide content.
Ideally, the concentration of the eluted radionuclide is distributed in the incremental
volumes as a Gaussian curve. For a distance of movement along the column (in this
case, the length of the resin column), l, cross-sectional area a, interstitial column
volume occupied by solution, i, and elutriant volume Ve measured to the peak of
the radionuclide concentration curve, Eq. (3.10) applies
Ve
DV = (3.10)
la−i
Because DV is dimensionless, the numerator and denominator on the right-hand
side of Eq. (3.10) must be in the same units of volume, say cm3 . If the radionuclide
can be measured while it is moving in the column (for example, with a gamma-
ray detector behind a shield that has a slit so that only a narrow width of column
is observed), then Eq. (3.10) can be applied for movement through the column
46 Jeffrey Lahr and Bernd Kahn
without monitoring the effluent. As indicated in the preceding paragraph, the same
backing ion and identical conditions must be used in all comparisons.
Column separations are more efficient than batch separations because a column
can be viewed as a series of sequential batch separations. For column separation of
two radionuclides, the difference between corresponding values of DV represents
the number of column volumes separating the two elution peaks; if the Gaussian
curves are sufficiently narrow, little contamination may occur. For batch separa-
tions, in contrast, the value of DV indicates that some contaminant always remains,
and sequential batch separations may be needed to reduce such contamination to an
acceptable amount. Batch separations are useful if the value of DV differs by, say,
2 orders of magnitude and the solid:liquid volume ratio is selected for acceptable
discrimination. The easiest separation is a cation from anions or vice versa, so that
one form can be essentially completely retained while the other form is only in
interstitial water.
3.3.2. Practice
The process of ion-exchange was first discovered and studied in natural inorganic
compounds, of which the most abundant are the clay minerals, especially zeolites.
The latter are microporous, crystalline aluminosilicate minerals. Numerous natu-
rally occurring and synthetic zeolites exist, each with a unique three-dimensional
structure that can host cations, water, or other molecules in its void space (cavities
or channels).
Several inorganic compounds were found to be useful for specific separations
involved in radioanalytical work (Amphlett 1964). These include hydrous oxides
of chromium(III), zirconium(IV), tin(IV), and thorium(IV); aluminum molyb-
dophosphate [(NH4 )3 PO.4 12MoO.3 3H2 O]; zirconium phosphate [Zr(HPO4 ).2 H2 O
or ZrO.2 P2 O5 ]; molecular sieves (activated synthetic crystalline zeolites); and sil-
ica gel. These inorganic substances are amorphous and eventually degrade to
fine powders. They are useful for batch operations but not for high performance
chromatography.
Paper can also function as an ion-exchange medium. It has very low capacity
but is suitable for separating radionuclides at their low concentrations. Typically,
paper chromatography is performed on strips through which a selected solvent
flows, and distinguishes radionuclides by their path lengths along the strip.
Once ion-exchange resins were synthesized in 1935, these organic exchangers
replaced zeolites in applications (industrial and analytical) and in scientific investi-
gation. The advent of nuclear power in mid-twentieth century stimulated renewed
interest in inorganic exchangers that were stable at high temperatures and could
withstand the effects of high doses of radiation.
The ion-exchange resins commonly used for separations in a radioanalytical
chemistry laboratory are solid organic structures with ion-exchange sites, or with
attached substances that function as exchange sites. Some common organic ion-
exchange media are listed in Table 3.3. The nature of the functional group defines
the properties of the resin. Resins with fixed positive charge groups can exchange
3. Analytical Chemistry Principles 47
anions and are thus called anion-exchange resins; resins with fixed negatively
charged groups exchange cations and are called cation-exchange resins.
A few commercially available ion-exchange resins are used for most published
separations. They are defined by functional group, ion-exchange capacity (in
meq g−1 ), mesh size, cross-linkage, density, and ionic form. Each descriptor af-
fects the value of DV in qualitatively understood ways (Korkisch 1989), so that the
analyst may select a resin with optimum characteristics for the intended separation.
The ionic form of the purchased resin can be replaced by washing the resin with
a solution that contains the new ion until complete conversion is demonstrated by
tests of the effluent solution.
Selection of ion-exchange column dimensions is based on the amount of the
resin needed to hold an amount of ion related to the sample (its capacity) and
to achieve separation of the radionuclide. A longer and narrower column permits
better separation, but the cross-sectional area must be sufficient to minimize wall
effects that lead to the liquid flowing along the wall in preference to passing through
the resin. Although the analyst may desire the most rapid flow permitted by column
flow resistance, the flow must be sufficiently slow to permit local equilibrium and
to avoid disrupting column structure by channeling or introducing air bubbles.
Flow may be either upward or downward; the former tends to decrease disruption
of column structure by air bubbles.
Columns must be filled carefully for uniformity of the resin bed without voids,
and maintained to prevent drying and the resulting separation within the column.
In many applications, columns can be reused for a limited number of times after
being washed with water and then regenerated with the ionic solution that re-
turns them to the original form. Resin reuse is not possible for applications that
partially destroy resin, clog the column with solids, coat the resin with emul-
sions, or load the ion-exchange sites with various ill-defined ions. Commercial
ion-exchange resins usually are relatively stable but decompose slowly with time,
or more rapidly when attacked by strong oxidizing agents (Rieman and Walton
1970).
The separation typically is performed by adding from a few milliliters to a few
liters of the radionuclide-bearing sample to a resin column of about 5–50 ml,
and then washing the radionuclides sorbed on the column with water or a dilute
reagent. In the sorption phase, radionuclides of interest either are retained at the
48 Jeffrey Lahr and Bernd Kahn
inflow end of the resin or move through the column but do not break through. The
radionuclide is then eluted with a selected reagent of experimentally determined
number of column volumes. Several radionuclides may be eluted in succession
with different types or strengths of elutriant. Used in the scavenging mode, the
radionuclide of interest passes through the column while contaminants remain on
the resin.
Researchers systematically examined values of the distribution coefficient for
various resins and solutions across the entire periodic table. These distribution
coefficients are available in tables or graphs of the distribution coefficient vs. acid
concentration. One well-known example, in Fig. 3.1, shows ln DV vs. concentra-
tion of HCl (Kraus and Nelson 1956). Similar figures are available for HBr, HNO3 ,
HClO4 , and H2 SO4 , as well as selected acid/alcohol combinations (Korkisch
1989).
The distinction among oxidation states in Fig. 3.1 suggests oxidation or reduc-
tion as a convenient technique for eluting a radionuclide that is strongly retained
on the ion-exchange resin at its original oxidation state. Taken to its extreme, a
cation held on the ion-exchange resin is removed completely by being converted to
an anion, and vice versa, or to a nonionic state. The same principle applies when a
metal ion, by addition of a complex-forming agent, is retained on or removed from
the ion-exchange resin. Such complexes may be uncharged, of the same charge as
the original ion, or of the opposite charge. The curves in Fig. 3.1 show the effect
of complex formation on retention by ion-exchange in the metal–chloride system.
Such complexes carry a negative charge when fully coordinated, and are adsorbed
by the anion exchanger. The analyte may be eluted by changing the concentration
to cause dissociation of the anionic complexes, or by changing the oxidation state
of the metal ion.
Use of complex-forming agents in solution can increase the selectivity of an
ion-exchange procedure. An example is given in Fig. 3.2 for separating rare earth
ions complexed with ethylene dinitrilotetraacetic acid (EDTA). The rare earth ions
are held on the cation-exchange resin when not complexed and are released by
changing the pH value so that they become complexed. The curves in Fig. 3.2
indicate that the radionuclide represented by the curve on the extreme right (La3+ )
can be completely retained on the column under the study conditions at pH from
0.4 to 2.8, and eluted completely above pH 3.9. The La+3 is separated completely
at pH 2.8 from the radionuclide represented by the fifth curve (Zn+2 ) from the
right, and from all ions to the left of the Zn+2 curve. It is partially separated from
Sm+3 and UO+2 2 ions represented by the other two curves on the right. By operating
at pH 3.4, approximately one-half of La+3 is retained, and is separated from all
radionuclides to the left except UO+2 2 .
Ion-exchange resins have also been loaded with counter ions for in situ pre-
cipitation; for example, an anion-exchange resin in the sulfate form can collect
226
Ra and 90 Sr, or one in the tetraphenyl borate form can collect 137 Cs (Cesarano
et al. 1965). The radionuclide of interest is then eluted with a solution in which it
dissolves.
No adsorption M 12
Slight adsorption M
Strong adsorption
3. Analytical Chemistry Principles
FIGURE 3.1. DV on anion-exchange resin as a function of HCl concentration. (Figure from Kraus and Nelson 1956.)
49
50 Jeffrey Lahr and Bernd Kahn
100
90
80
Metal ion on resin (%)
70 Th+4 Yb+3
50
Sc+3
40
30
20
Cu+4
10 Zr+4 Th+4
Bi+3 Yb+3
Fe+3
0
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0
pH
FIGURE 3.2. Elution of rare earths from a Dowex-50 cation-exchange resin, as a function
of EDTA solution pH. (Figure from Fritz and Umbreit 1958, p. 513.)
Corg
D= (3.11)
Caq
3. Analytical Chemistry Principles 51
The subscripts org and aq refer to the organic and aqueous phase, respectively, and
C is the concentration of the radioelement at equilibrium.
The extraction yield E is defined as the ratio of the amount of solute extracted
to its initial total amount. Relative to the volume of solution V ,
D
E= (3.12)
Vaq
D+
Vorg
For extractions repeated n times with fresh portions of the organic phase at
constant Vaq and Vorg , the total recovered fraction E r is defined as
Vorg −n
Er = 1 − 1 + D (3.13)
Vaq
From Eq. (3.13), the fraction of solute remaining in the aqueous phase is [1 +
D(Vorg /Vaq )]−n .
3.4.2. Practice
Extractant reagents and solvents are selected for optimum separation of the ra-
dionuclide of interest from contaminants, and also with regard to chemical stabil-
ity, purity, and minimal hazard potential. The reagents and solvents may have to
be stored under conditions that avoid degradation, and purified to remove minor
contaminants that can interfere with separations. Use of chemicals listed as haz-
ardous material may be feasible with appropriate care, but later disposal may be
difficult.
Equilibration usually is reached in a few minutes, but should be checked for
each method. Typically, the radionuclide in the aqueous phase is extracted into
the organic phase under one set of conditions, and is then back-extracted under a
second set of conditions. Washing steps commonly are inserted after extraction to
improve the specificity of the radionuclide transfer. A single extraction and back-
extraction cycle may suffice for purification or several cycles may be necessary. In
analogy to ion-exchange systems, column separations have been developed with
countercurrent flow of the two liquids.
Solvent extraction systems may be characterized by type of reagent used as
extractant or by chemical species extracted. Data for the extraction of elements is
in tabular form listed by element, plots of D vs. pH curves in periodic table format,
and distribution coefficient values for multiple elements plotted at specific phase
compositions for a given extraction system. The many solvent extraction systems
that have been reported provide many options to select a suitable system to solve
a given problem. A brief description of extractant categories is given here with
examples, noting that many reagents will fit more than one category according
to the environment in which they are used (Marcus and Kertes 1969, Sekine and
Hasegaw 1977).
Nonsolvating solvents may be used for nonelectrolyte molecular species such as
rare gases, halogens, interhalogen compounds, and some metal halide complexes.
52 Jeffrey Lahr and Bernd Kahn
This equation is analogous to Eq. (3.7). The amine and its inert solvent—for
example, benzene, toluene, xylene, or kerosene—strongly extract the mineral acids
HA and then extract anions B − that are retained by ion-exchange.
Figure 3.3 shows ln D with triisooctyl amine of many polyvalent metal ions
that form anionic chloro complexes in hydrochloric acid. Additional distribution
data are available for metal nitrates, sulfates, and oxy anions that can be extracted.
These ions can then be back-extracted into less acid systems in which they are no
longer anionic complexes.
Neutral extractants may be used for uncharged metal complexes, ionic salts, and
strong acids. Examples include ethers such as diethylether and diisopropylether;
ketones such as methylisobutyl ketone (MIBK or hexone); and neutral organophos-
phorous compounds such as tri-n-butylphosphate (TBP), tri-n-octylphosphine
oxide (TOPO), triphenylphosphate (TPP), and triphenyphosphine oxide (TPPO).
Such neutral extractants have long been used for extracting actinides and lan-
thanides from nitric acid solutions.
3. Analytical Chemistry Principles 53
One example is diethyl ether as an extractant for iron as the chloro complex
HFeCl4 or the cyano complex NH4 Fe(SCN)4 . Many of the radionuclides that form
these molecules can be separated from each other under different conditions related
to redox potential or pH.
54 Jeffrey Lahr and Bernd Kahn
1000
Plutonium(IV)
100
10 Uranium(VI)
Plutonium(VI)
D 1.0
0.1
Chromium(VI)
Chromium(VI)
0.01
Americium(III)
0.001
0 2 4 6 8 10 12 14
Nitric acid molarity
FIGURE 3.4. D as a function of nitric acid pH for extraction of metal ion into 0.1 M TOPO
in cyclohexane. (Figure from Martin et al. 1961, p. 99.)
56 Jeffrey Lahr and Bernd Kahn
1000
Uranium(VI)
Plutonium(IV)
100
Plutonium(VI)
D 10
1.0
Chromium(VI)
Plutonium(III)
0.1
0 2 4 6 8 10 12
FIGURE 3.5. D as a function of hydrochloric acid pH for extraction of metal ion into 0.1 M
TOPO in cyclohexane. (Figure from Martin et al. 1961, p. 102.)
H He
abcd a
Li Be B C N O F Ne
a bdf bcd abcd abcd abcd a
Na Mg Al Si P S Cl Ar
a d bd abcd abcd abcd a
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
a d d g e d bd bd abcd bcd abd ad
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
a d d d d cd cd a a a a bd bd bcd abd ad
Cs Ba La Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
a a d d d cd cd d a ad a ab ad ab ad
Fr Ra Ac
a
Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr
d d d d
FIGURE 3.6. Distillation separations by species. (Figure from Coomber 1975, p. 307.)
carrier gas flow or reagent addition. The condenser tube is cooled with water and
leads into the collector, which may either be dry or contain a solution to col-
lect the condensate. For optimum decontamination, the distillation process should
minimize liquid-droplet carry-over.
Tritium as tritiated water undoubtedly is the most common radionuclide purified
by distillation. For separating tritiated water from samples such as biological ma-
terial or vegetation, azeotropic distillation can be more effective (Moghissi et al.
1973). In this method, an organic solvent not miscible with water is mixed inti-
mately with the tritiated aqueous sample. An excellent website on azeotropic distil-
lation can be found at http://www.chemstations.net/documents/DISTILLATION.
PDF. Cyclohexane, which forms a constant-boiling mixture at 94◦ C is used, as
are other organic liquids, such as para-xylene and toluene. The distillation ap-
paratus is assembled with the azeotropic distillation receiver shown in Fig. 3.8,
rather than the usual collection flask. During distillation, the azeotrope separates
into an aqueous and an organic phase in the receiver. The aqueous phase is used
for analysis while the organic phase returns to the distilling flask. Distillation can
continue until essentially all of the water has been collected, achieving complete
distillation. This is particularly important in tritium purification, where incomplete
distillation will enrich the residue in 3 H (see Section 4.8) and yield results that are
underestimates by about 10%.
Many of the elements footnoted in Fig. 3.6 by the letters c or d have been
separated from compounds that have higher boiling points by heating the salts
in a volatilization apparatus at a relatively elevated temperature. The vapor is
3. Analytical Chemistry Principles 59
FIGURE 3.7. Simple distillation apparatus. (Figure modified from Chemweb 2005.)
To condenser
each radionuclide. The technique is occasionally used for rapidly separating a more
volatile radionuclide from an irradiated target, but is limited in the radioanalytical
chemistry laboratory to preparations of purified compounds in cases where the
apparatus and skill are available. An example of this separation is the volatilization
of plutonium chloride from soil in a gas chromatograph (Dienstbach and Bachmann
1980). Another example is the collection of 90 Sr as the chloride salt on a cooled
target when the source is heated between 900 and 1400◦ C; when heated above
1400◦ C, its daughter 90 Y is volatilized (Sherwin 1951).
Radionuclides that are volatilized during ashing or freeze-drying, can be col-
lected in a commercially available apparatus. For freeze-drying, the sample is
placed into a vacuum distillation apparatus and the vaporized radionuclide is col-
lected at a cooled surface or on a sorbent material.
1
= E o + 0.05916 log − E over (3.20)
Ag+
(aq)
62 Jeffrey Lahr and Bernd Kahn
H He
Li Be B C N O F Ne
Na Mg Al Si P S Cl Ar
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
X X X X X X
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
X X X X X X X X
Cs Ba La Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
X X X X X X X X X
Fr Ra Ac
X X
Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr
X X X X X X X
FIGURE 3.9. Electrodeposition separations (shown with X). (Data from Blanchard et al.
1960.)
The overvoltage or overpotential E over is inserted in Eq. (3.20) to adjust for other
processes that compete in the system and make electrodeposition less than ideally
efficient. These processes are irreversible and include the effects of the decompo-
sition of water, other solutes, and imperfections in the electrode surface. Because
of these processes, a greater potential difference than calculated from the reference
potential and the ionic concentration must be applied in order to achieve deposi-
tion. For the same reason, spontaneous deposition, inferred from a positive value
of E o , may not occur if the overvoltage exceeds it. Overvoltage effects occur at
both the cathode and the anode.
The positive value of E o for many metals (for Ag+ , E ◦ = 0.799 V) and the ad-
ditional effect of low concentration from the second term on the right in Eq. (3.20)
suggest that these metals will deposit readily at the cathode to which a low negative
voltage, i.e., a few tenths of a volt, is applied, or even spontaneously. The equi-
librium voltage E o applies to an electrode—in this case, the cathode–uniformly
constructed of the same metal as the ion; the initial voltage required for deposition
is different for a cathode of a different material. Only when a coating of the same
metal has been deposited on the cathode surface, the value of E o will apply.
The current, which is proportional to the amount deposited, can be calculated
to be small at low concentrations. Practical requirements for applied voltage and
current will differ because of the above-cited competing reactions by the water
and other substances in solution.
Most metals except the alkalis and alkaline earths can be electroplated at the
cathode with suitable applied voltage from acid solutions. Relatively early expe-
rience with electrodeposition of various metals is summarized in Fig. 3.9. The
process typically is applied for carrier-free or low-concentration samples to pre-
pare sources for alpha-particle spectral analysis. It is also useful for depositing thin
sources for counting radionuclides that emit beta-particles with low maximum en-
ergy.
Because electrodeposition usually is the last procedural step, the composition of
the solution can be controlled for ease of electrodeposition. The reagents that are
present may consist of an acid to avoid hydrolysis and a redox agent to retain the
radionuclide in a state suitable for final reduction. As shown in Table 3.4, metals
can be grouped for electrodeposition analysis based on their standard electrode
potentials.
Some elements are not suitable for electrodeposition from aqueous solution as
the metal. Among these are the radionuclides plutonium, uranium, and thorium,
which are prepared for alpha-particle spectral analysis by deposition of oxides.
Other metals, such as lead, can also be deposited as oxides under empirically
derived conditions (Laitinen and Watkins 1975).
4
Radioanalytical Chemistry Principles
and Practices
JEFFREY LAHR,1 BERND KAHN,1 and STAN MORTON2
4.1. Introduction
The preceding chapter describes methods that can be applied to the analysis of both
radionuclides and their stable element counterparts. One important difference to
consider is the extremely low amounts of radionuclides generally analyzed in the
radioanalytical chemistry laboratory. As noted in Section 13.1, the environmental
samples in a radioanalytical chemistry laboratory typically are in the picocurie to
nanocurie (0.037 to 37 Bq) range; this corresponds to a radionuclide sample mass
of around 10−15 g, depending on half-life and molecular weight of the radionuclide.
Precise measurement of such a low amount can be achieved because radionuclides
emit energetic radiation.
The low concentration of the typical radionuclide sample affects some aspects
of its behavior as a solute. One cause is believed to be interactions with surfaces
and other solutes that are not taken into consideration for species at the more
usual concentrations. The chemical status of a radionuclide sample may also be
affected by the energy liberated when a radionuclide is created and the energy de-
posited when the emitted radiation passes through the solvent medium (Vertes et al.
2003).
The peculiarities of radionuclides at extremely low concentration were of great
interest to radiochemists during the first half of the twentieth century and are
described in excellent texts such as Wahl and Bonner (1951) and Haissinsky (1964).
Some of these effects still are not clearly understood. Considered in this chapter
are issues pertinent to radioanalytical chemistry, including:
r Production of radionuclides
r Conventional equilibria values extended to very low concentrations
r Low-concentration “radiocolloidal” behavior
r Sample preservation
1
Environmental Radiation Branch, Georgia Tech Research Institute, Georgia Institute of
Technology, Atlanta, GA 30332
2
Radiobioassay Programs, General Engineering Laboratories, 1111 North Mission Park
Blvd. Suite #1070, Chandler, AZ 85224
64
4. Radioanalytical Chemistry Principles and Practices 65
r Sample dissolution
r Carrier or tracer addition
r Separation for purification
r Source preparation for radiation measurement
(>99%) dissolved by HF and HNO3 in 90-ml vessels in a few hours under pre-
scribed conditions (Garcia and Kahn 2001).
4.6.2. Fusion
Fusion is employed to decompose solids that are difficult to dissolve in acids but re-
act at high temperatures to form soluble compounds. To promote the fusion process,
a substance called “flux” is employed. Mixing the flux with the insoluble matrix
and applying heat can produce complete decomposition. A favorable reagent mix
and amount can be selected from published reports (Sulcek and Povondra 1989,
Bock 1979), but the sample matrix must be tested to confirm the effectiveness of
reagent choice and amount. Fusion can decompose environmental and biological
matrices (Sill et al. 1974, Williams and Grothaus 1984), even with high silica and
alumina content.
Fusion processes can be grouped into acid–base reactions (carbonates, borates,
hydroxides, disulfates, fluorides, and boron oxide) and redox reactions (alkaline
fusion agent plus oxidant or reductant). Common fluxes are listed in Table 4.2.
Fluoride-pyrosulfate and carbonate-bisulfate fusions are used to decompose soil
and fecal samples.
Disadvantages of the process are loss of volatile radionuclides at the elevated
temperature and the large amount of the fusion reagent added to the sample. Han-
dling the crucibles at the specified high temperatures is a safety concern. Platinum
crucibles are expensive, while other crucibles (e.g., nickel and iron) are attacked
by some reagents and contribute a contaminant to the sample.
may not respond identically to dissolution. Another gap may occur at the end of
the procedure if carrier yield is determined by a method other than precipitation,
such as spectrometric measurement of an aliquot, or if the precipitate is dissolved
for liquid scintillation counting. In such cases, the last step must be quantitatively
controlled.
28 mmol l−1 because the product of the concentrations of the two reagents of 1.7 ×
10−12 (mmol l−1 )2 is less than the solubility product for silver iodide of 8.49 ×
10−11 (mmol l−1 )2 . The precipitate may form, however, if even small amounts of
additional iodide were in the water from natural or man-made sources, or if the
thermodynamic activities exceeded component concentrations because of other
salts in solution.
Although chemical reactions of radionuclides are assumed, for practical pur-
poses, to be identical to those of their natural stable isotope mixtures, minor dif-
ferences in equilibrium distribution exists, notably in the case of tritium relative
to stable hydrogen, but also to a small extent for others, such as 14 C relative to
stable carbon. During water distillation to purify it for tritium analysis, the tri-
tium concentration of the residue is enriched by about 10% relative to the vapor.
Only when all of the water is distilled will the specific activity of tritium (and
the concentration in water, in units of Bq l−1 ) be the same in the distillate and
the sample (Baumgartner and Kim 1990). Moreover, because of such differences
in equilibrium constants, the tritium concentration can be enriched approximately
10-fold by electrolysis of a water sample from which the produced hydrogen gas
escapes into the air (NCRP 1976a).
Chemical procedures for purifying radionuclides that have no stable isotopes
are based on studies of radionuclide behavior performed over the years. A crucial
aspect of analytical procedures for these elements is addition of reagents that
maintain the radionuclide in the redox state appropriate for the separation step.
The essential tools of radioanalytical chemistry practice are the 50-ml glass
centrifuge tube and the stirring rod that are used for precipitation separations.
Carriers are added with pipettes, and reagents, from dropper bottles. Heat is applied
by a Bunsen or Meeker burner, and solutions are cooled in an ice bath. Plastic is
substituted for glass for convenience if no direct heat needs be applied, and by
necessity when hydrofluoric acid is used. The precipitate is separated from its
supernatant solution during the purification stage by centrifuging, and in the final
stage, on a 2.5-cm-diameter filter in a filtration apparatus.
5.1. Introduction
Sample collection and preparation are inextricably linked to the practice of radio-
analytical chemistry, despite the fact that two separate groups will perform these
activities. Ideally, the radioanalytical chemist is part of the team that plans the
sampling effort and prepares the quality assurance project plan (see Section 11.1).
Such participation in the planning phase benefits the entire analytical process. The
radioanalytical chemist can tailor the analytical approach to both the sample media
and the radionuclides in the media. Information on the purposes for which the lab-
oratory produces data can enable the radioanalytical chemist to devise laboratory
practices to match required detection sensitivity, sample submission rate, and re-
porting style. On the other hand, the analyst can describe the capabilities and limits
of the laboratory to the planning group. Such participation in the planning process
can prepare a cohesive sampling and laboratory effort that produces defensible
results for the client.
To serve as an effective member of the team, the radioanalytical chemist must
have some knowledge of sampling methods pertaining to various matrices. The
planning dialogue can influence specifications of sample collection by location,
frequency, size, and techniques, particularly when the suite of radionuclide samples
is from a complex environment. Sections 5.2 to 5.9 also briefly describe protocols
for sample preservation between collection and analysis.
77
78 Robert Rosson
Those planning the sampling effort should initially consider the following:
r The radionuclides produced or used in the facility or project
r Radionuclide activity levels
r Variability in radionuclide activity with time and location
r Chemical forms of the radionuclides
r Sampling circumstances, e.g., routine operation or specific incidents
r Radionuclide transport by various media.
These initial items of information guide sampling location, volume, and initial
preservation steps. Implicitly, they also guide laboratory selection; incident re-
sponse will likely require a quick turnaround time and the capability of handling
higher radioactivity levels (see Section 13.1), which narrows the choice of appro-
priate laboratories.
Once the sample has been taken, and a laboratory chosen, this information
goes with the sample as part of the chain of custody documentation. When the
laboratory accepts the sample, the sample information is reviewed to extract those
facts directly pertinent to the analytical process:
r Sample form and matrix
r Analytical requirements
r Format for results
r Sample load
The sample form and matrix type control the initial sample treatment in the labora-
tory. The analytical requirements dictate the chemical and instrumental procedures.
The format specifies units for reporting values, uncertainty, and detection limits,
and the context in which this information will be reported. The sample load controls
laboratory staffing, scheduling, and turnaround time.
The DQO checklist for the laboratory is a comprehensive compilation of all
these pieces of information. It will include descriptions of the following:
r Sample type
r Description of site parameters (e.g., flow rate of the sampled medium)
r Sample collection, storage, and preservation methods
r Sample matrix, including the radionuclides expected and the radionuclides of
interest
r Predicted radionuclide concentrations
r Required detection limits
r Available radiation detection instruments
r Available radioanalytical chemistry methods
r Estimated detection limits for selected method and instruments
r Requirements of sample size and counting times to meet specifications
r Personnel skill, time, and cost estimate
The usual types of samples are described in Table 5.1 and the following sections;
their analysis is discussed in Section 6.2. Some of these sample types are obtained
with sampling equipment such as air and water collectors placed at selected
5. Sample Collection and Preparation 79
locations (Budnitz et al. 1983), while other samples are collected by hand, as
in the case of soil and vegetation. Particularly important for hand-collected sam-
ples is the reported description of sampling site and process in sufficient detail
to permit comparisons among sampling periods and locations. All information
regarding the sample should be reported on the chain of custody form (see Section
11.2.5) that accompanies the sample to the laboratory.
Identifying the type of sample matrix and radionuclide of interest helps to de-
termine a pretreatment protocol. Some form of pretreatment is usually required
for environmental samples. The half-lives of both the radionuclides of interest and
interfering radionuclides must be considered to decide how quickly to perform
the analysis. The radiation type and energy must be known to select the radiation
detector, and possibly the radioanalytical chemistry method to prepare the source
for counting.
The concentration of the radionuclides of interest may be inferred from the sam-
pling location. Natural and man-made background radiation values are summa-
rized by the National Council on Radiation Protection and Measurements (NCRP
1987a) and the United Nations Scientific Committee on the effects of atomic ra-
diation (UNSCEAR 2000a,b).
Estimation of the detection limit of the measurement instrument in the time
period available for counting, linked with the required detectable concentration and
the expected concentration of the radionuclide to be determined, guides selection
of sample size for the analysis. Calculation of the minimum detectable activity
is discussed in Section 10.4.2. Additional documents of interest are Altshuler
and Pasternack (1963), Pasternack and Harley (1971), and Currie (1968). The
terms minimum detectable concentration and lower limit of detection also have
widespread use. A document that addresses these and other topics pertinent to
radiation monitoring and measurement was developed by a committee of the Health
Physics Society (EPA 1980a).
All sample information must be taken into account to develop an estimate of
the time and resources necessary to process the sample to the satisfaction of the
80 Robert Rosson
can be used only with low-flow-rate systems, while paper and glass fiber filters are
sturdier and more easily handled. These filters have high collection efficiency for
particles with sizes in the respirable range, i.e., those with an aerodynamic median
diameter of around 0.3 μm (Lockhart et al. 1964). All except membrane filters
have low-pressure drops that permit high flow rates for collecting large volumes
of air to achieve increased measurement sensitivity. The filters can meet standard
practice requirements of 99% removal of respirable particles at the operating air
velocity and pressure drop (NCRP 1976b, Corley et al. 1977, EPA 2004).
In view of the influence of particle size on inhalation radiation exposure, some
collectors of discrete particle sizes are deployed, typically for research rather than
routine monitoring. Impactors with metal plates or Mylar foils that are sprayed
with silicone to reduce particle bounce (ACGIH 2001) have proven useful, but can
challenge radioanalytical chemistry preparations. An aerosol spectrometer with a
series of metal screens (EPA 2004) may present less of a chemical preparation
problem.
Collector location and collection period must be selected to avoid excess mass
loading. Excessive deposition of matter on the filter will reduce the air flow-rate,
change retention efficiency, and result in some collected solids falling off when
the filter is handled.
The duration of sampling (start and end times) and the flow rate through the filter
must be recorded. The type of pump, the kind of filter substrate (cellulose, glass
fiber, polystyrene, etc.), and even the housing for the sampling station should be
recorded because these factors can affect data application. Interruption in sample
collection or the air-flow record should be avoided because the former causes a
gap in the monitoring program and the latter makes the radionuclide measurement
pointless because the air volume is unknown.
Radionuclides in particulate samples collected on filter substrates usually are
measured first by gamma-ray spectral analysis of the filter enclosed in a thin-walled
envelope. Gross alpha- and beta-particle analysis by low-background proportional
counter follows (see Section 7.2.4). The particles collected from the air stream
include radioactive progeny of natural 222 Rn and 220 Rn. These radionuclides always
present a radiation background to any targeted measurement of radionuclides. To
eliminate interference from these progeny in counting, the filters are held for a time
between collection and counting to allow for their decay; the short-lived daughters
of 222 Rn decay mostly in 2 h and those of 220 Rn decay in 4 days. The long-lived
progeny of 222 Rn—210 Pb, 210 Bi, and 210 Po—and cosmic-ray-produced 7 Be remain
on the filters after 4 days.
A specified fraction of the filter is taken for radiochemical analysis. The first
step is filter dissolution, as discussed in Section 6.2.1. The remainder of the filter is
archived for possible future analyses. The archive should also contain unexposed
(blank) filters.
Airborne particulate radionuclides are also collected in deposition trays. An
upward-facing adhesive “gummed film” in the tray has been used to retain the de-
posited solids during the collection period (NCRP 1976b). The tray collects solids
during dry periods and precipitation events unless a sliding cover is activated to
close the tray during precipitation. The cover can be moved by a sensor of increased
82 Robert Rosson
conductivity that causes the cover to close the tray at the beginning of precipita-
tion and opens it after the end (Chieco 1997). The measured deposition during dry
periods (for a collector with a cover) or during the entire period (for a collector
without cover) can be compared to the results measured for a rainwater collector
filter discussed in Section 5.4.3. The particles collected during precipitation may
include some dry deposition that the rainwater washed out of the collector and
onto the filter.
Glass wool
Silica gel
Glass wool FIGURE 5.2. Silica gel column for collecting tritiated
water vapor.
5. Sample Collection and Preparation 83
to sorb the maximum humidity expected for the volume of sampled air. Column
dimensions must accommodate this silica gel volume and the selected air flow
rate. The collected water or used silica gel is taken to the laboratory for tritium
analysis (see Section 6.4.1).
Special systems are used to collect separately the tritiated hydrogen gas (HT)
and organic-bound tritium from air. A two-component collection train consists of
a molecular sieve column, followed by addition of a stream of H2 carrier gas to the
sampled air before it passes through a palladium-coated molecular sieve column.
The first column collects tritiated water and the second column collects the two
other forms (Ostlund and Mason 1985). The two columns are treated separately at
the laboratory, as discussed in Section 6.2.2, to provide the three forms for separate
analyses.
Carbon-14 collection. Carbon-14 is formed in nature by cosmic-ray interactions.
It is in all carbon-containing compounds that are in equilibrium with 14 C in air at
a specific activity of 0.23 Bq g−1 carbon. Concentration measurements in carbon-
containing compounds that are no longer at equilibrium with air, such as dead trees,
are used to determine their “age”—the time period since the end of equilibrium
with airborne carbon—in terms of the fractional radioactive decay. Fluctuations of
cosmic-ray production of 14 C in air over the centuries must be considered in this
determination (NCRP 1985a). Carbon-14 is also produced at low rates in nuclear
reactors, mostly by the (n,p) reaction with 14 N and the (n,α) reaction with 17 O.
Collection of 14 C in air generally is a research effort. A large volume of air is
pumped through a collection train that consists of a bubbler with barium hydroxide
solution and a tube that contains aluminum–platinum (0.5%) catalyst, where it is
heated to 550◦ C, and then pumped through a second bubbler with barium hydroxide
solution. Air passes through the first solution, in which 14 C in the form of CO2
precipitates as barium carbonate and then passes with added CO carrier gas through
the oxidation tube in which 14 C as CO or CH4 is converted to CO2 . The gas finally
passes through the second solution in which the newly formed CO2 precipitates as
barium carbonate. The two 14 C samples are taken to the laboratory for measuring
the emitted beta particles (Kahn et al. 1971).
Radioiodine collection. Monitoring for radioiodine is associated with fallout,
nuclear facilities, and medical use. The most common radioiodine is 131 I, but
shorter-lived 132 I, 133 I, and 135 I are also fission products, and others, such as 125 I,
are also used in nuclear medicine. Some long-lived 129 I is produced in nature and
some is formed in fission and released when processing spent reactor fuel. Iodine
in air can be both gaseous and particulate. The existence of radioiodine in a number
of chemical states complicates sample collection and analysis.
The typical environmental airborne radionuclide collection train consists of a
filter followed by an activated-charcoal cartridge; particulate iodine is retained on
the filter and gaseous iodine on the cartridge (Corley et al. 1977). A more elaborate
collector train distinguishes the various forms of radioiodine. The collectors, in
order, are an air filter for particles, cadmium iodide on Chromosorb-P for I2 , 4-
iodophenol on alumina for HIO, and activated charcoal for organically bound
iodine, such as CH3 I (Keller et al. 1973).
84 Robert Rosson
Collection efficiency for radioiodine from the air stream by activated charcoal
is subject to type of charcoal, size of the granules, depth of the sampling bed, and
flow rate. The grain size should be in the range 12–30 mesh and the sampling rate,
0.03–0.09 m3 m−1 , for optimum collection (APHA 1972.). Air streams can also
be monitored for radioiodine with silver zeolite and molecular sieve, but activated
charcoal is the commonly used sorbent largely due to lower cost.
Noble gas collection. The radioactive argon, krypton, and xenon radioisotopes
are monitored near nuclear power plants and related facilities to determine the
magnitude of releases of gases generated by fission or neutron activation. Radon
radioisotopes and their particulate progeny are measured in homes and mines to
determine whether their airborne concentrations are below radiation protection
limits.
Grab sampling techniques for noble gas radionuclides include collection in
a previously folded large plastic bag or a previously evacuated sampling flask.
Low-volume samples can be taken with a hand pump; larger volume samples are
collected in metal containers under pressure of 10–30 atm (Corley et al.1977).
Flowing air is sampled for noble gases by sorption on cooled activated charcoal
or by cryogenic condensation.
Radon in air can be collected and measured with an evacuated Lucas cell (see
radium analysis in Section 6.4.1). The cell is taken to the sampling site and filled
by opening the stopcock. After closing the stopcock, the cell is returned to the
laboratory and counted between 2 h and a few days after collection. Radon is also
collected on charcoal cartridges that are measured directly by gamma-ray spectral
analysis. Radon progeny are collected on filters. They can be measured to esti-
mate radon parent concentrations. Numerous in situ detectors of radon and radon
progeny, including ionization detectors, proportional counters, and scintillation
detectors, are in use (EPA 2004).
5.4. Water
Types of samples include rain, surface water, ground water, drinking water, seawa-
ter, and wastewater. The various types have different chemical constituents, solids
content, and pH that may require distinctive collection and analysis techniques.
Information should be obtained on the expected radionuclide content of the water
to guide initial processing, sample preservation, and any requirement for rapid
shipping and prompt analysis.
Practical experience suggests that water is the easiest matrix to process. Prob-
lems with water samples include the distinction between dissolved and suspended
radionuclides and the preservation of dissolved radionuclides in the sample in
their original form. The sampling location and process can modify the sample; for
example, radionuclides in water may deposit on pipe and valve surfaces and later
may dissolve again.
Filtration of water samples at collection removes from the water the suspended
and colloidal fractions (for separate analysis, if desired) but some dissolved
5. Sample Collection and Preparation 85
the next collection period. Acid or another preservative is added to newly placed
collection bottles.
Depending on the program, rain samples are collected weekly, monthly, or quar-
terly. An adjacent rain gage is desirable to record precipitation data, or the collector
may be placed near a meteorological station so that radionuclide measurements can
be related to weather data. Radionuclide data are reported relative to the amount
of water, area of deposition, and rainfall for the period.
at locations and depths selected on the basis of groundwater flow models. Well
drilling and water sampling must follow established guidelines to yield acceptable
samples (Wood 1976, EPA 1986, Shuter and Teasdale 1989). Several wells may
be bundled at a given site to reach various depths. The number of wells drilled
is limited by their cost. A different sampling system consists of collecting water
from existing supply wells. The frequency of sampling is related to the estimated
subsurface flow rate; for example, if the estimated flow rate is a few meters per
year, only an annual sample is needed.
Samples that represent groundwater are collected after several well volumes have
been discarded unless little water is available. The sample is acidified promptly
for preservation. If one of the radionuclides is not stable in acid, a second sample
is collected and preserved as appropriate.
5.5. Milk
Milk is frequently analyzed for radionuclides in a monitoring program because
it is one of the few foods that reaches the market soon after collection. Milk
commonly is consumed within 7–11 days after milking and occasionally reaches
the consumer within 2 days. It may contain relatively short-lived radionuclides and
be a dietary source of fission-produced 89 Sr, 90 Sr, 131 I, 137 Cs, and 140 Ba, as well as
naturally occurring 40 K. A 4-l sample usually is collected for analysis. Cow’s milk
samples can be collected to represent a specific herd in the form of raw milk, or a
regional pool of pasteurized milk. Goat’s milk is collected when this medium may
88 Robert Rosson
Samples from aquatic species, i.e., fish and shellfish, may be the edible portion
to focus on human consumption, or the whole fish as indicator of radionuclide
levels in the system. Fish can be considered in several categories such as bottom
feeders (e.g., catfish), pan fish (blue gill), and predators (bass). Often, more than
one fish per species must be collected to accumulate the sample mass needed for
analysis. Fish are not good indicators of local radionuclide distribution because of
their mobility: i.e., fish caught at one location may have ingested the radionuclide
elsewhere, within the limits suggested by their normal movement or imposed by
dams.
Fish are sampled by electroshocking or with hoop nets, trout lines, or rod and
reel. Certain fish will float to the water surface after shocking (pan fish and preda-
tors), but bottom feeders such as catfish are more easily caught with hoop nets or
trout lines. Collection with rod and reel simulates sports angler catches.
Samples of meat and fish must be refrigerated for prompt analysis or frozen
for extended storage. Fish and other small animals may be dissected promptly
after collection or just before analysis. Samples should be weighed when fresh,
dried, and ashed as needed for the specified purpose in reporting radionuclide
concentrations.
Soil can accumulate radionuclides over long time periods but may not be useful for
observing short-term trends. Measurements can provide retrospective information
on radionuclide levels. Site description, location, collection date and time, sample
area and depth, and sampling method (core, template, trench) should be recorded
along with the sample identification number and name of sample collector.
Surface soil can be collected with a scoop or cookie cutter. The scoop sampler is
convenient for collecting a grab sample and transferring it directly into a collection
container. A cookie cutter is a rigidly framed apparatus that can be sunk into the
90 Robert Rosson
soil to a selected depth to select the material within the frame for removal and
analysis (FRMAC 2002). Vegetation, debris, roots, and stones should be removed
at the time of collection to obtain a sample pertinent only to soil and assure that
the soil mass is sufficient for analysis. Sandy soil usually retains only a small
fraction of radionuclides to which it is exposed, while clay can be highly retentive.
In agricultural land, surface soil with elevated radionuclide levels may have been
turned over and had chemicals added that affect soil retention capability.
If the interest in radionuclide distribution extends beneath the soil surface, cor-
ing is the preferred sampling technique. Cores are taken to a known depth at the
sampling site. Depth profiles are separated at selected vertical intervals to provide
information on area deposition and downward movement of a radionuclide. Nu-
merous cores may have to be taken because radionuclide retention can vary with
changes in soil constituents in a relatively small area. Samples at greater depths
can be obtained from well-drilling cores.
5.8.2. Sediment
Radionuclides in sediment are indicators of facility releases to bodies of water
and runoff from surfaces in the drainage region. Samples represent the fraction of
radionuclides precipitated or sorbed on suspended matter that has settled to the
bottom. If such radionuclides are in sediment, elevated levels usually can be found
where suspended material tends to settle as the water flow rate decreases, e.g., at
widening channels, on the inner side of bends, and behind dams.
Variables that affect radionuclide concentrations in sediment are water move-
ment, water quality, the presence of aquatic vegetation, and sediment characteris-
tics such as clay and organic fractions. Sediment samples collected at time intervals
should not be expected to have a consistent pattern of radionuclide concentrations
because the location sampled previously is difficult to find precisely, sediment
may have moved downstream, and older sediment may be covered by more recent
deposits.
A striking effect of the change in water quality is the increased deposition found
for 137 Cs in estuaries where fresh water contacts saltwater (Corley et al. 1977).
Impounded water that stratifies, with anoxic (oxygen deficient) conditions near
the bottom during winter and mixing in spring, can affect radionuclide solubility
(Carlton 1992). In “blackwater” streams, humic acids can complex radionuclides to
keep them soluble. Changes in water level may periodically cover soil, vegetation,
and debris that take up radionuclides, notably in swamps.
The dredge sampler for sediment acts like a scoop in collecting sediment from
the stream bed surface. The dredge has drawbacks because the sampling location
on the riverbed is uncertain, it collects only the top 15 cm of sediment, and fine
material can escape along with the water on return to the surface. A core sample
collected by the slide hammer method can be used to collect a 0.7-m-long sediment
sample in water as deep as 5 m (Byrnes 2001).
Collected samples are frozen if preservation is necessary. Both moist weight at
the time of collection and dried weight before analysis should be measured.
5. Sample Collection and Preparation 91
6.1. Introduction
A radioanalytical chemistry procedure is a series of steps that leads to the measure-
ment of radiation (or sometimes mass), with the goal of unambiguously identifying
the radionuclide of interest and determining its amount in a sample. The chem-
istry component of the analysis begins with the sample collection and handling
described in Chapter 5 and ends with the source preparation discussed in Chapter
7. The radiation emitted by the radionuclide is then measured (i.e., the sample is
“counted”), as discussed in Chapter 8. This chapter addresses the practical aspects
of the chemical and radiochemical separation processes surveyed in Chapters 3
and 4, which are central to separating interfering radionuclides and solids from a
radionuclide in preparation for counting.
Numerous separation methods of the types cited in Chapter 3 were developed and
applied in radioanalytical chemistry during the past century. The first 30 years were
devoted mostly to nuclear chemistry applications for identifying and characterizing
the naturally occurring radionuclides. In the following years, attention shifted to
the man-made ones; these activities continue, as exemplified by the work described
in Chapter 16. Currently, many methods are devoted to monitoring radionuclides
in the environment, facility effluent, process streams, and workers.
The radioanalytical chemist who is responsible for such monitoring selects or
designs procedures that meet the client’s specifications of sample type, list of
radionuclides, measurement reliability and sensitivity, and response time. The an-
alyst also considers the limits imposed by prescribed sample size, solution volume
appropriate for chemical separation, and radiation detection instruments at hand.
Potentially applicable procedures are selected for these criteria by literature review
and then evaluated in a methods development and testing process. A chemical sep-
aration procedure can be devised either by selecting the most applicable published
method and introducing any needed modifications or by combining pertinent sep-
aration steps.
One of the main efforts in measuring a radionuclide of interest is removing
interference due to other radionuclides by chemical separation, detector selection,
∗
This text owes the genesis of its content to Dr. Isabel Fisenne, Department of Homeland
Security.
1
Environmental Radiation Branch, Georgia Tech Research Institute, Georgia Institute of
Technology, Atlanta, GA 30332
93
94 Bernd Kahn, Robert Rosson, and Liz Thompson
In some instances, purification may not be necessary but can ease detection and
improve measurement.
Two or more radioisotopes of the same element that cannot be measured by
spectral analysis require integration of effective separation of impurities and radi-
ation detection of the selected distinguishing decay characteristics. To determine
the amounts of 89 Sr and 90 Sr in a sample, for example, interfering radionuclides
such as 226 Ra and 140 Ba must be removed; only then can the two strontium ra-
dioisotopes be distinguished in terms of the radioactive decay of 89 Sr, ingrowth of
the 90 Y daughter, and detector response to beta-particle energies, as discussed in
Section 6.4.1.
A testing program is needed to confirm the reliability of the selected proce-
dures, especially to achieve high chemical yield and a sufficient decontamination
factor for every interfering radionuclide. The analyst must estimate the allowable
concentrations of radionuclide impurities and matrix components; below this con-
centration, interference in measuring the radionuclide of interest is not of concern.
An ongoing quality assurance program provides feedback for maintaining or
improving each procedure during routine application. Changes are needed when
quality control results fall outside the limits of acceptability discussed in Section
10.5. Ambiguous results concerning the presence or absence of radionuclides of
interest require methods diagnostics, the topic of Chapter 12. Reported methods
that appear to be better should stimulate efforts to replace existing methods.
Whether some insoluble residue remains depends on the type of mass loading.
For example, soil on the filter may constitute a partially insoluble matrix that retains
some radionuclides, including naturally occurring uranium and thorium. Insoluble
residues of ashing and subsequent acid dissolution require fusion, as discussed in
Section 6.2.7. The two fractions are combined after the melt is dissolved in dilute
HNO3 ,and the solution is evaporated almost to dryness and is then taken up in
dilute HNO3 .
Some readily soluble filter materials are available (see Section 5.3.1), but the
most favorable material may not meet the specifications for large filter area, rapid
airflow, and long collection period. Paper filters generally are sufficiently rugged
to meet these requirements.
Removing the accumulated particles by washing is unlikely to be quantitative
or even consistent. Acid leaching can leave highly insoluble forms of several ra-
dionuclides, notably the actinides, on the filter. As an exception, some fallout
radionuclides (such as 90 Sr) are completely dissolved by acid treatment enhanced
by refluxing. The undissolved material (principally silica) is removed by filtration.
This process can be applied only to separations that have been verified experimen-
tally to show no radionuclide loss.
Tritium Samples
Airborne 3 H may be collected in any of the forms discussed in Section 5.3.2.
Samples may arrive with several forms in the same collector for separation at the
laboratory, or the forms may already be separated at collection. Most commonly,
tritiated water vapor is brought to the laboratory as a water sample. When collected
on silica gel or molecular sieve, tritiated water vapor is removed by heating the
collector material in a distillation flask above the boiling point of water. The
vapor is transferred to a separate flask and condensed. A measured amount of the
condensed water is analyzed for tritium in a liquid scintillation (LS) counter in
terms of becquerel per gram water. If the collector is silica gel, the result must
be corrected for dilution by a small amount of water originally in the silica gel
(Rosson et al. 2000). In addition, a small isotope effect, mentioned in Section 4.8,
results in distilling a greater fraction of H2 O than HTO unless the entire sample is
distilled.
6. Applied Radioanalytical Chemistry 97
The other two forms of tritium gas usually are submitted on a palladium-coated
molecular sieve collector. Any HT is converted to HTO by palladium-catalyzed
oxidation in the collector; the water is removed by heating it above its boiling
point and condensing the vapor. The sorbent with the remaining organical-bound
tritium is mixed with a Hopcalite catalyst and heated to 550◦ C to oxidize organic
gases to water, which is distilled and condensed (Ostlund and Mason 1985).
C-14 Samples
The sample for 14 C analysis, usually collected as part of a research project, has
been processed at the time of collection. The 14 C is in the form of barium carbonate,
precipitated from a barium hydroxide solution. Two samples may be submitted if
the CO2 form has been separated from the CO and organic-bound 14 C forms. The
barium carbonate precipitate can be measured directly in an LS or proportional
counter or further purified.
Radioiodine Samples
Gaseous forms of iodine generally are collected undifferentiated on an activated
charcoal cartridge after passing through a filter on which particulate iodine is col-
lected. In special studies, the different gaseous forms may be retained on separate
collectors that are brought to the laboratory for analysis of 131 I. Direct gamma-ray
spectral analysis of filter and cartridge for the characteristic 0.364-MeV (85%)
gamma ray of 131 I is the preferred method of measurement.
A fresh fission-product mixture also contains 132 I (2.3 h), 133 I (20.8 h), and 135 I
(6.6 h) that can be measured by gamma-ray spectral analysis. Measurement of 129 I
is discussed in Section 6.4.1.
6.2.3. Water
The sample may arrive unfiltered or separated as filtered water and the filter that
contains the solids. The water sample is preserved with dilute acid or a preservative
suitable for a radionuclide such as 131 I that may be lost from an acid solution.
Water without suspended solids is ready for evaporation to measure the gross
alpha- and beta-particle activity, measure gamma rays by spectral analysis, and
perform radiochemical analysis. The solids usually are counted similarly and then
processed for dissolution as described in Section 6.2.1 for subsequent radionuclide
analysis.
When a liquid volume of several liters is needed to achieve a specified sensitiv-
ity for radionuclide detection, the volume usually is reduced for ease of analysis.
Evaporation is a simple approach. A faster process—and one that is necessary for
samples of seawater or other high-salt-content solution—is precipitation of the
radionuclides of interest from the large sample volume with a carrier in bulky in-
soluble forms such as phosphate, hydroxide, or carbonate. The precipitate bearing
the radionuclides of interest is separated by decanting most of the solution and
filtering the rest, and is then dissolved for further radionuclide purification.
Certain radionuclides may be collected from even larger volumes (i.e., hundreds
of liters) by passing a stream of water through a large filter system to retain
insoluble forms and then through large ion-exchange columns to retain cations and
anions. The filter and columns can be analyzed with a gamma-ray spectrometer.
Radionuclides that do not emit gamma rays can be eluted from the ion-exchange
columns with a relatively small volume of a high-salt solution, a complexing agent,
or a redox reagent. The filter is processed as discussed in Section 6.2.1.
6.2.4. Milk
Many persons consume milk soon after collection, and the pathway for a few
radionuclides from origin to milk includes accumulation steps due to large-area
grazing by cows and their lactation process. Before the advent of gamma-ray
spectral analysis, the radionuclides 89 Sr, 90 Sr, 131 I, 137 Cs, and 140 Ba were measured
after chemical separation. Today, only the 89/90 Sr pair requires chemical separation.
The brute-force approach to sample preparation was evaporating and ashing a
liter of milk. The process requires care to avoid splattering and loss of ash from
6. Applied Radioanalytical Chemistry 99
the large mass of organic material. A major improvement was precipitation of the
organic phase of the milk with trichloroacetic acid, followed by radiochemical
analysis of radiostrontium in the aqueous phase (Murthy et al. 1960). Currently,
whole milk is passed through ion-exchange resin to sorb the radionuclide of inter-
est, notably radiostrontium, on cation-exchange resin and radioyttrium on anion-
exchange resin, at controlled pH. The radionuclide is then eluted from the resin
for measurement (Porter and Kahn 1964).
leachate/sample ratio, leach period, and temperature. The leached fraction varies
for different soils and radionuclides.
Total dissolution is needed for less soluble radionuclide forms. This goal may
be achieved by repeated treatment with solutions of the same acid or sequential
treatment with different acids. Total dissolution of gram amounts is achieved rela-
tively rapidly in a microwave oven with programmed temperatures and pressures
(Garcia and Kahn 2001). The factors discussed above also control total dissolution
and must be determined experimentally. Sample mass is limited to a few grams,
compared to order-of-magnitude larger amounts that can be treated by leaching.
Refractory species (such as zirconium oxide) are incompletely atomized at the
temperatures of a flame or a muffle furnace. If the radionuclide of interest may be
in a refractory form, the material must be mixed with fusion reagents and melted
at a high temperature. After cooling, the solidified melt is dissolved in a HNO3
solution. Fusion mixtures that have been tested for many types of soil are listed in
Table 4.2. Sample masses are limited by practical consideration of the final sample
size, taking into consideration the large amounts of fusion reagents added to the
sample.
6.2.9. Smears
Paper or cloth smears are used, often in response to regulations, to wipe surfaces of
specified area (e.g., 100 cm2 ) to check for removable radionuclides. The smears are
counted directly by gamma-ray spectral analysis. For gross alpha- or beta-particle
measurements, thin smears are counted in a proportional counter or immersed in a
cocktail for LS counting. Further analysis for a radionuclide of interest that emits
6. Applied Radioanalytical Chemistry 101
only alpha or beta particles requires ashing the smear and subsequent dissolution
with dilute mineral acid.
6.3.1. Carriers
Cationic carriers usually are obtained as chloride or nitrate forms, and anions, in
their sodium form. The reagents must be sufficiently pure so that any impurity
does not significantly affect its mass determination or radionuclide measurement.
Reagent blanks must be counted after carrier preparation to check purity, and any
significant source of radiation must be removed from the carrier.
The carrier solution should be maintained in a stable form, which may require
addition of a reagent such as a dilute acid to a carrier that may otherwise hydrolyze
over the period of use. The carrier should be filtered after preparation, checked
periodically, and discarded if solids have formed or other changes are apparent.
Care must be taken not to introduce contamination into a carrier; typically, a master
solution is stored separately from frequently used fractions.
The carrier concentration in the solution should be determined in replicate with
sufficient precision so that the uncertainty of yield measurement does not exceed
1%. The precipitate for carrier mass determination should in all ways be identical
to that obtained at the end of the procedure to avoid chemical form differences
such as water content.
The carrier is pipetted into the sample solution at the beginning of the analytical
process and mixed thoroughly with the solution. The chemical state of the carrier
must be identical to that of the radionuclide. If any difference is possible, steps
must be inserted into the procedure at this point to achieve identical form for carrier
and tracer. A common process is to oxidize and reduce the carrier and radionuclide
through all possible oxidation states. For example, if the radionuclide 131 I is not
known to be in the form of iodide and the carrier is added as iodide, the mixture is
oxidized through molecular iodine, iodate, and periodiate, and then reduced back
to iodide.
Nonisotopic carriers can be used when no isotopic carrier exists. Rhenium is a
carrier for technetium, lanthanum for promethium and plutonium, and barium for
radium. Such chemical homologs do not have identical solubility products, hence
their yields may differ and must be compared by testing. Group separations can be
102 Bernd Kahn, Robert Rosson, and Liz Thompson
performed (see Table 3.2), such as iron hydroxide precipitation to carry tetravalent
uranium and rare earth radionuclides.
So-called “holdback” carriers are added to improve separation of one radionu-
clide from another at very low concentration. A holdback carrier differs slightly
from a regular carrier in its function: a regular carrier functions to “carry” the ra-
dionuclide of interest out of solution, while a holdback carrier’s function is to keep
an interfering radionuclide in solution while the radionuclide of interest is pre-
cipitated with its regular carrier. Holdback carrier addition prevents radiocolloidal
behavior exhibited in sorption of a contaminant radionuclide on the precipitate of
the radionuclide of interest. The holdback carrier effectively maintains a consistent
decontamination factor.
6.3.2. Tracers
Tracers are used as radiometric means of determining the yield of the radionuclide
of interest. Tracers are selected for convenience in availability and counting the
emitted radiation. The two radiations may be distinguished from each other by
spectral analysis, e.g., 242 Pu tracer for 239 Pu by counting alpha particles, or by use
of different detectors, as in counting the gamma rays of the 85 Sr tracer for 90 Sr that
is determined by counting beta particles emitted by its 90 Y daughter.
Although isotopic tracers are preferred, application of nonisotopic tracers pro-
vides more choices for selecting a radionuclide with suitable radiation, e.g., 133 Ba
tracer by gamma-ray spectral analysis for 226 Ra by counting alpha particles. As
stated in the preceding section, tests must be performed to determine whether
yields for the two nonisotopic radionuclides are identical or have a constant ratio.
The radioactive tracer must not be contaminated with other radionuclides that
interfere with the yield measurement. Some actinide-series tracers have radioactive
progeny with slow ingrowth that should be identified from their decay scheme chain
and require periodic purification to remove these progeny.
The amount of tracer to be added needs to be balanced; too much may result
in some radiation cross-talk between the radionuclide of interest and the tracer,
while too little reduces the precision of yield determination. Typically, the tracer
concentration is selected to be a few times the expected concentration of the
radionuclide of interest.
element. Subsequent measurements with absorber foils (see Section 2.4.2) iden-
tified the type of emitted radiation and estimated its energy. The half-life was
determined by repeated measurements under the same conditions. This approach
was difficult for a radionuclide that emits radiation of several types and energies, for
more than one radioisotope per element, or for incompletely purified radionuclides.
These measurement techniques are reflected in the early compilations of radio-
analytical chemistry methods (Coryell and Sugarman 1951, NAS-NRC 1960). As
spectral analysis achieved better resolution and higher counting efficiency, chem-
ical separations were avoided when possible in favor of spectral analysis.
Currently, only a few routinely measured radionuclides such as 3 H, 14 C, 55 Fe,
90
Sr, 99 Tc, 129 I, 147 Pm, and many of the elements heavier than bismuth remain
dependent on chemical separations. Chemical separation still is needed for a few
reasons; a radionuclide emits no gamma rays, emits too low a fraction of gamma
rays to meet required detection levels, requires removal of solids for counting (e.g.,
alpha particles), or must be detected with greater sensitivity.
In the future, mass spectrometry (see Chapter 17) may supersede radiochemical
analysis for long-lived radionuclides and require a different set of chemical sepa-
rations. This trend is opposed to a certain extent by chemical separation processes
introduced to achieve ever lower minimum detectable activity requirements and
by the continued interest in identifying newly created radioelements.
H-3 Analysis
Tritium is produced in nuclear reactors and devices by ternary fission and neutron
activation of deuterium, boron, and lithium, and is produced naturally by cosmic
rays. It is analyzed in air, water, and biota samples and for bioassay. Water vapor in
air is collected by condensation on a cold surface, by bubbling through water, or on
a sorbent material such as silica gel from which it is then flushed above the boiling
point of water. Water from biological material is collected as vapor by heating
samples just above the boiling point of water and then condensing the vapor. The
NCRP has published a survey of tritium measurements, which provides detailed
information (NCRP 1976a).
The sample—typically about 20 ml water—is distilled for purification. Reagents
and holdback carriers may be added to the distilling flask to prevent volatile
forms of contaminant radionuclides from being distilled. Initial distillation may be
104 Bernd Kahn, Robert Rosson, and Liz Thompson
performed to separate the volatile forms, and the remaining water is then distilled
for tritium analysis. For extracting water from solid samples or tritium sorbents,
azeotropic distillation in a reflux system conveniently maintains a constant boiling
point and liquid cover. The organic solvent is refluxed while the heavier distilled
phase is 99.9% water, as described in Section 3.6.1.
The distilled water is counted in an LS system at a selected water-to-cocktail ratio
(usually 1:1). Samples may be counted without purification if other radionuclides
are known to be absent or can be differentiated clearly from tritium by pulse-
height discrimination in the detection system, and if chemicals that cause excessive
quenching or fluorescence are known to be absent. Section 15.4.3 illustrates an
extension of this measurement technique, where 3 H is measured in flowing water
with a scintillation counter.
Collection and analysis complexity introduced by forms other than tritiated
water is discussed in Section 6.2.2. Tritium is released in these forms from nuclear
facilities or converted to them in the environment. These forms generally occur
in lesser magnitude than tritiated water; they should be differentiated because of
their different pathways and radiation impacts after they enter the body. Biological
material can be dried to collect tritium as water vapor and then ashed to collect
organic tritium as HTO vapor.
6. Applied Radioanalytical Chemistry 105
Tritium has also been counted as a gas or vapor mixed with the counting gas
in a flow-through ionization or proportional counter. Flow-through detectors are
suitable as effluent monitors for tritium when tritium concentrations in air are high
relative to concentrations of other radionuclides. Discrimination between tritium
and radionuclides with more energetic beta-particle groups is achieved by pulse-
height control for the proportional counter.
Fe-55 Analysis
55
Fe is formed by the (n,γ ) reaction with 54 Fe. At the same time, 59 Fe is formed by
the (n,γ ) reaction with 58 Fe. Both radioisotopes are produced in iron and steel cas-
ings, vessels, or supports for nuclear weapons and reactors. 55 Fe (t1/2 = 2.73 years)
decays by electron capture with a K α X-ray energy of 5.89 keV (24.5%) and a K
Auger electron energy of 5.2 keV (61%). In contrast, 59 Fe (44.5 d) emits beta par-
ticles with maximum energies of 0.466 (53%) and 0.274 MeV (45%) and gamma
rays of 1.29 (43%) and 1.10 MeV (57%). The following procedure (Chieco 1997)
determines both radionuclides.
Iron chloride or nitrate carrier is added and precipitated to carry the radionuclides
as hydroxide, dissolved in strong HCl and collected as the chloro-complex on an
anion-exchange column. After washing to remove contaminants, iron is eluted
from the column with dilute HCl. Cobalt, manganese, and zinc holdback carriers
are added to the solution and iron is precipitated as the cupferate at 10◦ C. The
cupferate complex is destroyed by wet ashing and iron oxide is converted to the
chloride by boiling in HCl. The iron ion is complexed in an ammonium phosphate–
ammonium carbonate electrolyte and electroplated on a tared copper disc. The disk
is weighed to determine the iron yield and then is sprayed with a thin acrylic coating
to prevent oxidation of the iron.
59
Fe beta particles are counted with a proportional detector or its gamma rays
are analyzed with a Ge detector and spectrometer. The sample is then measured
for 55 Fe content with a thin Ge detector and spectrometer or xenon-filled X-ray
proportional detector with a thin (e.g., 140 mg cm−2 ) beryllium absorber. The 55 Fe
count rate is adjusted for background, the 59 Fe contribution, self-absorption in the
plated sample, and the chemical yield, and converted to the disintegration rate. The
activity of both radioisotopes is corrected for radioactive decay from the sampling
date.
If no 59 Fe is in the sample, as confirmed by sample analyses, it can be added
as yield tracer for 55 Fe analysis. After ion-exchange purification, the elutriant is
measured by LS counting 55 Fe and 59 Fe in different energy regions. Cross-talk
from the 59 Fe beta particles in the 55 Fe Auger electron region must be corrected,
as well as color quenching from chemicals.
Sr-90 Analysis
Analysis of 90 Sr is a classical example of a precipitation separation. Precipitation in
concentrated (around 60%) to fuming (>86%) nitric acid separates strontium and
106 Bernd Kahn, Robert Rosson, and Liz Thompson
barium from most other elements (Sunderman and Townley 1960), and is partic-
ularly useful because so many other fission-produced radionuclides are insoluble
in basic solution but soluble in acid.
After strontium carrier is added to a small volume (<10 ml) of 90 Sr solution,
sufficient fuming nitric acid is added to attain a nitric acid concentration of 14–
16 N. The solution with strontium nitrate precipitate is cooled in an ice bath
and then centrifuged. The supernatant solution is thoroughly decanted and the
strontium nitrate precipitate is dissolved in water. Barium and yttrium carriers
are added. Precipitation of barium chromate at pH 5.5 removes 140 Ba and natural
radium from the supernatant strontium solution (for counting, if needed, of these
two separated radioelements). Precipitation of yttrium hydroxide in basic solution
then removes the 90 Y daughter that has grown into the 90 Sr parent. Ammonium
oxalate is immediately added to the supernatant solution to precipitate strontium
oxalate. The precipitate is washed and dried in the filter holder with alcohol and
ether, promptly weighed for yield determination, and counted with a beta-particle
detector (Chieco 1997) such as a proportional detector.
The measured count rate must be corrected for contribution by 90 Y daughter
ingrowth that begins immediately after the hydroxide scavenging precipitation. If
time is not of the essence, the measurement can be delayed for 20 days until 90 Y
ingrowth is essentially complete, to reach equal disintegration rates for 90 Sr and
90
Y. Individual counting efficiencies must be applied for 90 Sr and 90 Y because the
two are not identical in the proportional counter.
An alternative is to store the supernatant solution after the barium chromate
scavenging precipitation for 20 days until 90 Y approaches equilibrium with 90 Sr,
and then precipitate yttrium hydroxide with the 90 Y. The hydroxide is dissolved in
dilute acid, and oxalic acid is added to precipitate yttrium oxalate. The precipitate
is washed and dried, promptly weighed for yield determination, and counted with a
proportional detector (Sunderman and Townley 1960). To the supernatant solution
from the yttrium hydroxide precipitation, ammonium oxalate is added to precipitate
strontium oxalate for weighing to determine the strontium yield. The 90 Sr activity
is calculated from the 90 Y count rate adjusted for 90 Y ingrowth into the 90 Sr parent,
90
Y decay since its separation from 90 Sr, yttrium and strontium yields, and 90 Y
counting efficiency.
One variant is that the radionuclides are dissolved and then measured with an LS
counter. Another variant is that 85 Sr tracer is used for strontium yield measurement
when 90 Y is measured to determine the 90 Sr activity.
Samples with high calcium content can be problematic because some calcium
nitrate may also precipitate in nitric acid at the indicated strength. The additional
calcium will cause overestimation in gravimetric strontium yield. The calcium
nitrate precipitate (after thorough draining to remove the explosion potential of
the nitric acid–alcohol mixture) can be dissolved in a few milliliters of absolute
alcohol with little loss of strontium nitrate precipitate. A possibly safer removal of
calcium contamination is by reducing the nitric acid strength to 60% or dissolving
the strontium and calcium nitrate precipitates in a small amount of water and then
precipitating strontium nitrate with concentrated nitric acid for a second or even
6. Applied Radioanalytical Chemistry 107
third time. One interference is a silica precipitate in the strong nitric acid solution
from samples such as dissolved soil; this precipitate can be removed by filtration
after dissolving the strontium nitrate in water.
As indicated in Section 6.4.2 below, the presence of 89 Sr requires calculation
by simultaneous equations to determine the disintegration rates of 89 Sr and 90 Sr in
the same sample. Both radionuclides can be expected in a mixture of shorter-lived
fission products within a year after formation.
Changes in 90 Sr purification were developed mainly to eliminate use of large
volumes of fuming or concentrated nitric acid that corrode hoods, ducts, and lab-
oratory equipment. Early modifications applied selective ion-exchange sorption
to separate strontium from calcium. For example, calcium in the sample is com-
plexed by EDTA at pH 4.6 so that strontium, barium, and radium can be retained
on a strong-base cation-exchange resin while calcium passes through the column
(Porter et al. 1967). The strontium is then selectively eluted with the same so-
lution at pH 5.1 and precipitated as the carbonate for yield determination and
counting.
A more recent technique uses the Eichrom column discussed in Section 3.5. After
initial precipitation and dissolution to reduce the solution volume, the solution is
adjusted to the specified acidity and passed through the column, which retains
strontium. Strontium is eluted and then precipitated as the carbonate.
Tc-99 Analysis
Technetium does not have a stable isotope. Of its long-lived isotopes, 95 Tc, 97 Tc,
98
Tc, and 99 Tc, only the latter is a fission product. Short-lived 99m Tc (6.0 h) pre-
cedes 99 Tc. It is the daughter of 99 Mo (66 h), a radionuclide widely used in nu-
clear medicine. The half-lives of all other technetium fission products are less
than 1 h.
The most stable oxidation states of technetium are +4 and +7. Hydrazine and
hydroxylamine reduce technetium to +4. Atmospheric oxygen, hydrogen perox-
ide, and strong nitric acid oxidize technetium to +7. Some reactions can be slow.
Other, less stable, oxidation states exist. In the lower oxidation state, technetium
as TcO2 is carried on ferric hydroxide precipitate. In the upper oxidation state,
technetium is the anion TcO− 4 . It is sorbed on anion-exchange resins, extracted
into oxygen-containing solvents such as hexone (methyl isobutyl ketone) and trib-
utylphosphate, and coprecipitated with rhenium as the phenylarsonium pertech-
netate, (C6 H5 )4 AsTcO4 . It can be distilled as Tc2 O7 from sulfuric and perchloric
acids (Anders 1960).
The above-cited reactions are used to separate 99 Tc for measurement. Rhenium is
a common nonisotopic carrier for precipitation or yield measurement. The shorter-
lived 95 Tc (60 days) has been used as isotopic tracer for yield measurement (Anders
1960). Technetium can be separated from rhenium by coprecipitating the +4 state
with ferric hydroxide while rhenium remains in solution. Technetium is separated
from the fission-produced radionuclides 103 Ru and 106 Ru by distillation with Re2 O7
from sulfuric acid while ruthenium remains behind because it is volatile only in its
108 Bernd Kahn, Robert Rosson, and Liz Thompson
I-129 Analysis
If a thermal neutron source is available, 129 I can be determined at low concentra-
tion by measuring activation-produced 130 I (12.4 h); at the same time, the specific
activity can be measured by activation of stable 127 I to 128 I (25 min). The cross-
sections for neutron activation are 6.1 barn for 127 I and 9 barn for 129 I. Although
129
I is a fission product, it is also produced in the environment by cosmic ray neu-
trons. Attribution to an anthropomorphic source must be checked by background
measurements of comparable matrices in terms of the specific activity. Data from
such measurements are available (NCRP 1983).
At a magnitude of 0.1 Bq per sample, 129 I can be measured directly with an LS
counter or a Ge gamma-ray spectrometer in its low-energy region. The radionuclide
emits 0.153 MeV (100%) maximum energy beta particles, 0.040 MeV (7.5%)
gamma rays, K X-rays in the 0.029–0.034 MeV (70%) region, and conversion and
Auger electrons in the 0.025–0.039 MeV (22%) region.
The radionuclide with added carrier can be separated from most other long-
lived radionuclides by precipitating silver or palladium iodide from dilute nitric
acid solution. As indicated in Section 6.3.1, the oxidation state of radioiodine,
of which there are several, must be well defined before a separation step can be
trusted. Other purification techniques are solvent extraction of iodine oxidized to
I2 into carbon tetrachloride followed by back extraction of the reduced iodide form
into water, or sorption of I− on anion-exchange resin followed by elution with a
strong chloride solution (Kleinberg and Cowan 1960). These processes also lend
themselves to concentrating the radionuclide from larger solution volumes to attain
a lower detection limit.
Cs-137 Analysis
Initially, 137 Cs was analyzed by adding cesium carrier and precipitating a com-
pound that is specific for cesium in dilute acid solution in the presence of sodium,
such as the cobaltinitrite, phosphomolybdate, silicotungstate, or tetraphenylbo-
rate. Known contaminants, such as 32 P or 99 Mo in phosphomolybdate, are then
removed after dissolving this precipitate by a second precipitation with a different
reagent or by a scavenging precipitation for the contaminant. The purified sample
is weighed for yield and then counted with a proportional or G-M counter (Finston
and Kinsley 1961).
137
Cs is collected from larger volumes of water on large cation-exchange resin
columns. Anion-exchange resin columns loaded with one of the above-cited pre-
cipitation reagents make these columns specific for cesium. A column loaded with
6. Applied Radioanalytical Chemistry 109
Pm-147 Analysis
147
Pm is one of three relatively long-lived rare earth fission products; the other
two are 144 Ce (284 days), which is discussed in the following section, and 155 Eu,
with a low fission yield. 147 Pm emits a 0.284-MeV maximum beta-particle energy
group and no gamma rays. In contrast, 144 Ce emits gamma rays of 0.134 MeV
(10.8%) and 155 Eu emits gamma rays of 0.087 MeV (31%) and 0.105 MeV (20%).
Contamination by 152 Eu and 154 Eu is possible if a sample that contains trace
amounts of europium, such as soil, was irradiated with neutrons. Both of these
radionuclides emit gamma rays.
The typical purification method for rare earths is coprecipitation with ferric
hydroxide, dissolution in dilute acid, precipitation as fluoride in strong mineral
acid solution, dissolution in strong nitric acid with boric acid to complex fluo-
ride, and precipitation for counting as the oxalate in dilute acid solution (Steven-
son and Nervik 1961). Because 147 Pm has no stable isotope, another rare earth
(such as lanthanum) is added as carrier. The 147 Pm precipitate can be counted
with a proportional counter, or can be dissolved and measured with an LS counter
because of the low beta-particle energy. If small amounts of the other rare earth
radionuclides are detected by gamma-ray spectrometric analysis, the beta-particle
count rate of 147 Pm can be calculated by difference.
147
Pm can be separated from cerium and europium because it remains in the +3
oxidation state when cerium is oxidized to +4 or europium is reduced to +2. 144 Ce
can be scavenged in dilute acid solution by adding cerium carrier, oxidizing to the
+4 form with sodium bromate, and then precipitating it with HIO3 as Ce(IO3 )4 .
Eu-155 can be separated by adding holdback europium carrier, reducing europium
to the +2 state with zinc powder, and then carrying 147 Pm on ferric hydroxide
precipitate (Stevenson and Nervik 1961).
Rare earths can be separated from each other on ion-exchange columns, as
illustrated in Fig. 3.2. Once separation conditions have been defined, the method
can be simplified for purifying one radionuclide such as 147 Pm from interfering
rare earth radionuclides.
pCi/l (0.037 Bq/l). Radium analysis may also be needed in soils, ores, manufactured
materials such as concrete, and various contaminated media.
The conventional analysis of 226 Ra in water is coprecipitation with barium sulfate
carrier from a 1-l volume. After decanting the supernatant solution, barium sulfate
is dissolved in EDTA solution and stored in a sealed container for 1–4 weeks to
allow for ingrowth of its 222 Rn (3.82 days) daughter. The 222 Rn gas is flushed with
air (sufficiently aged in a tank to assure the absence of radon) into an evacuated 100-
to 200-cc ZnS(Ag)-coated “Lucas” cell. A photomultiplier tube views scintillations
at the ZnS(Ag) surface through a light pipe at one end of the cell (Lucas 1957). The
measured alpha particles are emitted by 222 Rn and its progeny 218 Po and 214 Po. The
ingrowth of 214 Po, controlled by its precursors 214 Pb and 214 Bi, reaches equilibrium
with that of 222 Rn within 4 h.
The activity of 226 Ra is calculated from the count rate of the three alpha particles
of its progeny. Adjustments are made for the ingrowth and decay of 222 Rn, the
detector counting efficiency, and any losses during the precipitation of barium
sulfate and the transfer of the gas into the cell.
The analysis of 228 Ra conventionally followed that of 226 Ra in precipitating with
barium sulfate and dissolving with EDTA solution. After ingrowth of 228 Ac (6.13 h)
for 2 days, the 228 Ac is precipitated with yttrium carrier as yttrium oxalate. The
precipitate is washed and dried with alcohol and ether, weighed, and then counted
with a proportional detector. The barium that remained in the EDTA solution again
is precipitated as sulfate to determine its yield. The 228 Ra activity is calculated
from the 228 Ac count rate and counting efficiency, the yields of barium and yttrium
carriers, and the ingrowth and decay of 228 Ac (Percival and Martin 1974).
Various other methods of radium purification by coprecipitation, ion exchange,
and radon emanation are available (Kirby and Salutsky 1964). In a recent method,
both 226 Ra and 228 Ra are collected by a barium sulfate precipitate which is weighed
to determine the barium yield, and then counted by Ge gamma-ray spectral analysis.
The counted gamma rays are emitted by the 214 Pb and 214 Bi progeny of 226 Ra
and the 228 Ac progeny of 228 Ra. The measurement for 228 Ra can be performed
immediately because 228 Ac coprecipitates with barium sulfate, but delay by 1–4
weeks is needed for ingrowth of the 222 Rn daughter of 226 Ra in the barium sulfate
(Kahn et al. 1990). Corrections are required for yield and incomplete ingrowth to
calculate the activity.
Uranium Analysis
The uranium isotopes often encountered in the radioanalytical chemistry laboratory
are listed in Table 6.2. In natural uranium, the relative decay rates at equilibrium
are 1.0 Bq 238 U, 1.0 Bq 234 U, and 0.045 Bq 235 U. Enriched (containing relatively
more 235 U and 234 U) and depleted (relatively more 238 U) combinations are also
encountered, as are 236 U in neutron-irradiated mixtures and 233 U from some pro-
cesses. These uranium isotopes emit alpha particles, characteristic 13-keV L X
rays, and generally several weak gamma rays. Several isotopes have numerous
minor alpha-particle or gamma-ray transitions that are not listed.
6. Applied Radioanalytical Chemistry 111
C1284-94 listed in Appendix A-1. Other actinides that emit alpha particles are
also prepared with this method. After purification, uranium isotopes usually are
identified and measured by alpha-particle spectrometry with silicon diodes. Tracer
232
U can be added to the initial solution and counted with the uranium isotopes of
interest.
Table 6.2 indicates some uranium isotopes that emit gamma rays in a significant
fraction of decays; these can be measured directly by gamma-ray spectral analysis.
In measuring 235 U by its most intense gamma ray of 0.186 MeV, interference from
226
Ra is possible. For 238 U, the listed gamma rays are emitted by the daughter
234
Th (24.1 days) and its daughter 234 Pa (1.2 min) if they are at equilibrium.
Plutonium Analysis
Plutonium isotopes of interest are listed in Table 6.3. All except 241 Pu emit alpha
particles, 14-keV L X rays, and multiple weak gamma rays; 241 Pu emits beta par-
ticles that decay to 241 Am, which emits alpha particles and gamma rays. Single or
multiple neutron reactions with 238 U and 235 U form all of the isotopes. The isotopes
that emit alpha particles have numerous unlisted minor gamma-ray transitions.
Among plutonium oxidation states from +3 to +6, the most stable forms are
+3 and +4. Conversion between oxidation states is used for purification from
other radionuclides. Plutonium is oxidized to the +4 state by hydrogen peroxide,
permanganate, and nitrite, and reduced to the +3 state by bisulfite and ascorbic
acid. A strongly acidic or complexing solution is needed to maintain the selected
state and avoid hydrolysis with polymerization (Coleman 1965).
An early conventional method for plutonium analysis was coprecipitation in
acid solution with a rare earth fluoride, dissolution in aluminum nitrate solution
with sodium nitrite to maintain the +4 oxidation state, and extraction into thenoyl-
trifluoroacetone (TTA) in benzene. Plutonium was back-extracted into dilute HCl,
the acid was evaporated, and plutonium was taken up in HCl–NH4 Cl solution
Am-241 Analysis
The long-lived isotopes of americium are 241 Am (458 years), 242 Am (152 years),
and 243 Am (7400 years). All are formed by multiple neutron interactions with
uranium and plutonium. Of particular interest is 241 Am because it identifies the
erstwhile presence of its 241 Pu parent. 241 Am emits alpha particles but can also
be measured by its 0.059-MeV (36%) gamma ray. 242 Am has several isomers that
mostly emit beta particles, and 243 Am emits alpha particles.
Americium in aqueous solution is in the +3, +5, and +6 states, respectively,
as Am+3 , AmO+1 +2
2 , and AmO2 . The trivalent state is most common, but higher
oxidation states are achieved by strong oxidation. The highest oxidation states can
be used for separating americium from curium and rare earth radionuclides with
ammonium persulfate, (NH4 )2 S2 O8 (Penneman and Keenan 1960).
The trivalent actinides such as 241 Am follow the same precipitation reactions
as the trivalent rare earth radionuclides, notably with insoluble hydroxides, flu-
orides, and oxalates. Numerous solvent extraction and ion-exchange separations
from other trivalent radionuclides are reported. Americium radionuclides can be
114 Bernd Kahn, Robert Rosson, and Liz Thompson
separated from other rare earth and actinide radionuclides by meticulous control
of ion-exchange column elution.
A convenient tracer for 241 Am is 243 Am (if the latter is not in the sample).
Both conventionally are measured with a silicon diode by alpha-particle spectral
analysis. The disintegration rate is calculated as discussed above for plutonium.
the mix is purely natural, or consists of fission products; certain activation products
may also be identified. In principle, when the relative amounts of various fission
products resemble the distribution for one of the fissionable isotopes (see Fig. 2.2),
instantaneous fission can be inferred. The absence of shorter-lived fission product
suggests that a considerable interval between production and measurement has
occurred.
In practice, the origin of a radionuclide mixture usually is more difficult to
identify because of ongoing radioactive decay between origin and sample, differing
degrees of retention at the origin, and partial retention along the pathway. At a
nuclear reactor, for example, fuel element cladding retains most, but not all, of the
fission products, and reactor coolant controls and waste processing retain more.
Many of activation products that may originate from nuclear reactor or nuclear
medicine facilities can be predicted from experience, but not necessarily associated
with a specific facility.
Radioisotope multiples, e.g., 89/90 Sr, 103/106 Ru, 141/143/144 Ce, 134/136/137 Cs,
133/135
Xe, 235/238 U, or 238/239 Pu, are particularly useful for attributing origin and
inferring the time from formation to sampling. With the 89/90 Sr pair, for example, if
the generation rate of 89 Sr atoms relative to 90 Sr atoms were the same, the 89 Sr:90 Sr
activity ratio is inversely proportional to the half-lives, i.e., 208. The ratio is about
170 because of the higher fission yield (the relative atom generation rate) of 90 Sr
compared to 89 Sr. After accumulation for 500 days, the ratio is reduced to about
25 because 89 Sr no longer accumulates while 90 Sr still is far from equilibrium.
Even here, conditions usually are complex because various fuel elements may be
exposed for different periods and the two radioisotopes may have been released at
different times.
More confident attribution usually is possible when radionuclides in a sample
can have only limited origin from, say, atmospheric fallout and one or two nuclear
facilities. For tritium, 14 C, 129 I, or uranium in the environment, for example, the
specific activity (radioisotope/stable isotope ratio) can indicate the origin. Cer-
tain activation products can be attributed to specific nuclear medicine or reactor
facilities.
Screening measurements for gross alpha and beta particles may not be partic-
ularly useful for fission-product mixtures because generally alpha particles are at
extremely low concentration and beta particles are contributed by numerous ra-
dionuclides. A gross screening measurement is useful for suggesting additional
analyses if it clearly exceeds the sum of the radionuclides measured individually.
89
Sr-90 Sr Analysis
Determination of 89 Sr together with the long-lived 90 Sr (see Section 6.4.1) is a
widespread analytical endeavor for radiation protection because of the similarity
of strontium to calcium in their absorption into bone mass. Strontium separation
is performed as described in Section 6.4.1 whether or not 89 Sr is in the sample.
The distinction between treating a sample that is known to have no 89 Sr and one
116 Bernd Kahn, Robert Rosson, and Liz Thompson
instead of 2–4 weeks is feasible because 90 Y can be calculated for any time period,
but reduces the count rate, and hence the precision. In distinction, if results need
not be reported promptly, then the first measurements can be performed 3 weeks
after yttrium separation, when the 90 Sr/90 Y pair is in equilibrium, and the second
measurement, after 7 weeks when 89 Sr has decayed by about one half-life. The
difference in the count rate then is almost entirely due to 89 Sr decay.
The two radioisotopes of strontium and 90 Y can also be distinguished by energy
discrimination with absorber foils in a proportional counter or by spectrometer in
an LS counter. These methods usually are less precise than the two above-cited
alternatives.
The other cited radioisotope pairs among fission products are measured simul-
taneously by gamma-ray spectral analysis. Before such analysis became available
in the 1950s, these pairs were distinguished after chemical separation by dual mea-
surements such as those described above for 89/90 Sr before and after an interval
during which the shorter-lived radionuclides had decayed sufficiently to measure
the difference with the needed precision.
Numerous techniques have been reported (Kusaka and Meinke 1961). Fundamen-
tally, all processes for sample preparation, purification, and measurement must
be brief and highly effective. Sample purification before production is helpful.
Miniature ion exchange, solvent extraction, and precipitation systems have been
developed for processing small volumes. Mechanical transfer from purification to
the counting system enhances prompt measurement. Measurement results must
take into account the radioactive decay of the radionuclide during measurement
shown in Eq. (10.2).
Radionuclides with intermediate, short, and very short half-lives include progeny
of long-lived terrestrial radionuclides and products formed by cosmic-ray irradia-
tion of gases in the upper stratosphere. All three terrestrial decay chains—starting
from 238 U, 235 U, and 232 Th—include radionuclides with half-lives of seconds and
minutes that are readily detected because they are supported by their long-lived
precursors. Prompt analysis after collection is needed to analyze very short-lived
cosmic-ray-produced radionuclides such as the isotopes of chlorine, 34m Cl, 38 Cl,
and 39 Cl, with half-lives of minutes.
For radionuclides in the very long-lived to extremely long-lived category, the
primary consideration is that a longer half-life is associated with the lower prob-
ability of decay per atom and lesser decay energy. As a consequence, analysis
for such radionuclides requires large samples, thorough purification, and mini-
mal radiation background. Moreover, energetic gamma rays are not emitted. Mass
spectrometric separation and analysis (see Chapter 17) generally is the preferred
alternative.
committee. This committee has over 500 professionals from the ranks of the Amer-
ican Water Works Association, the American Public Health Association (APHA),
and the Water Environment Federation. Their combined effort produces the Stan-
dard Methods text; part 7000 of the text deals with the testing of water samples
for radionuclides. Appendix A-2 lists the Standard Methods, part 7000 proce-
dures. The APHA intersociety committee published Methods of Air Sampling and
Analysis (2nd ed.) in 1977.
ANSI is the U.S. representative of ISO. Appendices A-3 and A-4 list ANSI and
ISO procedures, respectively.
The benefit of a standard method is the scrutiny and evaluation that it has received
by fellow professionals. For that reason, a regulatory agency or customer may
require application of specified standard methods in the radioanalytical chemistry
laboratory. If another method is preferred because of its greater flexibility, rapidity,
or ease in application, it may be compared initially or periodically with a standard
method to confirm its reliability. Development and approval of these standards
takes much time, effort, and money. For that reason, the standards are available
from these agencies for a fee, generally ranging from $50 to $75.
7.1. Introduction
The transition from radiochemical separation described in Chapter 6 to instrumen-
tal radiation detection in Chapter 8 is source preparation for counting. The analyst
wants to prepare a source that represents the radionuclide in the collected sample,
can be measured reliably by its radiation, and is stable. The analyst selects a de-
tector that is sensitive to the radiation that characterizes the radionuclide, stable as
defined by its QA program, and calibrated for efficiency and—if needed—energy.
Source preparation concerns are addressed here for the four types of detectors that
are described in Chapters 2 and 8, but these considerations can apply to sources
prepared for measurement by other detectors.
121
122 Bernd Kahn
The carrier weight generally is in the range of 10–30 mg and the yield is specified
to be sufficiently high—typically above 50%—to reduce the uncertainty associated
with yield determination to about 1%. To maintain constant weight and counting
efficiency, the precipitate should not be volatile, hygroscopic, flaky, or powdery.
Any drying or heating during source preparation must not affect comparison with
initial carrier weight due to a change in chemical form or differences in degree of
hydration.
Sources are also prepared by pouring a liquid or slurry onto a planchet and then
evaporating it to dryness. Care must be taken to avoid nonuniform distribution
because solids tend to dry as rings and to accumulate near the lip of the planchet.
Adding a drop of a solvent such as alcohol that reduces surface tension in water
before the onset of dryness can reduce such uneven deposition (Chieco 1997).
Planchets have been designed with concentric ridges to distribute evaporated solids
more evenly. Evaporation on a level surface prevents lopsided accumulation of
solids. Evaporation must be sufficiently slow by controlling heat lamp distance to
avoid sample loss by spattering.
Electrodeposition on metal disks is used to prepare very thin sources of the
radionuclides discussed in Section 3.7. At very low concentrations in water,
certain radionuclides may be deposited on metal disks by sorption or sponta-
neous electrodeposition (Blanchard et al. 1960). The former yields a relatively
low percent deposition on more noble metals, as in the case of 144 Ce deposited
on gold. In contrast, spontaneous electrodepostion results in near 100% yield
on a less noble metal, for example, in depositing 110m Ag or divalent 59 Fe on
magnesium. Yield must be determined separately for this source preparation
step and then combined with the yield observed for the chemical purification
steps.
In the equation, A is the source activity in Bq, is the alpha particle range
in mg/cm2 , a is the source area in cm2 , m is the source mass in mg, and x is
the source thickness in mg/cm2 . The equation indicates that, for samples with a
constant activity per unit mass, the surface flux decreases linearly with thickness
to one-fourth the disintegration rate when the source thickness equals the range,
which is about 7 mg/cm2 for the 5.5-MeV alpha particle emitted by 241 Am. Beyond
this thickness, the surface flux is predicted to remain constant:
Aa
Rs = for x ≥ (7.1b)
4m
Equation (7.1b) shows that the counting efficiency varies inversely with the sam-
ple mass when the thickness exceeds the alpha-particle range, i.e., at “infinite
thickness.”
The counting efficiency, discussed in Section 8.2.1, is related to the flux at
the surface in terms of the source-detector geometry and attenuation of alpha
particles between source surface and the sensitive-detector volume. Table 8.3 pre-
dicts a 31% counting efficiency for the detector-source configuration shown in
Fig. 8.4 (but with the ring-and-disk replaced by the 20-cm2 planchet source).
In practice, the counting efficiency is considerably lower than 31%, as indi-
cated by the measured alpha-particle counting efficiency curve shown in Fig.
7.1. The reduction in efficiency even at 0 mg/cm2 is due to alpha particles that
pass through the source, air, and window at angles not perpendicular to the de-
tector for distances that exceed their range. The curve in Fig. 7.1 is approxi-
mately a straight line from 0 to 60 mg (i.e., 3 mg/cm2 ) and then curves, as the
count rate becomes constant at about 130 mg, the alpha-particle range in the
source.
The decrease in counting efficiency with thickness in Fig. 7.1 emphasizes the
benefit of a thin source for counting alpha particles at low levels. Thin sources can
be prepared by using only a small amount of carrier and a light precipitating reagent,
with purification processes other than precipitation such as electrodeposition, ion
exchange, and solvent extraction, and by flaming the source to remove organic
compounds.
For beta particles, reasonable estimates of the curve ε/ε0 for counting efficiency
as function of mass have been based on their approximately exponential attenuation
7. Preparation for Sample Measurement 125
35%
Geometric efficiency
30%
Alpha (Am-241)
20%
Efficiency
15%
10%
5%
0%
0 20 40 60 80 100 120 140 160
Weight (mg)
FIGURE 7.1. Alpha- and beta-particle efficiency self-absorption curves for 241Am and 137 Cs,
respectively, on 5-cm-diameter planchet measured with proportional counter.
ε0 (1 − e−μ x )
ε= (7.2)
μx
0.60
0.50
Counting efficiency
0.40 Yttrium-90
Strontium-89
0.30 Strontium-90
Carbon-14
0.20
0.10
0.00
0 10 20 30 40
−2
Sample thickness (mg SrCO3 cm )
FIGURE 7.2. Beta-particle efficiency self-absorption curves for 89 Sr, 90 Sr, 90 Y, and 14 C as
1.6-cm-diameter ring-and-disk source, calculated by Monte Carlo simulation. (from Nichols
2006).
For the first three applications, a radionuclide- and mass-specific counting ef-
ficiency must be selected. For the fourth application, a thin sample—below
2.5 mg/cm2 for alpha-particle counting—should be prepared so that efficiency
values are similar at commonly encountered energies. For counting beta particles,
the sample should not exceed 10 mg/cm2 . An intermediate-energy (e.g., 0.6–0.8
MeV β max ) radionuclide standard provides reasonable efficiency estimates except
that the activity of a radionuclide that emits only low-energy beta particles will be
underestimated.
An aliquot of the sample as liquid or slurry usually is dried on a planchet and
then counted for gross activity measurement. In some instances, the radionuclides
are coprecipitated by a carrier such as ferric hydroxide or manganese dioxide that
is expected to collect many of them (ICRU 1972), and this solid is filtered or
poured onto the planchet as slurry and then dried. Under favorable circumstance,
7. Preparation for Sample Measurement 127
the radionuclide concentration is high enough that a thin sample can be prepared
and attenuation by the sample itself is not a major concern. A thicker sample
may yield detectable beta-particle counts in the available time period from a less
radioactive sample, but the self-absorption fraction is in doubt if the radionuclides
are not identified.
Gross alpha- and beta-activity counting can be a useful screening process. It
should not be used to delineate radioactivity levels without associated specific
radionuclide measurements because of uncertainty about what and how much is
being measured. Moreover, some radionuclides may be lost in processing due to
volatility during evaporation or incomplete carrying during precipitation.
a consistent level of quenching leads to more reliable results than wildly variable
levels.
The main advantage of the LS counter over the gas proportional counter is the
intimate mixing of the radionuclide with the scintillating detection medium to elim-
inate loss of alpha and beta particles by absorption outside the detector. Moreover,
the LS geometrical efficiency is almost 100%. The practical counting efficiency is
above 90% for alpha and more energetic beta particles. Energy loss by quenching
is minor except for low-energy beta particles from radionuclides such as tritium.
One disadvantage of the sample–cocktail mix is impermanence. The stability
of the prepared vial contents should be tested over time. Typically, a sample can
be recounted after a few days or weeks, but should not be considered reliable after
longer periods.
is better than for beta-particle continua because alpha particles are emitted with
discrete energies. Resolution of alpha-particle energy peaks in conventional LS
systems is about 0.50 MeV. It is reduced to about 0.40 MeV by deoxygenating
the sample to reduce quenching. This resolution is far poorer than for solid state
detectors (see Section 7.4), but spectral analysis can be useful for radionuclide
mixtures in which only a few alpha-particle groups occur at widely separated
energies. A photon/electron rejecting alpha liquid scintillation (PERALS) system
has been developed that discriminates against beta particle scintillations by their
longer decay time, improves the resolution to about 0.25 MeV, and reduces the
background to about 0.001 c/m (McDowell 1992).
Solid scintillators such as anthracene have been tested as alternatives for contin-
uous measurements (ICRU 1972). The scintillation material and the radionuclide
collector are combined, as discussed above for scintillation beads (Winn 1993).
The operating characteristics of these devices, such as the sample volume, count
accumulation period, energy resolution, detection efficiency, and background must
be arranged to satisfy radionuclide concentration limit specifications.
The LS counter also measures Cherenkov radiation (see Section 2.4.3) in water
without addition of scintillation cocktail. The radiation is generated in water by
electrons at energies above 0.265 MeV. Hence, about one-half of beta particles in a
group with E max of 0.8 MeV generate this light, and the fraction is larger for groups
with higher maximum beta-particle energies. The advantages of this process for
detecting beta particles are that the water sample volume can be the full 20 ml and
the sample is stable. The drawbacks are that quenching still can occur, few photons
are produced per beta particle, and the PMT usually is not optimal for detecting
the wavelength of the Cherenkov radiation. A 1-MeV beta particle produces about
200 Cherenkov scintillations in water compared to 10,000 scintillations in an LS
cocktail (Knoll 1989).
7.5.3. Calibration
Source containers must be sufficiently rigid to maintain their shape. They must
resist attack and decomposition from contents such as organic solvents, acids,
bases, and biological material. They must be closed to prevent spillage, permit
storage, and protect the detector and its environment from contaminants. In some
7. Preparation for Sample Measurement 133
instances, samples are sealed to prevent inleakage of air or escape of gases. One
technique is to place a solid or liquid sample in a metal can and then seal its
lid (Chieco 1997). Gaseous samples are transferred to an evacuated container
connected by tubing, valves, and a pump with the sample container.
Efficiency calibration requires a sample of defined shape and volume placed
at a defined location relative to the detector. A sample frame may be needed for
reproducible placement. Calibration can be performed with a radioactivity standard
or by Monte Carlo simulation (see Sections 8.2 and 10.5).
Although gamma rays are much less subject to attenuation than alpha and beta
particles, a density correction is needed if the density of the sample deviates sig-
nificantly from the density of the calibration standards. The effect of density on
self-absorption for both the standard and the sample is estimated by Eq. (7.2); μ
for this purpose is the photon attenuation coefficient in cm2 /g and x is the sample
area density in g/cm2 . Values for μ in some common materials are listed in Table
2.2 and in its cited reference. If a large set of samples with consistent density is
analyzed, it may be possible to prepare radioactivity standards at the same density
to avoid the need for correction. Interpolating efficiency values as a function of
density is feasible at energies above 0.1 MeV because the effect of minor density
difference on counting efficiency is small.
When initial nonuniformity cannot be eliminated by mixing the sample, e.g., in
whole-body counting, the radionuclide distribution can be estimated by calibrating
with standards in various configurations and conditions. The results can establish
a range of possible counting efficiencies. Changing sample location relative to the
detector for separate counts may yield further information concerning radionuclide
distribution. Monte Carlo simulations can also provide needed efficiencies.
A simple example of sample nonuniformity is an air cartridge that may have
retained a radionuclide either on the front surface or with decreasing concentration
in depth, or distributed uniformly throughout. Counting the cartridge from both
sides and comparing the results to a simulation model will permit calculating the
activity as a function of depth distribution.
8
Applied Radiation Measurements
JOHN M. KELLER
8.1. Introduction
The radiation detection systems employed in radioanalytical chemistry labora-
tories have changed considerably over the past sixty years, with significant im-
provement realized since the early 1980s. Advancements in the areas of material
science, electronics, and computer technology have contributed to the develop-
ment of more sensitive, reliable, and user-friendly laboratory instruments. The
four primary radiation measurement systems considered to be necessary for the
modern radionuclide measurement laboratory are gas-flow proportional counters,
liquid scintillation (LS) counters, Si alpha-particle spectrometer systems, and Ge
gamma-ray spectrometer systems. These four systems are the tools used to identify
and measure most forms of nuclear radiation.
Some operating parameters can be considered in common for all these detec-
tors, notably detection efficiency and the radiation background. Additional param-
eters pertain only to detectors with associated spectrometers. These parameters
concern the radiation energy peaks that identify and quantify radionuclides, and
include energy calibration, energy resolution, peak-to-Compton ratio, and peak
shape.
Each of the detection instruments has a specialized field of application. The gas-
flow proportional counter is the laboratory workhorse for measuring radionuclides
that emit alpha and beta particles in samples before radionuclide separation and
then again in the purified fractions. The LS counter also measures radionuclides that
emit alpha and beta particles and is particularly useful for measuring low-energy
beta particles. The alpha-particle spectrometer is applied for purified samples to
measure radioisotopes, notably of the heavy elements, that emit alpha particles. The
gamma-ray spectrometer is the laborsaving device for measuring photon-emitting
radionuclides, in most cases without chemical separation.
Many ingenious applications of these detectors go beyond such specialization
so that their use can be a matter of matching sample characteristics, radiation
types, the radionuclide of interest and associated impurities, required sensitivity,
and convenience in sample preparation, or even analyst preference. Various other
detection systems and instrumental techniques that are mentioned at the end of
134
8. Applied Radiation Measurements 135
this chapter should be considered when available. Some of these other detectors
may be more appropriate than the four conventional ones for analyzing samples
with very high or low activity and for radionuclides that otherwise would require
complex radiochemical separations prior to analysis.
A careful analyst will apply more than one of these techniques to assure reliable
calibration. The first three will be discussed here to review the basic elements of
counting. The fourth approach is a numerical analysis technique that should be
given proper grounding in an appropriate text (see Briesmeister 1990) and proper
introduction to the student after the basics of counting are understood.
The crucial characteristic that applies to all detector calibration is the practical
counting efficiency—the ratio of the observed measurement to the activity of the
sample in Bq (or pCi) per sample that may be converted to activity per unit mass
(or volume or flow rate). In the simplest case, a measure of the practical counting
efficiency ε is given by the relationship
RG = Aε + RB (8.1)
where A is the disintegration rate (or activity), RG is the observed count rate of the
sample, and ε is the counting efficiency. A measured background count rate RB
must be subtracted from the observed count rate to calculate the actual count rate
of a sample. The observed count rate is also referred to as the gross count rate and
the value after subtracting the background, the net count rate. Most measurements
of the count rate of a sample are performed to determine its disintegration rate, but
some measurements are performed only to compare count rates, e.g., in half-life
measurements or for samples taken to evaluate a purification step.
In the first two calibration approaches cited above, a certified radionuclide
standard is measured to obtain the practical counting efficiency of the detector,
136 John M. Keller
TABLE 8.1. Typical values of radiation detector counting efficiency and background
Sensitive
Detector volume Radiation Energy Efficiency Background
type (cm3 ) type (keV) (%) (c/m) Conditions
Proportional 28 alpha 4000–9000 20 0.05 Anticoincidence
beta 400–3500 45 1
LS 20 alpha 4000–9000 >95
20 beta 18 25 2 10 ml H2 O
20 beta 250–3500 >90 5 20
LS 1 alpha 4000–9000 99.7 0.001 PERALS
Si spec 0.06 alpha 4000–9000 25 0.0001 Electrodeposited
source
Ge spec 85 gamma 100–103 0.15 0.5 Flat source
500–505 0.063 0.3 Flat source
1000–1006 0.035 0.15 Flat source
2-000–2007 0.011 0.011 Flat source
Here d is the distance from the point source to the detector face and ρ d is the
radius of the detector window. The solid angle approaches 2π when d approaches
0, i.e., approaches 0.5 for a point source near the detector surface. Such opti-
mized geometry is desirable in low-level measurements, but not necessary under
other circumstances. Some geometry factors for a circular source with radius b,
calculated as a function of d/ρ d and b/ρ d (Bland 1984), are given in Table 8.3.
Reducing the sample mass and the amount of materials between the sample and
the detector reduces radiation attenuation l. This effort is most important for alpha
particles and least important for energetic gamma rays, as discussed in Chapter 7.
The fractional self-absorption within the sample can be estimated for beta particles
and gamma rays by using Eq. (7.2). The fractional attenuation of gamma rays in
8. Applied Radiation Measurements 139
and self-absorption effect in G-M counters also differ from those in end-window
proportional counters because of the typically larger sample-to-detector distances
and smaller detector window (Steinberg 1962). Scattered energetic gamma rays do
not affect the efficiency in gamma ray spectrometers because they only increase
the Compton-scattering region. Material between source and detector also scatters
direct radiation away from the detector, but this effect is small when, as is common,
the amount of intervening material is small.
Source radiation that produces relatively weak pulses is lost (to the extent of frac-
tion 1 − p) when the pulse processing system of a detector discriminates against
the low-energy pulses that constitute noise; discriminators are, however, set to min-
imize count loss. Such discrimination prevents large Ge detectors from measuring
low-energy (<30 keV or <3 keV, depending on construction) X rays, reduces the
counting efficiency of LS counters in measuring low-energy radionuclides such as
tritium, and causes a loss of <1% in proportional counters. Measurements with a
pulse generator or recording low-energy electron or X-ray pulses can identify the
energy cutoff to evaluate whether the discriminator setting is acceptable.
The factor p also refers to loss of counts at all energies if radiation arrives at
the detector more rapidly than individual pulses can be processed. The dead time
(τ ) is the detector recovery period before it is capable of recording a second event.
The dead time depends on the period and shape of the pulse and is an inherent
property of the detecting systems (e.g., gas, liquid, or solid state detectors) and
their electronic support. If the count rate is sufficiently low, the average interval
between pulses is sufficiently long that the dead time is insignificant. For higher
count rates, the following “nonparalyzable” model for correction can be applied:
Robs
Rcor = (8.6)
(1 − Robs τ )
Rcor is the corrected count rate and Robs is the observed rate. Count rates should
be limited so that the correction does not add more than about 30%.
The dead time is long in a G-M counter because of its large and wide pulses,
but short in a proportional counter. For a G-M detector with a dead time of 0.5 μs,
a measured count rate of 200 c/s implies a loss of 11% according to Eq. (8.6). In
contrast, a proportional counter with a dead time of 5 μs has a loss of only 0.1%
at the same measured count rate.
The loss in count rate can be determined by counting a set of samples with
increasing known activity and plotting the relation of count rate to activity. The
extent of deviation from a straight line at higher count rates indicates the loss.
Another technique for determining the dead time compares the count from a split
source when combined with the sum of the separately counted components (Knoll
1989).
The distinction between a second event that only adds to the pulse height of
the first event or is counted separately is a function of pulse height discrimination
and random pulse arrival time, as shown in Fig. 8.1. Further distinctive behavior,
also shown in the figure, is a second event that is counted although the first pulse
has not completely died away, i.e., the second pulse has some contribution from
8. Applied Radiation Measurements 141
both events. The time required for the first pulse to die away completely is termed
the resolving time. The terms “dead time” and “resolving time” are often used
interchangeably.
The impact of such loss depends on pulse shape and pulse application. In integral
counting, all distinct pulses with height above an acceptance value are counted.
In spectral analysis, the pulse height matters. A clock records the live time for
spectrometers as the time interval to use in calculating the count rate. A pulse
that results from two radiations detected within the resolving time is known as a
“coincidence”; the coincidence is recorded at the summed energy and results in two
lost pulses at the energies corresponding to the individual radiations. Coincidences
occur randomly, but can be expected with much greater frequency when counting
radionuclides that emit two radiations simultaneously.
Radiations that are emitted simultaneously and measured for standardization
purposes (see Section 9.3.5) include two gamma rays and beta particles with
gamma rays. Occurrence of coincidences can be inferred from information in
the compilations of radionuclide decay characteristics cited in Chapter 9. The
counting efficiency ε for such coincidences is the product of the efficiency for
each radiation. In simplified terms, the fractional detection of a coincident gamma
ray relative to either of the two gamma rays is ε1 ε 2 /ε 1 = ε 2 , which is significant
only for large values of ε1 and ε 2 .
In recent years, calibration by cascade summing has become a more recognized
measurement technique with the introduction of the larger high-purity germanium
(HPGe) detectors which have a higher intrinsic efficiency. Cascade summing or
true coincidence summing results when two gamma rays are emitted from the
nucleus at nearly the same time and are then detected simultaneously at an energy
that is the sum of the two incident gamma rays. True coincidence summing is
geometry-dependent and should not be confused with the count rate-dependent
random summing mentioned above. The probability of detecting both gamma rays
142 John M. Keller
is proportional to the square of the solid angle subtended by the detector, hence
the sum peak count rate will increase as the source to detector distance decreases.
Coincidence summing can introduce the problem of misattributing summed
gamma rays as single gamma rays when counting sources close to large HPGe de-
tectors. Use of n-type HPGe detectors and Be windows can result in cascade
summing problems at lower energy X rays. Gamma rays can even sum with
Bremsstrahlung radiation from associated beta particles (Gilmore and Hemingway
1995). Cascade summing problems are observed with radionuclides commonly
used as calibration standards such as 60 Co, 134 Cs, and 152 Eu.
The decay fraction f of an emitted radiation and its half-life that are used for
measuring a radionuclide and calculating the extent of decay or ingrowth are tab-
ulated in tables of isotopes and on the Internet at the BNL nuclear data bases
(http://www.nndc.bnl.gov/ index.jsp). The decay scheme should be examined, as
discussed in Section 9.3, before selecting the decay-fraction value. When mea-
suring gamma rays, for example, the gamma-ray fraction must be used instead
of a decay fraction from an excited state that includes conversion electrons. Any
applicable conversion-electron fraction must be included when converting the beta-
particle count rate to activity. Characteristic X rays may be detected when counting
beta particles or gamma rays.
Processing a radionuclide with minimal decay between collection and
measurement—that is, D is almost 1.0—optimizes detection sensitivity and min-
imizes the uncertainty of decay correction. Some radioactive decay is inevitable
and may be useful for reducing interference from shorter-lived radionuclides.
A separation procedure with the yield Y below unity is expected in radioanalyt-
ical chemistry. For quality assurance purposes, the range of acceptable yields may
be set between 0.5 and 1.0. Lower or higher yield fractions suggest occurrence
of analytical process problems. Carrier or tracer addition to determine yield is
described in Section 6.3.
Appearance of the mass (or volume) term, m, in the denominator of Eq. (8.2)
suggests that large samples permit more sensitive measurements. Analyzing ever
larger samples is a tempting solution for attaining lower detection levels, but
larger samples require ever more complex processing and may cause reduction
in factors for the other variables in Eq. (8.3), notably those related to detector-
sample geometry, radiation attenuation, and chemical yield.
1000
511
800
600
Counts
400
352
1461
200
609
1764
0
0 500 1,000 1,500 2,000
Energy (keV)
The extremely low background count rates given in Table 8.1 for two types of
alpha-particle detectors suggest that few, if any, counts will be accumulated during a
background measurement. In that case, statistical considerations in Sections 10.3.1
and 10.4.1 for calculating the uncertainty due to the background and the minimum
detectable activity do not apply. To decide whether a measured value near zero is
due to the background may depend on the presence of a recognizable peak or on
inferences based on experience concerning background fluctuation in the energy
region of interest.
Gamma rays from the thermal neutron capture of cosmic ray neutrons within the
detector and shielding materials are emitted by cadmium in graded shield liners
for gamma-ray spectrometry systems. One of the naturally occurring isotopes of
cadmium (113 Cd) has an extremely high cross section (∼20,000 barns) for thermal-
neutron capture and emits a 558-keV prompt gamma ray. Background due to such
prompt gamma rays should be significantly lower than the 511-keV annihilation
peak and the 1461-keV 40 K peak in the detector background.
The detector, its neighboring structures, and the counted source itself also con-
tribute to the background. Reducing the detector volume can decrease the back-
ground count rate, but may also reduce the counting efficiency. This interaction
can be evaluated in terms of maximizing the figure of merit ε 2 /RB , where ε is
the counting efficiency and RB is the radiation background count rate. The fig-
ure of merit also permits comparison of sensitivity of two detector systems. For
example, if 89 Sr is counted with an efficiency of 0.96 and a background of 20
c/m in an LS counter and with an efficiency of 0.40 and a background of 1 c/m
in an anti-coincidence-shielded gas proportional counter, the respective figures
of merit are 0.046 and 0.16. In this comparison, the gas proportional counter is
more sensitive when other factors, such as sample volume and counting period, are
equal.
Radioactive contaminants in filters, planchets, detectors, and shields may be
primordial radionuclides and their progeny in aluminum (e.g., thorium), lead (e.g.,
uranium progeny), and filters (40 K). Materials with low radionuclide content should
be selected from carefully screened supplies. Man-made radionuclides have con-
taminated steel and other metals during processing (from airborne fallout radionu-
clides) or reprocessing (from radioactive tracers or medical irradiation sources
such as 60 Co, 137 Cs, or 226 Ra).
A common and effective way to reduce the external background is use of an
anticoincidence system (see Figs. 8.3 and 8.4). A separate detector shields the
sample detector. When the shield counter detects a pulse, the electronics system
momentarily turns off the detector to reduce the background count rate.
Research that requires extreme gamma-ray detection sensitivity, such as neutrino
and double beta decay measurements, has led to enormous reduction in gamma-
ray spectrometer background. Contributing to this improvement are extremely low
radionuclide contents in materials for the germanium detector, container and shield,
and reduction of the cosmic-ray background by underground location, electronic
anticoincidence, muon shielding, and shield material with low photon production
in muon interactions (Heusser 2003).
146 John M. Keller
Amp
Amp Gate
to stop all beta particles, it will still record the lower-intensity Bremsstrahlung
produced in the shield, as described in Section 2.4.3.
The reverse—gamma-ray background in beta-particle detectors—also occurs
due to electron-producing interactions of gamma rays in the walls of the gas-filled
proportional and G-M detectors. The magnitude of the detection efficiency for
energetic gamma rays, i.e., above 0.1 MeV, typically is about 1% of that for beta
particles (Knoll 1989).
An effective but expensive technique for monitoring background radiation is
operation of one of each set of detectors in the background measurement mode
at all times to alert the operator to fluctuations due to line noise, radon in air, or
external radiation sources. Another approach is to monitor the count rates recorded
in anticoincidence detectors for identifying fluctuations in the external background.
Detector contamination between regular background tests may be inferred from a
sudden increase to unusually high sample count rates.
8.3. Detectors
8.3.1. Gas-filled Detectors
Different types of radiation detectors that use a counting gas are based on a similar
design (see Fig. 2.10). Detectors are either sealed or with a continuous flow of gas.
A high voltage applied across the volume of gas in a conducting container forms an
electrical field between two electrodes—the detector wall and a central electrode.
As radiation passes through this electrical field with enough energy (∼25–35 eV
per ion pair) to ionize the counting gas, a flow of electrons to the wire anode creates
a current pulse between the electrodes. Based upon detector design, this electrical
signal is measured as a current, accumulated charge, or pulse which can be related
to the incident radiation. Examples of operating parameters for the three basic
types of gas-filled detectors are listed in Table 8.4. As described in Section 2.4.1,
detector design and operating voltage determine detection performance.
The ionization chamber generally integrates the current over a brief period and
measures it with a sensitive electrometer. An early version was an electroscope
charged by friction, in which the current was measured by the rate of separation of
two light metal foils that was a measure of their accumulation of electric charge.
The relatively large pulses from alpha particles are measured with a Frisch grid
chamber (Knoll 1989) with a resolution of about 40 keV. Its advantage is the
capability to measure and perform spectral analysis for samples with relatively
large areas, e.g., 500 cm2 .
A pressurized and sealed ionization chamber is used as calibrator for individual
radionuclides that emit gamma rays. Their activity is measured in terms of the
radiation dose, as discussed in Section 8.3.5.
The G-M counter is a simple and relatively inexpensive gas-filled tube with
a count-rate meter; an amplifier may also be present. The G-M detector counts
alpha particles, beta particles, and gamma rays with a very thin window, counts
beta particles and gamma rays with a thicker window, and counts gamma rays only
with a thick shield. Alpha and beta particles interact in the gas; gamma rays interact
mostly in the walls, from which electrons enter the gas. The intrinsic efficiency for
counting gamma rays relative to beta particles depends on the amount and type of
solids surrounding the detection gas.
The fill gas differs among types of detector. Almost any gas is suitable for an
ionization chamber. The counting gas for a proportional counter must have an
added organic quenching agent. The most commonly used gas is a mixture of
argon fill gas (90%) and methane quench gas (10%) readily available as P-10 gas.
In the G-M counter, if the tubes are sealed, the quenching agent usually is a small
amount of Br2 or Cl2 ; if the detectors are flow type, the quenching agent may be
organic compounds such as ethyl alcohol and ethyl formate. The fill gas is usually
a noble gas; for example, Q-gas is 97% He and 3% an organic compound.
The most common detector is the gas-flow proportional counter, although ion-
ization chambers and G-M counters can be usefully applied in the laboratory.
Several types of gas-flow proportional counters are used in modern counting
laboratories.
The first type is a manual detector, either with a thin (typically 80 μg/cm2 )
window beneath which the sample is placed, or without a window but with a
sample tray that slides under the detector and becomes the detector bottom. Use of
the internal proportional counter eliminates external attenuation of low-energy beta
particles but introducing the source into the detector can cause contamination or
reduce the potential difference to a value below the applied voltage. The detection
volume may be cylindrical (typically about 1 cm high and 3 cm in radius), or may
be a hemisphere in the classical 2πgroportional counter. For either a thin window
or no window, continuous gas flow is maintained. A thicker window (typically 0.5
mg/cm2 ) can seal the counting gas in the detector.
The second commonly available system is a detector coupled to an automatic
sample changer. In this arrangement, planchets or disks from a stack of 30 or
50 are fed individually by tray to the location beneath the detector window. The
controlling computer sets the counting time of typically 10 to 100 min, moves the
tray, and records the time and accumulated counts. The computer also controls
whether alpha or beta particles or both are measured and whether samples are to
8. Applied Radiation Measurements 149
5.70 cm.
Window and detector diameter
Guard detector
80 m g cm–2 Window
Depth of retaining
detector
window Sample detector ring Window to
carrier
distance
0.97 cm.
0.32 cm.
Slide Assembly
Horizontal view
Plastic Plastic
spacer spacer
Air
Plastic Plastic disk Stainless
ring steel pan
Expanded source mount
FIGURE 8.4. Configuration of the source and detector for the Tennelec LB5100, an end-
window, gas-flow, anticoincidence proportional counter (from Nichols 2006).
the linear amplifier gain and vice versa. The optimal operating voltage for each
mode is selected after finding the plateau by plotting count rate vs. bias voltage with
an alpha-particle and a beta-particle source (see Section 2.4.1). The pulses require a
preamplifier and an amplifier for counting. Pulses are proportional to energy depo-
sition, hence are hundreds of times larger for alpha particles than for beta particles.
Plateau curves are generated with a pure alpha-particle emitter (238 Pu, 244 Cm)
and a pure beta-particle emitter (63 Ni, 90 Sr/90 Y, 99 Tc) unless a combination of
radionuclides is selected to represent an actual sample. The analyst chooses, on
the basis of the two curves, a bias voltage for “alpha only” mode that is below the
point where beta particles just begin to be detected (see Fig. 2.11). At this voltage,
and also by application of pulse-shape discrimination, beta-particle “cross talk” is
minimized. The voltage for α + β counting is selected near the midpoint of the
beta-particle plateau. This plateau is not quite horizontal because an increase in
the applied voltage adds a few counts that previously had just been excluded by
the lower-energy discriminator.
The next step is to determine the counting efficiency for alpha- and beta-particle
measurements. The alpha-particle counting efficiency for thin samples with a pro-
portional counter is almost constant over the energy range of 4–9 MeV. This
energy range includes the naturally occurring isotopes of thorium, uranium and
their progeny, and also the actinides in global fallout from nuclear weapon test-
ing. A few naturally occurring rare earth elements (e.g., 147 Sm, 144 Nd, and 152 Gd)
emit alpha particles in the 2–3 MeV range, but these radionuclides have very low
decay rates associated with their long half-lives and are not of interest at most
radioanalytical chemistry laboratories.
Alpha particles that enter the gas proportional detector deposit energies of the
order of 1 MeV, far more energy than is needed for detection, hence alpha-particle
counting efficiency is a function of the source–detector geometry minus losses from
alpha-particle attenuation in the sample mass, intervening air, and detector window.
The ratio of count rate to disintegration rate (cpm/dpm) for alpha particles from a
very thin source mounted on stainless steel and counted within a windowless 2π
gas-flow proportional counter is ∼51.5%. The additional 1.5% over the expected
50% from the 2π geometry is due to backscattering of the alpha particles from a
stainless steel source mount.
End-window proportional counters with thinwindows that separate sample from
detector have the alpha-particle counting efficiency shown in Table 8.1 for thin
samples on planchets. The lesser counter efficiency relative to the internal detector
is due to less than 2π geometry and attenuation in sample, air space, and window.
Beta particles deposit typically about 2 keV of energy, in the range of 0.5–10
keV, in the gas proportional counter. Pulse height discrimination readily distin-
guishes these beta-particle pulses from the much more energetic ones generated
by alpha particles to provide a “beta only” mode. Cross-talk with alpha particles
can occur because the various angles at which the particles enter the detector
lead to considerably lower energy deposition for a small fraction of the particles.
Beta particles emitted with near-zero energy are not detected when their energy is
deposited in the sample, air, and detector window.
8. Applied Radiation Measurements 151
Unlike alpha particles, a significant fraction of beta particles that enter the
detector has been backscattered. Due to the energy distribution of beta particles,
the overall beta-particle counting efficiency for a proportional counter depends on
the fraction of low-energy beta particles absorbed in the sample, air, and window.
To quantify the activity for 63 Ni (66.9 keV) or 99 Tc (292 keV), for example, a
separate standard is needed for each radionuclide. The beta counting efficiency is
in the 65–75% range for radionuclides with maximum energies of 0.3–3.5 MeV
on steel planchets in a windowless gas-flow proportional counter. The efficiency is
less in an end-window counter (see Table 8.1), and can reach zero for beta-particle
groups with maximum energy below about 30 keV.
Other versions of the gas proportional counter system permit measurements
of special sources or radionuclides. Radioactive gas samples can be counted in
a detector that has valves for drawing a vacuum or admitting gases. A measured
volume of the gas sample is mixed with counter gas and counted in the detec-
tor (ICRU 1972). Alternatively, a flowing gas stream can be mixed with flowing
counting gas. Radioactive liquid samples can be placed in a thin-walled container
beneath an end-window counter, or a flowing stream can be passed through this
container.
Modified versions of the gas proportional counter are used for spectral analysis
of X rays in the energy region of a few kiloelectron volts. Detectors are constructed
with a greater depth and a window with a low-Z material such as Be, and with a
high-Z counting gas such as Xe at a pressure of several atmospheres. The resolution
is about 0.6 keV at 5 keV (Knoll 1989).
PMT 1
PMT 2
(dN/dE )
Unquenched
Emax
Quenched
sample or scintillation cocktail interferes with energy transfer between the solvent
and the scintillator. Color quenching is the attenuation of scintillator photons in the
solution due to light absorption. The difference between the two types of quenching
lies in the type of energy absorbed; chemical quenching absorbs deposited charged-
particle energy while color quenching absorbs emitted-photon energy.
Loss of energy, i.e., a reduction in the pulse height recorded at the PMT, is the
result in both cases as quenching shifts the energy spectrum to lower energies (see
Fig. 8.6). Any pulses shifted to below the low-energy discrimination are lost and
reduce the counting efficiency. Factors that can influence the degree of quenching
include the sample matrix, sample preparation, and choice of scintillation cocktails.
Several techniques have been developed to correct the effect of quenching.
The traditional method for quench correction is to add an internal standard to
each sample to determine the counting efficiency for each sample matrix. This
method continues to be one of the most accurate but is labor intensive and expen-
sive. A set of samples is counted without the internal standard. A duplicate set,
with a small volume (e.g., 0.1 ml or less), of known activity of the internal stan-
dard spiking solution is added to each sample, and is then counted. The counting
efficiency for each sample is as follows:
The values of cpms and cpms+i are the net count rates for each sample without
and with the added internal standard, respectively, and dpmi is the activity of
the internal standard added. The activity, dpms , for each sample is calculated by
dividing the cpms for each sample by the efficiency for each sample. This method
8. Applied Radiation Measurements 155
assumes that the chemical form of the radionuclide added as the internal standard
is the same as in the sample and that the added internal standard spiking solution
does not affect sample quenching.
Instrument vendors developed various techniques to measure the degree of
quenching in each sample with quench correction curves. The quench correc-
tion curve is prepared from a set of increasingly quenched standards with the same
level of activity. Then a plot of efficiency vs. the degree of quenching is prepared
for each radionuclide of interest. The success of quench correction curves depends
on skill in identifying the quenching agent and measuring the degree of quenching,
which is typically represented by a quench indicating parameter (QIP). Techniques
developed to define various QIPs are discussed in LS instrument manuals, vendor
literature, and other LS application descriptions (Horrocks 1974). The main types
of QIP used for quench correction curves are either based on the sample spectrum
or an external standard spectrum. The QIPs based on sample spectra are derived
from mathematical transformations of spectral data that provide a measurement
related to either the average or maximum spectral energy.
An external standard spectrum to determine the QIP is popular with users and
instrument vendors. The external gamma-ray source (e.g., 137 Cs, 133 Ba, or 152 Eu)
that is part of the detector system induces a Compton-electron spectrum in the
scintillation cocktail. Each sample is automatically counted with and without the
external standard. The Compton-electron spectrum produced in each sample vial
is applied with mathematical techniques to derive a QIP for a quench correction
curve.
One method offered by instrument vendors to determine the activity of a radionu-
clide is efficiency tracing (ET) (Takiue and Ishikawa 1978) for most beta-particle
radionuclides except tritium. The ET method uses a single unquenched 14 C stan-
dard to calibrate the LS system by collecting the spectrum and determining the
efficiency in several energy regions. Then, an unknown sample is counted in the
same energy regions. The cpm measured for the unknown in each counting region
is plotted vs. the efficiency determined for the 14 C standard in the same regions.
The generated curve is extrapolated to 100% efficiency to obtain the dpm for the
unknown activity.
As an example, the information in Table 8.6 summarizes the data collected for
a 14 C standard (134,900 dpm) in six energy regions of the beta-particle spectrum.
The table also includes the data from a simulated unknown sample with 90 Sr and
90
Y in secular equilibrium with a total beta-particle-activity of 1037 dpm.
1050
1000
y = 2.8057x + 765.31
950
Y (CPM)
900
90
Sr +
90
850
800
750
0 10 20 30 40 50 60 70 80 90 100 110
14
C Efficiency
The cpm of the “unknown” is plotted vs. the 14 C efficiency from each energy
region as shown in Fig. 8.7. The equation in Fig. 8.7 is the best fit to the data as
determined by linear regression. This linear equation is then used to extrapolate
the count rate of the “unknown” at 100% 14 C efficiency, which is equivalent to the
activity (dpm) for the 90 Sr + 90 Y in the sample. The calculated activity for this
example of 1045 dpm is within −0.8% of the known value.
Chemiluminescence and photoluminescence are other forms of interference
that can reduce the accuracy of LS techniques. “Chemiluminescence” describes
the emission within the scintillation cocktail of photons that result from a chemical
reaction; common initiators are samples with an alkaline pH or the presence of
peroxides. Photoluminescence can occur when the scintillation cocktail is exposed
to ultraviolet light. Some substances in the cocktail, notably the scintillator, are
excited and then emit light when the species return to ground state. The effect of
photoluminescence is reduced in LS systems by decay when the sample train is
held in a dark environment for a few minutes prior to counting. On the other hand,
chemiluminescence may have a slow decay rate that requires a change in sample
preparation to eliminate the chemical that causes it. Some LS systems identify
luminescence by pulse shape and indicate its relative extent.
Charged particle
Preamp Amplifier ADC MCA
detector
Bias
Supply
(depletion) zone of the detector. Only a very thin dead region is acceptable because
of the low penetrating power of alpha particles. The sample is counted in a small
vacuum chamber that provides a nearly air-free path between the sample support
plate and the detector. Without the vacuum, the alpha-particle spectrum is degraded
by interactions with air molecules.
Commercially available systems integrate in a single unit the high-voltage power
supply, amplifier, analog-to-digital converter (ADC) and multichannel analyzer
with the counting chamber, detector, and preamplifier, as shown in Fig. 8.8.
In general, larger detectors have worse resolution. A good compromise for most
laboratory applications is a Si diode detector with an active area of 3–5 cm2 , a
depletion zone of about 0.02 cm, and a resolution of 15–20 keV. This resolution
is sufficient for baseline separation of 239 Pu (5.15 MeV) from 238 Pu (5.50 MeV)
despite complications due to multiple alpha-particle groups per radionuclide. The
counting efficiency for surface barrier detectors depends on the geometrical factor
and can be expected to be constant at fixed geometry for very thin sources that
emit alpha particles in the typical energy range of 4–9 MeV. As described Section
7.2.3, preparation of a uniform and very thin source, usually by electrodeposition,
is essential for obtaining good peak resolution. This requirement may be relaxed
to the extent that low-energy tailing at a peak does not interfere with calculating
the activity, identifying the peak energy for the alpha particle, or measuring lower-
energy alpha particles.
Spectrometer energy is calibrated for these energies with available thin sources
of alpha particles at energies known within 0.1 keV. The energy peaks in an
acceptable system are almost symmetrical. Many radionuclides emit additional
alpha-particle groups at energies slightly lower by 30–50 keV than the main char-
acteristic peak. These are not clearly resolved from the characteristic peak, but do
not interfere seriously with calibration or detection because they are at much lesser
intensity.
Detectors within this range of dimension and resolution are suitable for most en-
vironmental radioanalytical chemistry applications. The alpha-particle emitters of
interest include isotopes of thorium, uranium, neptunium, plutonium, americium,
and sometimes curium (244 Cm at 5.80 MeV). Most of the major alpha particles
on this list can be resolved. Exceptions are 233 U from 234 U and 239 Pu from 240 Pu,
which require mass spectrometric separation to resolve them (see Table 6.3). A
158 John M. Keller
pair that can be distinguished only after chemical separation is 238 Pu and 241 Am
at 5.50 MeV.
Computation of activity is simplified because the same counting efficiency ap-
plies to all alpha particles in the usual energy range. The activity of a radionuclide
is calculated simply from the activity of the added tracer multiplied by the net
accumulated counts for the peak of the radionuclide of interest and divided by the
net counts for the tracer peak. Separate values of the counting efficiency and the
yield are not needed, although they may be of interest to monitor tracer activity
and process yield, respectively.
Alpha particles are also measured by spectral analysis in Frisch grid cham-
bers (see Section 8.3.1) and specially designed LS counting systems (see Section
7.3.3). Although solid-state detectors are most commonly used, the other detection
systems, if available, can simplify the processing of certain samples.
Energy (keV) εabs (%) FWHM (keV) ε abs (%) FWHM (keV)
60 (241Am) 50 8 22.4 1.2
(14%)
662 (137 Cs) 32 50 5.0 1.5
(7.5%)
1332 (60 Co) 25 95 2.5 1.9
(7.1%)
Application of this relation is shown in Table 8.7 for comparing the efficiency
and the resolution of two types of detectors with different dimensions. The com-
parison has been extended to three photon energies to demonstrate that comparison
at 1332 keV may not be sufficient for the purposes of the user because both rel-
ative efficiency and resolution change with energy. The comparison for practical
application also depends on sample dimensions and source-to-detector distance.
Table 8.7 data support the observations of better resolution but worse efficiency in
the HPGe detector than the NaI(Tl) detector.
The peak-to-Compton ratio is an important indication of detector performance
for measuring lower-energy gamma rays in the presence of higher energies. The
Compton continuum that is from 0 to ∼200 keV below the full-energy peak (see
Eq. 2.16) of 60 Co increases the spectral background across the entire low energy
region below the Compton edges. Most detector manufacturers report the peak-to-
Compton ratio as the full-energy peak count rate divided by the net count rate for
the same number of channels in the middle of the Compton continuum. This ratio
typically ranges from 40 to 70 for 60 Co measured with coaxial HPGe detectors
currently produced. The peak-to-Compton ratio is higher for detectors with better
efficiency and resolution.
The spectral response of a detector is more complex than described in Section
2.4.4 because of the bulk of the detector. The observed Compton continuum con-
sists of single plus multiple successive scattering interactions. When such multiple
Compton scattering interactions are terminated by a photoelectric interaction, the
pulse is added to the full-energy peak. Most of the counts in a full-energy peak
for gamma rays above 100 keV are due to such multiple scattering plus a final
photoelectric interaction.
Photoelectric interactions are recorded at the full energy peak despite the binding
energy loss (see Eq. 2.17) because the emitted X ray that follows electron emission
is detected simultaneously with the gamma ray. In small detectors, where the X
ray has a high probability of escape, a small second (“escape”) peak is seen at
the energy predicted by Eq. (2.17), which is below the gamma-ray energy by 28
keV in the NaI(Tl) detector and by 10 keV in the Ge detector. The dual peaks are
readily apparent at low gamma-ray energies.
Interaction by pair production results in a spectrum that includes escape peaks
at the full energy minus 511 keV and minus 1022 keV, when either one or both
of the positron annihilation photons do not interact with the detector. Compton
scattering of these photons adds to the continuum.
8. Applied Radiation Measurements 161
76 mm
IR window
1.3 mm 5 mm
Ge crystal
22.5
Lithium contact
(1.0 mm Ge equivalent)
1.6 mm
Core hole
(length:45.5 mm)
7.5 mm 0.75 mm
15.4 mm
25.4 mm 3 mm
25.4 mm
The previously cited sum peaks occur for two or more coincident gamma rays,
for example, at 2505 keV for 60 Co. Interactions outside the detector commonly
are detected as a peak at 511 keV due to annihilation radiation and at about 200
keV due to Compton scattering at 180◦ . Gamma rays produced by cosmic-ray
interactions in or near the detector are observed as discussed in Section 8.2.2.
All of these interactions by gamma rays with the detector can be modeled by a
gamma-ray simulation program such as the previously cited Monte Carlo n-particle
code, version 4. Modeling requires precise information on the location and material
of the source, detector, and surroundings. A description of the detector such as that
shown in Fig. 8.9 must be obtained from the supplier because the detector container
is sealed.
Selection of the type of detector for gamma-ray counting should be informed by
the analyst’s intention for the measurement. For example, NaI(Tl) detectors have
excellent efficiency and are useful for counting low-activity samples with few
radionuclides that emit gamma rays. Germanium detectors have superior energy
resolution, as shown in Fig. 2.14, for identifying numerous radionuclides that
emit gamma rays in a sample. Both systems must be evaluated in terms of their
background characteristics. Sodium iodide crystals of the same size will have
higher backgrounds inherently in spectral analysis because the peaks cover a wider
energy range; larger detectors have higher backgrounds because of their size.
Current efforts to select better detectors for spectral analysis have identi-
fied several materials, notably cerium-doped lanthanum chloride or bromide for
162 John M. Keller
scintillation systems and cadmium zinc telluride for solid-state systems. The en-
ergy resolution of the new crystalline scintillators is better than two-fold that of
thallium-doped sodium iodide and the detectors are less responsive to temperature
change. One drawback is the 138 La radionuclide that constitutes 0.09% of lan-
thanum in nature; moreover, thorough purification from thorium is necessary. The
new solid-state material has much better counting efficiency (due to its higher Z)
than the same volume of germanium and can be used at room temperature, but
resolution comparable to germanium has not yet been achieved.
9.1. Introduction
Systematic identification of potential radionuclides of interest usually is considered
at the beginning of a project, with further needs developing if results are questioned
in the course of the project. In practice, one identifies a radionuclide by finding
a match of measured decay characteristics to listed values. This comparison may
not be a simple matter. The effort entails selecting appropriate radiation detectors,
correctly interpreting the resulting data, and being aware of distinctive forma-
tion and decay characteristics that can distinguish otherwise similar radionuclides.
Correct radionuclide identification can be crucial to planning protective measures,
especially in emergency situations, by defining the type of radiation source and its
radiological hazard. Discussed here are the information used for radionuclide iden-
tification, the sources of this information, and the application of the information
in the radioanalytical chemistry laboratory.
A radionuclide is defined by its half-life and type, fraction, and energy of emit-
ted radiation. Also of interest are its atomic number (element) Z , its mass number
(isotope) A, whether it exists in an excited or ground state, and any parent–progeny
relation, as discussed in Chapter 2. The radioanalytical chemist needs this informa-
tion initially to plan the analytical program for measuring expected radionuclides
appropriately, and later to identify and quantify any unexpected radionuclides. The
professional should be familiar with the decay schemes of commonly encountered
radionuclides and with sources of information for all decay schemes.
Each radionuclide among the more than one thousand that are known has a
unique decay scheme by which it is identified. For this reason, among others,
researchers have studied decay schemes over the years and their reported infor-
mation has been compiled and periodically updated. The compiler surveys the
reported information for each radionuclide and attempts to select the most reliable
information for constructing a self-consistent decay scheme. The fraction of beta
particles that feed an excited state must match the fraction of gamma rays plus
conversion electrons emitted by the excited state. The energy difference between
any two states must be consistent with the energies of the transition radiations plus
the recoil energy of the atom that emitted the radiations.
163
164 Bernd Kahn
internal transition can be summed from this information on the assumption that
contributions are minimal from electron shells more distant from the nucleus. The
energies and fractions of characteristic X rays that accompany conversion electron
transitions may be reported in compilations or the energies and relative intensities
from the various K and L subshells are tabulated separately in compilations such
as Table 7 in Appendix F of the Table of Isotopes (Firestone 1996). The fractions
of characteristic X rays per conversion electron from that shell can be calculated
by multiplying the relative intensities from the shell by the K or L fluorescence
yield (see Table 3 in the same Appendix F).
The uncertainty information for these values that is needed to estimate the
uncertainty in the calculated radionuclide activity may be given in the compila-
tion or in the cited original studies. Any consistent deviations beyond this un-
certainty in quality assurance test comparisons from the reported values may
suggest, among other causes, an erroneous decay fraction in current use by the
laboratory.
sublevels such as Kα1, Kα2, and Kβ that differ by several kiloelectron volts for
the more energetic K X rays. Low-energy X rays are also spectrally analyzed with
gas proportional counters designed for that purpose, as described at the end of
Section 8.3.1.
Care should be taken to avoid misidentifying a radionuclide by peaks that are
artifacts of the detector. Most important are the escape peaks from annihilation
radiation cited above, and from low-energy radiation at the full energy minus
10 keV for the K X ray and 1.2 keV for the L X ray in Ge. Sum peaks are
displayed for a radionuclide that emits two or more gamma rays simultaneously
(“in coincidence”), as discussed in Section 9.3.5 for 60 Co, i.e., a 2505-keV peak
for two gamma rays with energies of 1173 and 1332 keV. Although not a peak,
the Compton edge, approximately 200 keV less than the full-energy peak (see
Eq. 216) in the usual gamma-ray energy region, may be misinterpreted as such
when seeking barely detectable radionuclides. Similarly, a slight increase in the
Compton continuum due to the backscattered electrons is associated with this
edge at about 200 keV. The radiation background in this energy region consists
of peaks for annihilation radiation, progeny of 222 Ra and 220 Rn gas, the terrestrial
radionuclides 40 K, uranium, and thorium, and cosmic-ray-produced radionuclides,
as discussed in Section 8.2.2.
The Si detector with spectrometer is used with thin sources to identify and
quantify radionuclides that emit alpha particles. All alpha particles are in the ap-
propriate energy range for detection unless attenuated in a thick source. Chemical
separation of the element of interest and meticulous preparation of the source usu-
ally are needed to obtain well-resolved peaks. Figure 9.1 shows the spectrum of a
100 234
238
U U
4200 4776
232
U
5320
75
Counts
50
4723
4150 5264
25
235
U
4392
0
3000 4000 5000 6000 7000
Energy (keV)
FIGURE 9.1. Alpha-partial spectrum of uranium with 232 U tracer, counted for 60,000 s.
170 Bernd Kahn
FIGURE 9.2. Conversion electron s measured whith a Si(Li) detector (from Knoll 1989,
p. 464).
separated uranium source, with the large peaks for 238 U and 234 U, the usual small
peak for 235 U, and the peak for the radioanalytical tracer 232 U. Many radionuclides
emit several alpha-particle groups at energies that—because of the small energy
difference and the lesser quality of the detector and the source cannot be resolved.
Some smaller peaks of lower-energy alpha particles are shown in Fig. 9.1. The
combined count rates of the unresolved peaks must be matched to the same com-
posite group in a known source for calibration, or the combined decay fractions of
all unresolved peaks must be used for calculating activity.
For some radionuclide mixtures, a group separation, e.g., for actinides, is sat-
isfactory for measuring its component radionuclides by alpha-particle spectral
analysis. As discussed in Section 6.4.1, further chemical separation is needed for
radionuclides that emit alpha particles of almost the same energies, or even a
mass spectrometer for radioisotopes of the same element with almost identical
alpha-particle energies such as 239 Pu and 240 Pu.
A thin solid-state detector with spectrometer is also useful for identifying and
quantifying conversion electrons. Figure 9.2 shows the spectrum of conversion
electrons from a thin source of 244 Cm. For radionuclides that also emit beta par-
ticles, the conversion electron spectrum may be underlain by the beta-particle
continuum.
selected on the basis of the chemical behavior of the radioelement of interest and of
possible interfering radioelements. When little is known about the presence of var-
ious radionuclides in the sample, an attempt is made to separate the radioelement
of interest from all conceivable radioactive contaminants. At best, the identity and
approximate amounts of contaminant radionuclides are known, and a purification
scheme can be developed as discussed in Chapter 6.
A purification procedure can be considered appropriate if separation from other
radionuclides is so effective that, in the absence of the radioelement of interest, no
radioactivity is detected in the purified sample. That is, the measured radioactivity is
zero within the measurement uncertainty. If the radioelement of interest is present,
then contaminants should not observably increase its measured activity. To plan
the procedure, the amount of each contaminant radionuclide that remains after
applying the separation methods is calculated in terms of its decontamination
factor (DF) (see Section 3.1). Meeting the stated criteria depends on the initial
concentrations of the radioelement of interest and of each contaminant radionuclide
in the sample, and on the achieved DF. Because these values generally are not
known initially, reasonable concentrations must be assumed and subsequently
revised when initial measurement results become available.
The other important purpose of chemical separation is to remove nonradioactive
substances that interfere with the selected separation procedures and measurement
of the purified source. In many instances, dissolved solids must be removed from
the sample solution to measure carrier yield without interference and to obtain thin
sources for counting alpha and beta particles, as indicated in Sections 7.2 and 7.4.
energy and half-life are scanned to look for other radionuclides with the measured
decay scheme.
FIGURE 9.3. Decay scheme for 89 Sr (modified from Lederer et al. 1967, p. 392).
9. Radionuclide Identification 173
FIGURE 9.4. Decay scheme for 90 Sr/90 Y (modified from Lederer et al. 1967, p. 392).
FIGURE 9.5. Decay scheme for 137 Cs (modified from Lederer et al. 1967, p. 399).
174 Bernd Kahn
FIGURE 9.6. Decay scheme for 40 K (modified from Lederer et al. 1967, p. 384).
FIGURE 9.7. Decay scheme for 60 Co (modified from Lederer et al. 1967, p. 389).
FIGURE 9.8. Decay scheme for 22 Na (modified from Lederer et al. 1967, p. 382).
9. Radionuclide Identification 175
241
FIGURE 9.9. Decay scheme for Am (modified from Martin and Blichert-Tolf, 1970,
p. 144).
r nuclear spin, e.g., 5/2−, 2+, shown at left on energy level line, that relates to the
intensity of transition
r gamma ray transition polarity, e.g. M4, E0, shown sideways on internal transition
vertical arrow
r log ft value, e.g., 12.1, 8.6, listed after the beta-particle energy, that defines the
spectral distribution of beta particles
Applications of these items are discussed in nuclear physics texts and summarized
in the nuclear and radiochemistry texts cited in Chapter 1.
176 Bernd Kahn
FIGURE 9.10. Radioactive daughter ingrowth—secular case (from Vertes et al. 2003,
p. 275).
Transient equilibrium occurs when the half-life of the daughter is only somewhat
shorter than that of the parent (Fig. 9.11). The slopes of the disintegration rate of
the daughter approaches that of the parent after about six half-lives. The important
distinction from secular equilibrium is that the disintegration rate of the daughter
FIGURE 9.11. Radioactive daughter ingrowth—transient case (from Vertes et al. 2003,
p. 273).
178 Bernd Kahn
FIGURE 9.12. Radioactive daughter ingrowth—no equilibrium case (from Vertes et al. 2003,
p. 276).
ultimately exceeds that of the parent by the factor t1 /(t1 – t2 ), where t is the half-life
and the subscripts 1 and 2 refer to the parent and daughter, respectively, as inferred
from Eq. (2.8).
No equilibrium is reached when the daughter half-life exceeds that of the parent.
As shown in Fig. 9.12, the daughter first accumulates as the parent decays, and
will remain for some time after the parent no longer can be detected. The daughter
will decay with its own half-life when, for practical purposes, the parent no longer
remains.
The decay constant of a daughter in equilibrium with its parent may be inferred
from repeated measurements of the count rate during ingrowth by use of Eq. (2.8),
or directly determined by chemically separating the daughter and then performing
repeated measurements of the count rate as the daughter decays.
half-life. Each emits a pure beta-particle group in 100% of the decay, which is
directly to the ground state. The maximum beta-particle energies are 546 keV for
90
Sr and 2280 keV for 90 Y. The minor (0.0115%) radiation emitted by 90 Y of
a 519-keV maximum-energy beta-particle group followed by 1760-keV gamma
rays is usually of no practical consequence. If the 90 Y is separated from 90 Sr, the
subsequent daughter ingrowth is described as secular equilibrium discussed in
Section 9.3.1.
of 7.3%. Positron decay produces two 511-keV gamma rays in the course of
annihilation, or 0.002%.
The 1461-keV gamma ray at 0.107 per disintegration and the beta-particle group
at 0.891 per disintegration are commonly measured. The other radiations—notably
positron decay—are at too low intensity for practical purposes.
several beta particle groups or gamma rays that have different counting efficien-
cies may be emitted, some in coincidence and others not in coincidence.
For 60 Co, gamma–gamma coincidence counting can be performed with a single
Ge detector and gamma-ray spectrometer by recording separately the peak count
rates for the two coincident gamma rays and their coincidences at the summed
full-energy peak. Corrections are necessary for the nonpeak count rate and any
significant angular correlation between the two gamma rays (NCRP 1985b). Coin-
cidences between other pairs of radiations are also used for absolute measurement.
spectrometric readings. The activity of total alpha and beta particles per sample
is estimated by assuming a typical or intermediate counting efficiency. These ef-
ficiency values must be revised once the samples are better characterized because
factors such as beta-particle energy, sample self-absorption (in the proportional
counter), and quenching (in the LS counter) affect counting efficiency.
If the gross alpha-particle activity is sufficiently high, a thin sample—of the order
of 1 mg/cm2 —is prepared for alpha-particle spectral analysis with a Si detector
plus spectrometer. The energy range of interest usually is 4–10 MeV.
Measurements are repeated at specified intervals to observe radioactive decay
or ingrowth. Counting usually is performed after increasing intervals, e.g., 0.5, 2,
8, and 24 h, as the initially rapid overall decay becomes slower. Decay or ingrowth
measurements for screening are ended when no further change is observed in the
count rate.
Any change in the net count rate associated with alpha- and beta-particle count-
ing provides a composite decay–rate curve that may be analyzed for a few com-
ponent half-lives by the procedure discussed in Section 9.3.1. These half-lives are
used to identify the radionuclides in the sample by scanning a list of radionuclides
tabulated by half-life. If the sample has many radionuclides with short half-lives,
only a general conclusion can be drawn concerning the presence of such radionu-
clides.
Initial measurements may stimulate further analysis with different sample mass
or type of measurement. A larger sample is prepared if the count rate was too low
for reliable measurement, and a more favorable geometry or longer counting period
may be selected. Too intense a sample suggests use of a smaller sample, greater
distance of source from detector, and shorter counting period. If the gamma-ray
spectrum suggests the presence of gamma rays below 30 keV, the sample can
be measured with an X-ray spectrometer to energies as low as 1 keV. A limited
table of gamma-ray energies and radionuclide half-lives used for identification—
either on paper by computer—may be modified to include additional radionuclides
and remove ones inappropriate for the recorded spectra. Other detectors that are
considered more useful for the type of sample or expected radionuclides may be
selected for further measurements.
When no radiation is observed above the detector background, the estimated
detection limit is compared to the regulatory limit or radiation protection guidance
to determine whether sample analysis results are sufficiently sensitive. If not,
larger samples, longer counting periods, and more efficient detection are required
for more sensitive measurements. Any radionuclides that have been tentatively
identified and quantified are compared to initial predictions to assure consistency
or consider reasons for differences.
Additional samples may have to be collected if an observed radionuclide can
be attributed either to the source of interest or to the ambient background due to
another source. These samples are analyzed to examine the pattern and concentra-
tion range of the radionuclide and establish as reliably as possible the distinction
between the two sources.
9. Radionuclide Identification 185
results rapidly by reducing the number of steps and simplifying the remaining
steps, but the reduced effort must provide sufficient DFs for anticipated impuri-
ties. Incomplete decontamination is acceptable only if the impurity is minor and
the measurement technique corrects for the impurity.
Radionuclides that emit only low-energy beta particles may not be detected by
gross beta-particle analysis with an end-window proportional counter, but can be
detected, although at low efficiency, by LS counting. These radionuclides include
3
H, 14 C, 32 Si, 33 P, 35 S, 45 Ca, 59 Ni, 63 Ni, 228 Ra, and 241 Pu. The long-lived radionu-
clides among these have very low decay rates and usually are not of concern as
immediate health hazards.
10.1. Introduction
The reputation of the radioanalytical chemistry laboratory is based on the extent to
which its reported results are judged to be reliable and reported in a form responsive
to customer needs. The effort to obtain and provide such data requires the competent
execution of every analytical step in the process, as well as adherence to the quality
assurance tenets discussed in Chapter 11. The result of these interlocking activities
should be an accurate and defensible data set. This chapter addresses the processing
and evaluation of those data.
Four types of calculations are typically performed that pertain to measured
radionuclide values:
r Conversion of the instrument signal (counts or count rate) and other measured
quantities to an activity concentration, typically expressed in becquerels (Bq) or
picocuries (pCi) per unit mass (or per unit volume, area, or time);
r Calculation of the uncertainty of the result in terms of an estimated standard
deviation or a multiple of the standard deviation;
r Calculation of the critical value, which is used to decide whether the radionuclide
is “detected”; and
r Calculation of the minimum detectable activity (MDA).
1
National Air and Radiation Environmental Laboratory, USEPA, Montgomery, AL 36115
2
Oak Ridge National Laboratory, Oak Ridge, TN 37831
189
190 Keith D. McCroan and John M. Keller
the information required by the initial plan, and conveying to the reader a clear
understanding of the radiological impact of the subject. Section 10.6 describes the
process of data review, the steps that ensure the viability of the data. Section 10.7
addresses some of the important details of data presentation.
The mass m in Eq. (10.1) may represent mass of material that is moist, dried, or
ashed. The concentration may be reported on the basis of moist weight to indicate
the radionuclide concentration in the environment, or on the basis of dried or
ashed weight for reliable comparison with measurements of other samples or by
others. For such comparisons, aliquots of a sample are weighed moist as collected,
then after drying at 110◦ C, and then again after thorough ashing, both at least
overnight.
The decay or ingrowth factor D is a correction factor for decay or ingrowth of
the radionuclide before and during counting. For the simple situation where the
radionuclide of interest is not supported by a parent in the sample, the decay factor
is given by
1 − e−λtR
D = e−λtD × (10.2)
λtR
Here λ denotes the radionuclide decay constant,tD denotes the time from sample
collection to the start of sample counting, and tR denotes the real time, or clock
time, of the sample counting measurement, which may be longer than the live
time tG . When the time tR is short relative to the half-life of the radionuclide, the
product λtR may be very small, and as λtR approaches zero, the correction factor
for decay during counting, (1 − e−λtR )/λtR , approaches 1. If the value of this factor
is sufficiently close to 1, it may be omitted from the expression for D. When it is
included, it must be calculated carefully to avoid large round off errors when λtR
is small.
Note: For small values of λtR , the factor (1 − e−λtR )/λtR may be approximated
either by e−λtR /2 or by the sum of the first few terms of the series 1 − λtR / 2! +
(λtR )2 / 3! − (λtR )3 / 4! + · · ·.
Instead, one uses a mathematical model of the measurement to relate the values of
input quantities, which are observed or previously measured, to the value of the
desired output quantity (the measurand), which must be calculated. The mathe-
matical model is an equation or a set of equations that describe exactly how one
calculates the value of the measurand from the observed values of the input quan-
tities. The model for a radioanalytical measurement often resembles Eq. (10.1),
but for the purpose of explaining uncertainty evaluation, a model that follows the
GUM is written abstractly in the form
Y = f (X 1 , X 2 , . . . , X N ) (10.3)
where X 1 , X 2 , . . . , X N denote the input quantities, and Y denotes the output quan-
tity (not the yield in this case), or measurand.
A particular observed value of an input quantity X i is called an input estimate.
If we denote the input estimates by lowercase variables x1 , x2 , . . . , x N , then the
output estimate, y, is calculated as follows:
y = f (x1 , x2 , . . . , x N ) (10.4)
Each input estimate xi has an uncertainty, u(xi ). In principle the standard uncer-
tainty of an input estimate can be evaluated in many ways. For example, if the value
of the input quantity can be measured repeatedly, then a series of observations of
the quantity, xi,1 , xi,2 , . . . , xi,n , can be obtained to calculate the arithmetic mean
x̄i and experimental standard deviation s(xi,k ):
1 n 1 n
x̄i = xi,k and s(xi,k ) = (xi,k − x̄i )2 (10.5)
n k=1 n − 1 k=1
The experimental
√ standard deviation of the mean s(x̄i ) is obtained by dividing
s(xi,k ) by n:
s(xi,k ) 1 n
s(x̄i ) = √ = (xi,k − x̄i )2 (10.6)
n n(n − 1) k=1
One then lets xi equal the arithmetic mean of the observations and lets u(xi ) be
the experimental standard deviation of the mean:
xi = x̄i
u(xi ) = s(x̄i ) (10.7)
This type of uncertainty evaluation is an example of what the GUM calls a Type A
evaluation of uncertainty, which is defined as an evaluation of uncertainty based
on the statistical analysis of series of observations (ISO 1995).
Many methods of uncertainty evaluation do not involve the statistical analysis
of series of observations; these are called Type B evaluations. One of the most
common Type B methods of uncertainty evaluation in radioanalytical chemistry is
the common practice√of estimating the standard uncertainty of an observed count
C by its square root C (see Section 10.3.4).
194 Keith D. McCroan and John M. Keller
1 n
xi = x̄i = xi,k
n k=1
1 n
x j = x̄ j = x j,k
n k=1
1 n
u(xi , x j ) = (xi,k − x̄i )(x j,k − x̄ j ) (10.8)
n(n − 1) k=1
Other possible methods for estimating the covariance u(xi , x j ) are described by
MARLAP (EPA 2004).
All the uncertainties u(xi ) and covariances u(xi , x j ) of the input estimates com-
bine to produce the total uncertainty of the output estimate, y. The mathematical
operation of combining the standard uncertainties and covariances of the input
estimates xi to obtain the standard uncertainty of the output estimate y is called
propagation of uncertainty. The standard uncertainty of y obtained by uncer-
tainty propagation is called the combined standard uncertainty of y and is de-
noted by u c (y). The following general equation, which the GUM calls the “law of
10. Data Calculation, Analysis, and Reporting 195
The following applications show how Eq. (10.10) can be applied to some simple
models when the input estimates are uncorrelated. In these equations, the variable
xi denotes input estimates and the variable y denotes the calculated output estimate;
a and b are constants.
Application 1: Addition and Subtraction
y = x1 ± x2
u c (y) = u 2 (x1 ) + u 2 (x2 )
Application 2: Multiplication
y = x1 x2
u c (y) u 2 (x1 ) u 2 (x2 )
u c (y) = x22 u 2 (x1 ) + x12 u 2 (x2 ) = +
y x12 x22
Application 3: Division
x1
y=
x2
+ +2 +3
The uncertainty of the input quantities in this model is discussed below in Section
10.3.8. Some quantities, such as the decay factor D and the yield Y , are not directly
observed but are calculated from other observed quantities. Their uncertainties are
calculated by uncertainty propagation before being used in the equation above.
The uncertainties of the count times, u(tG ) and u(tB ), and the uncertainty of the
decay factor, u(D), usually are negligible and can be omitted from the uncertainty
equation. Then the simplified uncertainty equation becomes
2
u 2 (CG )/tG2 +u 2 (CB )/tB2 u (ε) u 2 (Y ) u 2 (m) u 2 (F)
u c (a) = +a ×
2 + 2 + +
ε2 Y 2 m 2 D 2 F 2 ε2 Y m2 F2
(10.12)
The uncertainty of a count time may be relatively significant if the count time is
very short (say less than 1 min), if the number of counts is very large, or if the dead
time rate is large. The uncertainty of the decay factor is most significant when the
radionuclide is short-lived and the decay time is either uncertain or long.
198 Keith D. McCroan and John M. Keller
The preceding uncertainty equations presume that all pairs of input estimates are
uncorrelated, which may or may not be true. One of the most common examples of
correlated input estimates in radioanalytical chemistry occurs when the chemical
yield Y is calculated from an equation involving the counting efficiency ε. This
happens in measurements by alpha-particle spectrometry with an isotopic tracer. In
this case, the uncertainty equation can be simplified by treating the product ε × Y
as a single variable. What happens in effect is that the efficiency cancels out of the
activity equation and for this reason its uncertainty can be considered to be zero:
2
u 2 (CG )/tG2 + u 2 (CB )/tB2 u (ε × Y ) u 2 (m) u 2 (F)
u c (a) = +a ×
2 + +
ε2 Y 2 m 2 D 2 F 2 (ε × Y )2 m2 F2
(10.13)
Another example of correlated input estimates occurs when both the counting
efficiency ε and the yield Y depend on the mass of a precipitate or residue on the
prepared sample source. In this case, dealing with the correlation is less simple. It
may be necessary to replace the variables ε and Y in the activity equation by the
expressions used to calculate them, or to include the covariance term for ε and Y
in the uncertainty equation.
r The source in the detector contains a large number, N , of atoms of some long-
lived radionuclide.
r Each atom of the radionuclide has the same probability p of decaying during the
counting period, emitting the radiation of interest, and producing a count. When
the half-life of the radionuclide is long, p is a very small number.
r All atoms of the radionuclide decay and produce counts independently of each
other. Decay by one atom that produces a count has no impact on whether another
atom decays and produces a count.
r No atom produces more than one count during the counting period.
r Counts are not produced by other sources.
Given these assumptions, the distribution of the observed total number of counts
according to probability theory should be binomial with parameters N and p.
Because p is so small, this binomial distribution is approximated very well by the
Poisson distribution
√ with parameter Np, which has a mean of Np, and a standard
deviation of N p. The mean and variance of a Poisson distribution are numerically
equal; so, a single counting measurement provides an estimate of the mean of
the distribution Np and its square root is an estimate of the standard deviation
√
N p. When this Poisson approximation is valid, one may estimate the standard
uncertainty of the counting measurement without repeating the measurement (a
Type B evaluation of uncertainty).
Although all the assumptions stated above are needed to ensure that the distri-
bution of the total count is binomial, not all the assumptions are needed to ensure
that the Poisson approximation is valid. In particular, if the source contains several
long-lived radionuclides, or if long-lived radionuclides are present in the back-
ground, but all atoms decay and produce counts independently of each other, and
no atom can produce more than one count, then the Poisson approximation is still
useful, and the standard deviation of the total count is approximately the square
root of the mean.
Note: One does not evaluate the uncertainty of a count rate, R = C/t, by taking the
square root of R. If the total count C has a Poisson distribution √
and t has √
negligible
uncertainty, the standard uncertainty of R is given instead by C/t or R/t.
The Poisson approximation is not always valid. For example, if the half-life of
one or more radionuclides in a source is short relative to the count time t, the Pois-
son distribution may not be a good approximation of the binomial distribution. Also
note that some radiation-counting measurements involve the counting of particles
emitted by a radionuclide and its short-lived progeny, where one atom may produce
several counts as it decays through a series of short-lived states. A well-known
example of this type of measurement is alpha-counting 222 Rn and its progeny in
an alpha scintillation cell, or “Lucas cell.” For such measurements, neither the
binomial nor Poisson model is valid. Another example is the use of the Poisson
model to estimate the total uncertainties of gross and background counts measured
by gamma-ray spectrometry. While the use of the Poisson model is applicable in
this case, there are some complications to consider (see Sections 10.2 and 10.3.7).
200 Keith D. McCroan and John M. Keller
μx e−μ
Pr(X = x) = (10.14)
x!
When the mean μ is small, the distribution of values about μ is skewed because
Pr(X = x) cannot be negative. When μ is large—say 20 or greater—the Poisson
distribution approaches a Gaussian distribution like the one shown in Fig. 10.1 (but
√
with the standard deviation σ = μ). The probability function for the Gaussian
approximation of the Poisson distribution is shown in Eq. (10.15):
1 (x−μ)2
Pr(X = x) = √ e− 2μ (10.15)
2πμ
Figure 10.2 shows how the Poisson distribution approaches the Gaussian ap-
proximation as the mean μ increases. In the figure, the height of a bar over a
value x represents the actual Poisson probability p(x) of observing that value,
while the height of the smooth curve over the point x represents the Gaussian
approximation of this probability. When μ is only 3, it is clear that the Pois-
son distribution is skewed and the Gaussian approximation does not describe it
well. When μ = 10, the distribution is more symmetric and superficially Gaus-
sian in shape. By the time μ reaches 20, the Gaussian approximation is good
enough for many purposes, and the approximation only improves as μ increases
further.
p(x) p(x)
=3 = 10
x x
0 1 2 3 4 5 6 7 8 9 10 11 12 0 2 4 6 8 10 12 14 16 18 20 22 24
p(x) p(x)
= 20 = 36
x x
0 4 8 12 16 20 24 28 32 36 40 44 48 0 6 12 18 24 30 36 42 48 54 60 66 72
10.3.7. Spectrometry
Uncertainty evaluation for gamma-ray spectral analysis is more complicated
than for most other radiation measurements. First, gamma-ray counts for one
202 Keith D. McCroan and John M. Keller
where
ur denotes the relative standard uncertainty due to subsampling
mS is the mass of the subsample (aliquot)
mL is the mass of the entire sample
k is, by default, 0.4 g cm−3
d is the maximum particle diameter (size of a square mesh), in cm
For example, if a very large sample (m L = ∞) is ground until it passes through
a sieve with mesh size d = 0.1 cm, and a subsample of mass m S = 0.25 g is
removed for analysis, this equation estimates the relative subsampling uncertainty
to be u r = 0.04, i.e., 4%.
in Eq. (10.18):
1
L C = t1−α (n − 1) × s(Bi ) × 1+ (10.18)
n
In this equation t1−α (n − 1) denotes the (1 − α)-quantile of the Student’s
t-distribution with n − 1 degrees of freedom. For example, if α = 0.05 and n = 10,
then t1−α (n − 1) = t0.95 (9) = 1.833.
If one performs the series of blank measurements described above, calculates
L C by Eq. (10.18), and subsequently calculates the blank-corrected activity A for
an analyte-free sample, the probability of observing a value of A greater than L C
is approximately equal to α. If one makes the detection decision by comparing A
to L C , the false positive rate should be approximately α.
If α = β and the number of blank measurements n is not too small (say
at least 5), and if the measurement variance does not increase rapidly with activity,
the minimum detectable absolute activity L D may be approximated as shown in
Eq. (10.19):
LD = 2 × LC (10.19)
Typically, one expresses the MDA as a massic or volumic activity by dividing
L D by the amount of sample analyzed. If α = β or if variance increases too rapidly
with activity, Eq. (10.19) is inappropriate. For these cases see MARLAP (EPA
2004).
When the radiation counting statistics are essentially Poisson for the background
count and there are no interferences, the critical net count rate (critical value of
the net count rate) may be calculated as follows:
CB 1 1
SC = z 1−α + (10.20)
tB tG tB
where
SC is the critical net count rate in s−1
CB is the observed background count
tB is the background count time in s
tG is the sample (gross) count time in s
α is the significance level (e.g., α = 0.05 by default)
z 1−α is the (1 − α)-quantile of the standard normal distribution (equal to
1.645 when α = 0.05).
In this case one makes the detection decision for a sample by comparing the
observed net count rate RN to the critical net count rate SC .
When Eq. (10.20) is used for the critical value and α = β, a commonly used
approximation formula for the MDA is
z2 1 1
+ 2z RB +
tG tG tB
MDA = (10.21)
K
10. Data Calculation, Analysis, and Reporting 207
where
z = z 1−α = z 1−β
RB is the background count rate in s
tG is the sample (gross) count time
tB is the background count time
K is an appropriate estimate of the denominator by which the net count
rate is divided to calculate the final result of the analysis.
The denominator represented by K above might, for example, be the product
ε × Y × m × D × F, which appeared in the previous example of a mathematical
model for a radioanalytical measurement (Eq. 10.1). If there is significant vari-
ability in this denominator, a somewhat low estimate of its value should be used
to avoid underestimating the MDA.
In many cases, the assumption of pure Poisson counting statistics is invalid
because other sources of variability affect the distribution of measurement results.
In these cases, the laboratory should consider determining the critical value and
MDA based on a series of blank samples, as in Eqs. (10.18) and (10.19). When
the Poisson model is valid but the background level is so low that the Poisson
distribution is not approximately normal, other formulas for the critical value and
MDA tend to give the best results, as discussed in MARLAP (EPA 2004).
Note: Radioanalytical chemists should avoid the common mistake of using the
MDA as a critical value.
15,000
14,000
Counts
13,500
13,000
s limit s limit
12,500
01/01/2000 12/31/2000 01/01/2002 01/01/2003 01/01/2004 12/31/2004
Date
the mean and bounding values for the source must curve downward with the slope
defined by the radioactive decay constant.
The instrument control source is not a standard. Its main criteria are stability—
i.e., no observable change in the radionuclide content (other than by decay) or
configuration—despite frequent handling over the period of use, and known ra-
dioactive constituents. Its radionuclide content is selected to have a count rate that
is high but not excessive, as discussed above, so that a 1- to 10-min counting pe-
riod is sufficient to accumulate about 10,000 counts. The background count usually
requires a much longer time period—50 to 1000 min, depending on the type of
detector—but still will have a larger relative standard deviation than the source.
Hence, the upper and lower bounds for the source control chart will be much closer
together, i.e., have a smaller relative standard deviation, than for the background
control chart.
During routine operation of the instrument, the control source and the back-
ground are measured at selected intervals and the number of counts is entered on
the control chart. This may be done by hand or by computer, as in Fig. 10.3. A
measurement point that falls within the 3σ control limits is acceptable.
Any value outside the 3σ limits indicates that a problem exists and correction is
required. The instrument should be checked for easily corrected problems such as
control settings, computer instructions, applied voltage, radioactive contamination,
ambient radiation levels, temperature changes, dirt, and noise. Points outside of
the 3σ limit may occur, for example, if small errors in the decay correction cause
an upward or downward trend in the data that results in outlying values over time.
If the problem cannot be corrected at this level of trouble shooting, the instrument
is scheduled for repair. Values that are too low generally result from component
failures or a decrease in the applied voltage. Values that are too high often result
210 Keith D. McCroan and John M. Keller
Here A is the measured sample activity and u c (A) is its combined standard un-
certainty. Acceptance limits for Z rb may be analogously to 3σ control limits on
a control chart or at other values based on the desired false rejection rate; they
could, for example, be based on the estimated number of effective degrees of free-
dom for the combined standard uncertainty u c (A), described in the note in Section
10.3.2. Thus, the result of the reagent blank analysis A is found to be acceptable
if its absolute value does not exceed a specified multiple of its combined standard
uncertainty u c (A).
Laboratory-fortified blanks and matrix spikes both test the analyst’s ability to
obtain the expected result. The extent to which the net radionuclide concentration
of the fortified blank (corrected for yield and radioactive decay) deviates from the
expected value for the tracer radionuclide concentration is a measure of analytical
bias. Any consistent deviation from the expected value should be investigated to
eliminate the cause. Typical causes are the wrong counting efficiency, an analytical
problem with interchange between carrier and tracer, unreliable yield determina-
tion, or erroneous tracer radionuclide concentration.
The recommended test statistic for a blank spike is computed as follows:
A−K
Z bs = (10.23)
u 2c (A) + u 2c (K )
Here A is the measured sample activity, K is the added spike activity, and u c (A)
and u c (K ) are the combined standard uncertainties of A and K , respectively.
Acceptance limits for Zbs may be set at value such as 3.
The percent recovery (%R) for a blank spike is computed as in Eq. (10.24):
A
%R = × 100% (10.24)
K
Here Arepresents the measured spiked sample activity and K is the actual activity
of the spike added.
The difference between the activity of the matrix spike and that of the unspiked
matrix should be approximately equal to the activity of the added radioactive tracer.
The recommended test statistic for a matrix spike analysis is computed as in Eq.
(10.25):
aS − aU − aK
Z ms = (10.25)
u 2c (aS ) + u 2c (aU ) + u 2c (aK )
Here aS is the measured spiked sample result, aU is the unspiked sample result,
aK is the spike added, and u c (aS ), u c (aU ), and u c (aK ) are the estimated standard
deviations in aS , aU , and aK , respectively. All variables are again in the same units.
If required, the percent recovery for a matrix spike is computed as in Eq. (10.26):
aS − aU
%R = × 100% (10.26)
aK
where aS represents the measured spiked sample result, aU the unspiked sample
result, and aK the actual concentration of spike added.
212 Keith D. McCroan and John M. Keller
Any difference in tracer results between the spiked and unspiked blanks and
the spiked and unspiked matrix usually can be attributed to the effect of different
chemical and physical forms of the radionuclide of interest and its tracer. In that
case, the tracer does not represent the radionuclide of interest, and some treatment
of the tracer plus radionuclide of interest, such as a redox process, will be needed
to place them into identical forms. The cause of such inconsistency needs to be
examined and resolved if the difference exceeds the uncertainty of the two values.
Accumulated results from laboratory-fortified blanks and matrix spikes are a
measure of both precision and bias. To determine precision, numerous replicate
values can be examined and the standard deviation can be calculated. The mean
value is compared to the expected tracer concentration to determine bias. For this
application, the results collected at different periods must be seen to belong to the
same set, i.e., there is no obvious temporal difference among measurements due
to analytical or measurement problems. The calculated standard deviation value
can be compared to the suitably propagated components for counting and for the
rest of the analytical process. Causes of unexpectedly large or small values of the
standard deviation should be examined.
To obtain reliable values of the standard deviation, the accumulated tracer-
related count must be sufficiently large, e.g., at least several hundred counts. The
amount of tracer that is added must be similar to the amount of the radionuclide
of interest in order to represent that level of activity but large enough to achieve a
relatively small standard deviation of counting in a reasonable period. The latter is
especially of concern for the matrix spike, where the result is the difference between
counts that are often relatively small for the spiked and unspiked samples.
Combined analytical and measurement bias is also evaluated by periodic anal-
yses of interlaboratory comparison samples described in Section 11.2.10. The
radionuclide concentrations in these samples should have been measured with
great care. Samples are submitted for blind analysis, i.e., they are not identified as
test samples.
Here a1 and a2 are the two measured sample results, and u c (a1 ) and u c (a2 ) are their
combined standard uncertainties. Acceptance limits for Z dup may be set at ±3 or
at other values if desired.
If required, the relative percent difference (RPD) for a pair of duplicate analyses
may be computed using Eq. (10.28):
|a1 − a2 |
RPD = × 100% (10.28)
(a1 + a2 )/2
where a1 and a2 represent the two measured results. If the average of the two
results a1 and a2 is less than or equal to zero, the RPD is undefined.
Applicability of radioactive tracer tests to actual radionuclide measurements is
always in question because the radionuclide in the sample may be in a different
chemical or physical form. If the radionuclide has multiple oxidation states in
nature, e.g., iodine or plutonium, a tracer study will be applicable only if the initial
step in the procedure provides for the interchange of all possible oxidation states,
or if tracers have been used in all possible oxidation states. Tests can become even
more elaborate if the radionuclide of interest is an integral part of a solid sample
matrix, e.g., biological material or soil.
In such cases, it may be desirable to reproduce the actual sample with tracers; for
example, radioactive tracer is added to the roots of the growing vegetation that is to
be analyzed. If going to such lengths is not feasible, all portions usually discarded
during the testing of the procedure should be analyzed for the radionuclide to check
for losses not indicated by chemical yield measurements.
The listed items emphasize that each sample must be processed with alertness at
every step. At collection, the sample must be accompanied by all of the informa-
tion required to calculate results and to assure that the correct sample is collected
in the specified manner. The sample must be preserved and stored appropriately.
Familiarity with the sampling plan and previous shipments is desirable to enable
recognition of deviations from the plan. Care in sample identification avoids sam-
ples being switched, lost, or analyzed for the wrong constituents. Observation
of strange behavior during analysis may indicate reasons for unusual yields or
counting results. Application of inappropriate calculations and factors, by hand or
computer, are common causes of error.
Problems are best recognized and corrected at the time of occurrence. If found
later, during data compilation, they may be difficult or even impossible to remedy.
At the very least, however, false results will not be reported.
Findings based on QC results of unacceptable circumstances during a period
of operation (e.g., unduly elevated sample radiation background or deviation of
reported tracer concentration addition from the expected value) invalidate sample
data for the entire period to which the QC data pertain. This period extends back to
the previous QC tests, although an investigation may narrow the problem period by
dating the cause of error, e.g., onset of contamination or error in reagent preparation.
The inverse inference, that correct results in the QC program validate the processes
during the period since the previous QC test, is not necessarily true because errors
may have occurred between QC tests.
One type of reality check is comparing a radionuclide measurement with values
for similar samples collected at the same time or previously, either in the same
program or reported by others. It is usually invoked when the result under consid-
eration appears to be unusually low or high. As indicated in Chapter 12, the best
response is reanalysis. If the questioned value is confirmed, an unexpected situa-
tion has been found. If the second result does not confirm the questioned value,
the source of error should be sought in the analysis, radiation detection, sample
identification, preservation, or possible contamination.
Reports of false positives or false negatives are a common problem, especially
for gamma-ray spectral measurements with data analysis by computer. A false
negative value may be suspected because the radionuclide is known to be at the
monitoring site. A false positive value may be suspected because the radionuclide
is unlikely to be at the location or in the sample—for example, a short-lived fission
product where a nuclear reactor has been inactive for years. Results and calculations
should be reexamined if the reviewer has doubts about the presence or absence of
a reported radionuclide. Repetition of the measurement for a longer period or with
10. Data Calculation, Analysis, and Reporting 215
a more sensitive detector can be ordered; ultimately, the sample can be reanalyzed
if portions are still available.
In a gamma-ray spectrum, the reexamination may consist of considering details
of the sample plus detector background. For example, peaks in soil samples at
766 and 835 keV that may be falsely attributed to 95 Nb and 54 Mn, respectively,
actually can be minor gamma rays in the natural uranium and thorium decay
chains. For measurements near or below the MDA, the usual peak identification
and quantification software can be replaced with one that is more or less responsive
to channel-by-channel fluctuations.
A reality check may also focus on scanning the raw data and initial calcu-
lations to seek major deviations from the norm. This search typically unearths
typographical errors such as decimal-point shifts and digit reversals, or shifts
of entire lines or columns. Obvious identifiers of error are improbable weights
and volumes, unacceptable yields, and unusual detection efficiencies. More subtle
sources of error are wrong decay schemes or interference by the natural back-
ground.
The chemical and physical characteristics of a radionuclide that affect its inter-
action with the sample matrix may offer suggestions whether its presence can be
expected. For example, thorium and radiozirconium compounds are insoluble in
water except under highly acid conditions and would not be expected in drinking
water. Anions generally are not sorbed on soil and remain in groundwater. Gener-
alities such as these, however, can be invalid under local conditions. Radiocesium
ions are strongly retained in some soil types but not in others. Radioiodine can be
found in airborne particles and also as a gas. Plutonium and the radiolanthanide
compounds are insoluble in water near neutral pH values but can be soluble as
organic complexes at the same pH values.
Uranium and technetium are soluble in water when oxidized but insoluble in
reduced form. Radioisotope pairs, parent–progeny relations, and specific activity
may provide guidance in assigning the origin of radioactive material and identify
questionable results for a specified location. Certain radioactive materials have
“signatures” of uranium or plutonium isotopes at known ratios. Ratios for other
radionuclide pairs, such as 89/90 Sr and 103/106 Ru, can suggest the time interval
since formation. The time interval can also be suggested by ingrowth of shorter-
lived progeny into long-lived parents. The 134/137 Cs ratio provides information on
the neutron environment at origin. The specific activity of 3 H, 14 C, and 129 I can
distinguish between natural and anthropomorphic creation.
The data reviewer may be guided in expecting a certain radionuclide concentra-
tion in a sample matrix by the radionuclide concentration in a related matrix and
the reported transfer factor between these matrices. Relations exist among environ-
mental samples of air, rain, vegetation, and milk, and between industrial samples
of airborne particles and workers’ urine. These relations are complex and often
time- and site-related, but calculations ranging from rule of thumb to an elaborate
computer program may suggest the magnitude of radionuclide concentration ratios
among such matrices. A major inconsistency in the concentration of a radionuclide
among related sample matrices suggests reevaluating the analytical data, although
216 Keith D. McCroan and John M. Keller
the conclusion may be that the data are reliable but the applied transfer coefficients
are inappropriate.
A common problem in radioanalytical chemistry laboratories, especially those
that analyze environmental samples, is the existence of an overwhelming number
of samples without detectable radionuclide activity, which can lead analysts and
data reviewers to expect such results as a matter of course. Reviewers who expect
undetectable results for samples may tend to interpret small positive values as false
positives even when they are not. The tendency to let one’s preconceptions influ-
ence one’s judgment in data interpretation must be resisted. Detection decisions
and other evaluations of the data should always be based on objective criteria. If
the laboratory’s objective criteria indicate the presence of an unexpected radionu-
clide, an investigation may be needed to confirm that the measurement process is
performing properly.
prepare both types of table, the former for the professional reader and the latter
for overview by management and the public. Typically, the client of a laboratory
should be knowledgeable enough to interpret a full report that includes all data,
but tables and data summaries prepared for others may be tailored to the needs of
the audience.
Multiple results, such as weekly data in an annual report, can be tabulated
as the mean value with its standard deviation or values of the mean, highest,
and lowest value by location and sample medium. Additional information should
be provided if elevated radionuclide levels occur at specific times or locations.
Patterns of changing radionuclide concentrations can be presented as time-line
graphs. Spatial patterns of radionuclide concentrations can be presented as maps
with concentration isopleths, which indicate the points at which each radionuclide
has a specified constant concentration.
Presentations in figures can clarify and emphasize relations of radionuclide
concentrations to release levels or release points. Figure 10.4 compares the higher
radionuclide concentrations in river water downstream (“US 301 bridge”) from a
10. Data Calculation, Analysis, and Reporting 219
11.1. Introduction
Quality assurance (QA) describes the effort by a laboratory organization to pro-
duce trustworthy results. Every laboratory, no matter how small, must maintain
a continuing effort to confirm instrument calibration, measurement reproducibil-
ity, and applicability of analytical methods. These efforts must be documented so
that the results achieved by the laboratory can be used confidently in a decision-
making program. Various decision-making programs are supported by radioana-
lytical chemistry. A partial list of these programs includes
r protection of workers, population, and the environment in accord with regulations
and professional guidance;
r determination that contractual requirements have been met, with regard to facility
operation or decommissioning;
r verification that a decontamination response is necessary for a nuclear incident
with the potential for environmental contamination;
r litigation regarding nuclear attribution or worker compensation;
r demonstration of compliance with guides or regulations for pollution control,
e.g., for drinking water or air quality;
r comparison of results among laboratories;
r confirmation of compliance with international trade agreements, e.g., food
import;
verification of compliance with nuclear nonproliferation treaties such as com-
prehensive test ban monitoring; and
r publication of research.
1
Environmental Radiation Branch, Georgia Tech Research Institute, Georgia Institute of
Technology, Atlanta, GA
2
Department of Homeland Security, Washington, DC
3
Centennial High School, Boise, ID
4
National Institute of Standards and Technology, 100 Bureau Drive, Gaithersburg, MD
20899-8462
220
11. Quality Assurance 221
These categories are also addressed by other standards and guidance documents,
as mentioned in Section 11.1; however, the order, title, description, degree of
importance, and amount of detail given each topic differs among publications. For
a point-by-point comparison of different standards, see MARSSIM (EPA, 2000c,
Appendix K) or Appendix B of ASQC (1995).
11. Quality Assurance 223
The categories listed above are described in the following subsections. The
degree of detail is intended as guidance for management to formulate an effective
QA laboratory organization and to enable an auditor to overview the pertinent
components of the organization in a single document.
11.2.1. Organization
The QAP must contain a staffing chart that presents the laboratory organization
and describes the function and responsibility of each individual within the organi-
zational hierarchy (see Fig. 13.7). The resume of each staff member should also
be copied and filed in the plan. This information defines the chains of authority
and communication and records the qualifications for analyzing samples and re-
porting results so that responsibilities can be assigned to the appropriate analytical
and management staff persons. Every staff person is considered responsible for
performing duties and reporting in accord with the QAP. The information must be
kept current.
A QA officer must be designated to be directly responsible for managing the
QAP with respect to preparing the plan, supervising its application, and review-
ing its results. The QA officer must be competent in the fields of radioanalytical
chemistry and QA. He or she assists in reviewing analytical data reports to address
problems with analyses and radiation detection instruments and recommend prob-
lem solutions. The QA officer must be free to discuss results and findings directly
with supervisors, analysts, and operators. He or she must report directly to upper
management, with authority to recommend and initiate corrective action. In a large
laboratory, the QA officer supervises a separate and independent QA staff; in a
small laboratory, the QA officer may also fulfill other functions.
The QA officer and staff procure, store, and dispense QC samples, report QC
results, and evaluate the implication of these results for the analytical program.
The QA officer and staff prepare radionuclide sources for counting and solutions
for “spiking,” i.e., tracing the radionuclide of interest.
and reporting. Guidance must be presented for ancillary activities for worker safety,
laboratory security, and emergency response.
FIGURE 11.2. Summary SRM certificate for 133 Ba (Certificate available at https://srmors.
nist.gov/certificates/view cert2gif.cfm?certificate=4241C) (December 2005).
great many SRMs, all of which are included in this list; the 4000 series consists of
radionuclide SRMs. Click an SRM number—e.g., 4241c for the Barium-133 point
source—and then click on the “C” to view the certificate for specific radionuclide
or matrix. The full NIST certificate contains more than the information shown in
Fig. 11.2 and may be several pages of documentation and instructions on use.
Radionuclides that are used as comparison sources, e.g., for determining instru-
ment count rate stability and as tracers, must be of sufficient radiochemical purity
and activity to eliminate interference in counting and permit correction for decay,
but in most instances they need not have an accurately known disintegration rate.
The absolute count rate also is unimportant for energy calibration because only
the energy must be accurately known.
by the user. Each analyst enters all pertinent information in a personal notebook,
including the sample ID, what operations were performed (at least by reference to
a method or an instrument), calculations, and any pertinent observations. Ancillary
activities, such as QC measurements, reagent preparations, or instrument modifi-
cations, are also recorded. Unanticipated occurrences and findings also should be
recorded. Changes or corrections should not be erased or whited out, but rather
indicated by restating the information and drawing a line through the superseded
information.
Separate notebooks are associated with sample log-in and each instrument, such
as analytical balance and pH meter in the laboratory and radiation detector in the
counting room. Each use is recorded in terms of operator, date, sample ID, and
pertinent notes. All completed notebooks are filed permanently.
More recently, storage of information in computers has tended to superseded use
of notebooks. The computerized LIM system can include all such information and
relate it to sample ID and QC data presentation. Use of notebooks often continues
on a limited basis. This dual data record may be a lag in technology development
or reasoned record duplication in recognition of computer imperfections. In either
case, storage of all analytical records, including raw data printouts, whether hard
copy, disk, or tape, is mandatory, and the information must be retrievable, as
dictated by the QAP.
All but the fourth QC sample type are inserted in batches of routine samples. Al-
though inserting these QC samples without the knowledge of the operator (“blind”)
ensures that they are not treated with special care, this may not be possible in small
laboratories, where the analyst may recognize the QC samples. Distinctive features
are their radionuclide concentration (i.e., radionuclide concentrations are zero in
blank samples and elevated in spiked samples), salt content (blank samples may
have none), or the manner of their insertion into the sample flow. The last of the
above types also is used for external QC in the form of interlaboratory comparison
samples (see Section 11.2.11).
The results of QC sample analyses are evaluated with statistical tests (see Sec-
tion 10.5) and presented in tandem with the sample batch results to which they
correspond. Good QC results give confidence that the laboratory procedures and
personnel operated satisfactorily for that batch. Bad QC results show the need for
investigation and remeasurement. The QA officer keeps track of the laboratory’s
QC efforts, and the results are filed as specified by the QAP.
The fraction of QC samples depends on the number of samples per batch. If a
batch consists of about 20 samples, then as much as 20% of the sample load is
devoted to QC samples. Larger batches not only reduce the fractional cost of QC
analysis but also place more sample results at risk of a bad QC result.
The QA manual should have a directory for file locations of instrument manuals
and control records. Control records for each instrument should contain the date of
the test, the name of the tester, the results, and any pertinent observations. Among
these observations may be notes related to deviations and actions taken to correct
these deviations. The QAP should require that all instrument control records are
retained for the life of the instrument, and longer if subsequent reviews may occur.
Once radiation detection instruments are operational, they must be calibrated
with radionuclide standards for response in terms of counting efficiency and, for
some detectors, with radionuclides in terms of energy and peak resolution. The
counting efficiency is affected by radiation type and energy and by sample di-
mensions, density, and placement. Counting efficiency results obtained with one
standard should be confirmed by replicate measurements and with other standards.
Such additional measurements also can be used to estimate the standard deviation
of the efficiency values. In many instances, the efficiency values can be checked by
calculations based on the geometry and associated factors (see Section 8.2.1) and
by computer simulation. Curves or equations that represent trends in the variation
of efficiency with factors such as radiation energy and sample thickness (see Figs.
7.1 and 7.2) can provide efficiency values by interpolation where none were mea-
sured and call attention to efficiency measurements that are questionable because
of their difference from interpolated values.
The results are recorded in permanent form, defined by estimated uncertainty
values (see Section 10.3), and applied in the calculations that are used for re-
porting radioactivity levels. Calibrations are repeated at specified intervals. More
recent values are compared with earlier ones to resolve whether the new values are
within the uncertainty of the old values, should supersede them, or require further
measurements.
Background QC: At specified intervals, in many instances daily or for each
batch, the background count rate for each system must be measured. The count
rate is recorded and plotted on control charts, either by hand or by computer. The
mean value of the background is found by averaging at least 20 measurements.
The ±2σ and 3σ geviations are calculated from the individual and mean values,
and these multiple-standard-deviation lines are plotted on the control chart (see
Section 10.5.1). Once the control chart is established, each newly measured value
is recorded. The measurement should be repeated if it falls outside the 2σ band to
distinguish between a random event and an instrumental problem. Remedial action
with the detector or its environment is necessary if the repeated measurement is
beyond the 3σ band.
Instrumental QC: For a routinely operating radiation detector, a count rate com-
parison source is measured initially about 20 times to establish the mean value and
the standard deviation. A control chart with lines for the mean value and ±2σ and
3σ values is prepared. The source then is counted at frequent intervals—typically
daily or per batch—to demonstrate that the system is functioning properly. The
count rate is recorded and plotted on the control chart for the specific system,
as discussed above. This value is compared with the 2σ warning) limits and the
3σ (out-of-control) limits, and the procedure is repeated if the 2σ boundary is
232 Liz Thompson, Pamela Greenlaw, Linda Selvig, and Ken Inn
exceeded. Appropriate action must be taken if the measurement is beyond the out-
of-control levels. One option is to continue use of the detector but recalculate the
mean value and warning limits for the control chart. This approach is reasonable
only if the values have reached another stable level. The other option is instrument
repair.
Spectrometers have additional parameters for which response should be tracked.
These include the count rate in spectral energy regions of interest, peak channel
numbers at selected energies, difference in channel numbers between two spec-
ified energy peaks, peak energy resolution at specified energies, and the “peak
to Compton” ratio (see Section 8.3.4). A QC chart can be established for each
parameter by replicate measurements. This chart will display each data point as
well as lines at the mean value and at ±2σ and 3σ values.
The number of instrument QC checks typically represents about 5% of the total
number of measurements made for a set of samples. The QC data and records of
problems and responses should be kept accessible for the life of the instrument, and
possibly beyond, to indicate the extent of instrument reliability. Changes in values
may be examined for periodic patterns related to parameters such as temperature
or airborne radon concentrations to consider remedial measures.
1.6
1.4 Upper
Upper
1.2 middle
1
Ratio value
Lower
middle
0.8
Lower
0.6
0.4
0.2
0
0 20 40 60 80 100 120
Lab ID number
95th percentile) and lower middle (5th–15th percentile) limits define the warning
levels; depending on whether a result lies inside or outside the middle limit lines,
it is labeled Acceptable (A) or Acceptable With Warning (W), respectively. Other
parameters, such as the mean of all submitted values and the mean of acceptable
submitted values (all submitted values except outlying values), may be reported.
Any results in the W or N categories for a participating laboratory should trigger
a thorough QA effort within that laboratory to remedy the situation (see Chapter
12). The simplest causes are typographic, data transfer, and calculating error. A
second line of consideration is an error in analysis or counting that affects only
the intercomparison sample and can be checked by reanalysis. If these causes are
not applicable, the bad result reflects an inherent defect in analysis or counting
efficiency. This conclusion is supported if previous interlaboratory comparison
results were at or near the warning level. The results may hint at what problem to
seek by being biased low or high.
In view of the importance placed on agreement of results by a participating
laboratory with those reported by the testing laboratory, the testing laboratory must
take great care in preparing test samples and calculating its published radionuclide
concentration value and uncertainty. Nevertheless, errors have occurred in standard
certificates and testing laboratories. A significant difference between the mean of
acceptable values by many participating laboratories and the testing laboratory
value suggests that such an error may have occurred. For the values in Fig. 11.3,
for example, the ratio for the mean of all acceptable values of 1.07 indicates
reasonable agreement of results by the testing and tested laboratories.
The existence of a testing laboratory for interlaboratory comparison is so impor-
tant for assuring reliability of radioanalytical chemistry laboratories that it should
be supported by either a Federal agency or a group of laboratories. In the absence
of a testing laboratory, a cooperative venture may be arranged among several lab-
oratories to perform round-robin testing of a single procedure or a collaborative
study to develop and test standard methods. In these cases, results are compared
among several laboratories. Agreement among all participants within specified un-
certainty values usually satisfies the participants but disagreements cannot always
be resolved by attributing bad results to a given laboratory.
Even acceptable interlaboratory comparison results may require further QC ef-
forts at a laboratory if the test sample matrix is not identical to routine sample
matrices. Possible concerns are that the radionuclide in routine samples is in a dif-
ferent chemical form, that the radionuclide is at much lower concentration, or that
the matrix has different constituents. The analytical and counting procedure should
be examined to determine which factors are important, and whether additional QC
tests are indicated.
Acceptable interlaboratory comparison results reinforce acceptable internal QC
results by generating confidence in analytical and measurement processes. The
information usually is required by the client and for laboratory accreditation. Par-
ticipation in these comparisons should be at least on an annual basis for every
matrix type and radionuclide that is analyzed, and would be beneficial on a more
frequent cycle if the workforce turns over frequently.
236 Liz Thompson, Pamela Greenlaw, Linda Selvig, and Ken Inn
The audit plan, including the schedule, the personnel involved, procedures, and
checklist is the responsibility of the audit organization. A copy of this document
should be filed with the QAP to record the extent of the audit. All information,
comments, and recommendations provided by the audit team at the exit interview
and in the visit report should be added to the filed audit plan.
Every item in the report that requires a response must be addressed. Information
must be provided for action items to describe the outcome of the action taken. All
responses should be filed with the audit plan, together with resulting documents
such as certificates of accreditation or notification that all outstanding items have
been resolved.
determines what will be done with the resulting data after analysis by defining the
reporting format and the distribution and/or presentation of reports.
The QAPP is not to be confused with the QAP discussed in Section 11.2. The
QAPP is distinct from a laboratory QAP in that it defines an entire project, from
start to finish, rather than governing only the handling of samples once they reach
the laboratory. Like the QAP, the QAPP should be considered a work in progress
because results of analyses are to be used to revise the sampling and analysis
program for a more effective response to the information needs of the client.
for those samples that were collected was not satisfactory, given the geography
of the site.
r Acceptable control sites were not established so that the data collected from the
contaminated area could not be compared with those collected from a reliably
“clean” area.
r Too few replicate samples were taken to define the statistical uncertainty of
contaminant concentration at a given sampling site.
r Insufficient effort was devoted to interpreting the data in terms of its impact on
public health. While the effect of many of the individual chemicals on humans
was well documented, no research was conducted to determine the effect—the
synergy—of many toxic chemicals in concert.
The EPA clearly needed to develop technical guidelines for environmental mon-
itoring; this included the implementation of sampling and analytical protocols, and
the establishment of acceptable techniques for the documentation and presentation
of analytical results. The alternative was to continue to produce results that were
indefensible and expensive. The EPA realized that protocols were lacking, and
published some interim guidelines on QA (EPA, 1980b), but these had not yet
been widely implemented by the time the Love Canal situation came to light.
The problems faced by the EPA were symptomatic of the entire field of chemical
and radiochemical analysis at that time. Laboratories were increasingly scrutinized
and questioned: “How can we (the recipients of the data) be sure that the data are
what you (the generators of the data) say they are?” Formal proof of the data and
the process that produced it was required to answer this question. This need led
directly to the formulation of the DQO and QAPP systems.
Table 11.4 displays an expanded listing of costs within each category. Most
expenditure results from the time spent on QA measures, rather than from the
cost of materials. Time is of great value, and management should make efforts to
quantify the time spent on each of these QA elements. The analyst can help in this
effort by communicating with management regarding the allocation of time for
each element. For instance, it may be apparent to those in the laboratory that more
time should be spent on training or establishing a more efficient document filing
system. On the other hand, the staff may believe that some elements are consuming
too much time; perhaps too many reports are being prepared or report formats are
too elaborate. The dialogue of itself can be informative.
242 Liz Thompson, Pamela Greenlaw, Linda Selvig, and Ken Inn
Another aspect of QA made apparent by Table 11.4 is that the elements cat-
egorized as Internal and External failure are fewer, but more expensive. If QA
efforts in the Prevention and Appraisal categories are managed properly, the cost
of failure can be avoided.
The less quantifiable result of successful QA measures is the confidence they
awaken in the laboratory output of analytical results. Absent such trust, at best an
entire set of measurements may have to be repeated; at worst, the conclusion con-
cerning an entire monitoring program supported by the analyses may be rejected.
If these alternatives are unacceptable, then sufficient QA measures are worth the
effort.
Beyond reassuring regulators and stakeholders that analytical results reflect
reality, a thorough QA program described by a QAP manual has other benefits
(Ratliff, 2003):
r Some agencies, such as the NRC, require implementation of a QA system before
an operating license is granted.
r Accreditation by various groups requires that a QA system be in place. The QAP
manual serves as documentation that this has been done.
r A QAP manual serves as a historical record for the laboratory. It may be referred
to in times of investigation, or merely as an aid to memory.
r A QAP manual may be utilized as a promotional tool to emphasize ongoing
laboratory commitment to quality.
The QAP manual must, of course, reflect the efforts of effectively trained and re-
sponsible analysts and operators, QA staff, and laboratory management committed
to producing reliable results.
broadly applicable Supreme Court ruling that addresses this issue is related to
Daubert v. Merrill Dow (1993). This ruling makes clear that data and testimony
must meet a standard that the trial judge deems reasonable to admit the testimony,
and indicated the questions that would be used to establish their admissibility. As
stated in Daubert:
Many considerations will bear on the inquiry, including whether the theory or technique
in question can be (and has been) tested, whether it has been subjected to peer review
and publication, its known or potential error rate, and the existence and maintenance
of standards controlling its operation, and whether it has attracted widespread accep-
tance within a relevant scientific community. The inquiry is a flexible one, and its focus
must be solely on principles and methodology, not on the conclusions that they generate.
(http://www.daubertontheweb.com. Accessed August 27, 2005)
These are some of the same questions that are used to evaluate a standard method
(see Sections 11.2.8 and 6.5). Accurate record-keeping in the QAP ensures that
standard laboratory procedures are documented, and accurate record-keeping in a
laboratory notebook ensures that a written record exists to indicate that said pro-
cedures were indeed followed. Although it may seem exaggerated, the importance
of instituting QA procedures and keeping QA records cannot be overemphasized.
12
Methods Diagnostics
BERND KAHN
12.1. Introduction
Ideally, a laboratory processes every sample successfully and then correctly re-
ports its results. In practice, laboratory supervisors and staff devote much time
to examining the causes of failed analyses and attempting to avoid such failures
in the future. Such efforts in the laboratory to identify and then prevent or rem-
edy problems are usually not considered as a formal topic. This chapter proposes a
systematic approach to investigating the cause of a situation, i.e., diagnosis of labo-
ratory methodologies, to solve the current problem and prevent its recurrence. The
laboratory management structure must accommodate response to current prob-
lems in terms of staff, time, and budget. It must also take measures to prevent
future incidents by instituting and adhering to a quality assurance (QA) program,
as discussed in Chapter 11.
QA is the institutional system of methods diagnosis. Implementation of a QA
program formalizes analyst training, written procedures, and the careful review of
results. This formalization demands that each staff member perform the duties of
his/her position and take direct responsibility for the result. Within this framework,
the analyst takes responsibility for the analytical process, the detector operator for
the measurement process, and the supervisor for producing the results. The effect
is to create a laboratory environment that pervasively supports reliable analysis
and dependable reporting.
The quality control (QC) tests discussed in Sections 10.5 and 11.2.9 are integral
parts of QA designed to check results. Some QC measures are prompt indicators
that warn of problem occurrence at the time of analysis; others are delayed indi-
cators that require backtracking to find when a problem first arose. Control charts
for radiation detector operation are an example of a prompt indicator of reliability.
Records of deviations from the norm in an analysis or a measurement may also
be prompt indicators if immediately considered. Periodic blank, blind, and repli-
cate analyses, especially interlaboratory comparisons, are delayed indicators for
which results may not be available for days or weeks after a problem has arisen.
Review and assessment of compiled data are delayed indicators of information
quality.
244
12. Methods Diagnostics 245
The magnitude of the current effort devoted to QA (see Sections 10.5 and 11.4)
is a measure of the concern about problems that are encountered in radioanalyti-
cal chemistry. The goals of this chapter are twofold: first, to emphasize that QA
requirements are instituted not just to convince others of the dependability of the
laboratory but also to provide information for correcting problems and second, to
provide practical insight into laboratory problem solving.
cover all aspects of laboratory operation, as described in Section 11.1 for the QA
plan. Training is given to inculcate laboratory practices and methods application
as well as to review underlying chemical and physical principles. “Cold” (nonra-
dioactive) tests are performed in which the trainer first demonstrates the process,
then supervises every step to check analyst performance and eliminate faults, and
finally challenges the analyst to perform well independently. During this training,
the analyst should accept the responsibility to perform reliable work.
Analytical unreliability may arise when new analysts or operators, samples,
methods, or instruments are introduced. Loss of an analyst or operator and re-
placement by another is a common predictor of unreliability in analytical results,
as is the shift of an analyst or operator to a new assignment or a temporary one for
a vacation period or heavy workload. Thorough education and training can help
prevent bad analyses that are caused by unfamiliarity with a process.
12. Methods Diagnostics 247
In the same category lies the procurement of new instrumentation and the appli-
cation of new computer software for measurement, data processing, or reporting.
If the new instrument or software is radically different than that used previously,
a class may have to be scheduled to familiarize the staff with the particulars of
the equipment. Less drastic changes, such as the updating of a model or version,
probably can be handled by reading the manuals accompanying the new product.
Even subtle changes can cause problems. Examples are preparation of new
reagents and interactions by different persons in the chain of sample preparation,
analysis, and measurement. Here again, thorough training is the greatest preventive
measure in avoiding mistakes. Analysts should be notified in writing when the lab-
oratory makes any changes in reagent concentration or identity, and these changes
should be entered in the QA manual. Changes in the procedure for sample transfer
or alteration in the chain-of-command in the laboratory should be conveyed to the
staff so that everyone knows their new duties within the management framework.
addition, and amount of reagent may be specified too imprecisely to achieve con-
sistently the conditions for successful analysis.
Such inconsistency may be associated with unexpectedly wide fluctuations in
yield or decontamination factor, but may be difficult to assign to the imprecise
instruction at fault unless the analyst is sufficiently observant. Without such ob-
servation, a step-by-step test of the analysis may be required.
Occasional method failure can occur before or after chemical separation. A
variable fraction of the radionuclide may be lost during storage or initial treatment
before the carrier or tracer is added, or interchange between the carrier or tracer
and the radionuclide of interest may be incomplete. During counting, instrumental
effects such as quenching may be inaccurately assessed.
The first two tests measure the fraction of radionuclide loss to walls, while the
others only show relative retention in solution. A deficiency in these tests for
samples with low radionuclide content is that the obtained count rate often is too
low for precise determination of loss. A radioactive tracer solution can provide
higher count rates but may not represent the conditions in the actual sample.
Another common problem is the separation of samples into two phases: solids
and liquids for solids such as sediment, vegetation, and tissue; solution and sus-
pended solids for water. For such separation
r radionuclides distribute disproportionately and nonreproducibly between the
phases;
r aliquots taken for analysis do not represent the entire sample;
r sample weight at analysis differs from weight at collection.
Freezing samples or initially filtering liquids can reduce these problems. In the
absence of freezing, a preservative should be added to prevent decomposition of
biota and accumulation of decomposition products. The analyst should plan to
minimize the problem by prompt analysis and either mix the sample before taking
an aliquot or sample the liquid and solid phases in proportion.
One concern is partial or complete loss of the radionuclide before carrier or tracer
addition to monitor such loss. Carrier or tracer often are added after these con-
centration and dissolution processes to achieve complete interchange with the
radionuclide in solution (see Section 4.5.2). If, on the other hand, carrier or tracer
is added before dissolution, interchange of carrier or tracer with the radionuclide
of interest may not occur.
Possible loss may be anticipated and prevented by considering the known be-
havior of its chemical form. For example, if its form has relatively high vapor
pressure, either the processing temperature must be kept well below its boiling
point or its redox environment should be controlled to prevent formation of the
volatile form.
Sample nonuniformity must be avoided when a source is prepared for direct
radionuclide measurement by gamma-ray spectral analysis (see Section 7.5.1)
or gross alpha- and beta-particle counting. Some samples that were thoroughly
mixed just before measurement may remain mixed during the measurement, while
others separate into fractions by solids settling or gas emanation. For a mixture
of gas with liquid or solid, e.g., radioactive noble gases in solution or radon in
soil, the container must be filled completely to avoid a gas phase at the top and
sealed to avoid gas loss. Canning a sample is one way to retain gaseous radionu-
clides.
The first case strongly suggests a problem of recent origin: a reagent was badly
prepared, a fatal change was inadvertently introduced into the procedure, or a new
and incompletely trained analyst is at work. The second case suggests that at least
one of the separation steps is unreliable as written when the sample composi-
tion varies. Infrequent failure suggests a highly unusual sample matrix or analyst
aberration, as discussed in Section 12.2.3.
The simplest cause of failure—and the one most readily resolved—is some
temporary aberration or absentmindedness, which may be conceded by the analyst
or operator and can be confirmed by repeating the analysis or measurement. For a
more serious analyst problem, repeated analyses or measurements can be assigned
to a more experienced staff member.
Use of a reagent or carrier that is badly prepared, incorrectly labeled, in-
appropriate to the method, or no longer effective, can be pinpointed by first
repeating the analysis with a completely different set of reagents and, if success-
ful, then checking each original reagent in turn. In tandem with chemical failure,
254 Bernd Kahn
instrumental failure also must be considered. When a detector in the counting room
becomes suspect, use a second detector to check the earlier results from the first
detector.
Carrier and radionuclide tracer yields are convenient criteria of method efficacy.
Low yields suggest the action of interfering substances, but may also be caused by
analyst error, as discussed above. Yields that are unexpectedly high, especially in
excess of 100%, suggest the presence of the carrier or tracer in the original sample.
For example, some samples analyzed for radiostrontium to which strontium carrier
has been added may already contain a few milligrams of strontium. Excessively
high yields can be caused by the presence of relatively massive amounts of chemi-
cals that behave similarly to the radionuclide of interest, e.g., calcium or barium in
a process that purifies strontium in the sulfate or carbonate form, when the calcium
or barium is incompletely separated during purification. Imperfect exchange with
the radionuclide (see Section 4.5.2) is a condition that makes the yield irregular
and yield measurement, unreliable.
Sometimes the sample parameters simply do not mesh with the constraints of a
given method. Attempted analysis of a sample in which one or more constituents
exceed the amounts in the matrix for which the method was reported to be appli-
cable can result in failure. As discussed in Section 12.2.2, sample matrices such
as water, soil, and biota can be so variable that a procedure developed for one set
of samples may not be able to control interference from the different constituents
of the samples under consideration. In response, analyses can be performed with
smaller samples or more reagents.
Analysts should consider the method’s written statements of limitations with
regard to stable substances and the decontamination factor list for important in-
terfering radionuclides and compare those limitations with the characteristics of
the analyzed sample. For methods that do not have sufficiently defined limits,
the step at which the method fails will indicate the point of departure for a
study of possible interfering substances. This point can be identified by mea-
suring carrier or tracer levels at each step to find where serious loss occurs.
Tests of the inferred cause of failure include increasing reagents, inserting new
purification steps, or repeating existing steps. The presence of interfering ra-
dionuclides is shown by special measurements, e.g., for spectra or radioactive
decay.
The evaluation may show that the method is imprecisely written such that the
analyst views it differently than the writer. This oversight will become apparent
only in discussions with the analyst and must then be rectified by rewriting the
procedure.
to NIST (see Section 8.2.1). These sources are used to calibrate the detection sys-
tem for counting efficiency and, for a spectrometer, energy per channel and peak
resolution. Errors occur if
r a commercial supplier provides the wrong activity value for the standard solution;
r the user prepares the source badly from the solution;
r the source location relative to the detector is not accurately reproduced;
r the measurement is recorded inaccurately;
r a calculation is erroneous.
A correct calibration source may become less reliable with time by incorrect
radioactive decay adjustment, increase in the contaminant fraction, damage in
handling, or poorer statistical power due to radioactive decay.
In some cases, efficiency and energy calibrations are obtained from a standard
source of the radionuclide of interest that has a reported uncertainty value. In other
cases, these values are interpolations between data points obtained with several
measured radionuclide standards and plotted as function of energy so that the
uncertainty of curve fitting must be considered. Monte Carlo simulation has be-
come sufficiently accurate for energy calibration at the usually attained standard
deviation of 1–2% for reliability. This approach eliminates the need for interpo-
lation, but depends for accuracy on detailed information on the dimensions of
the detector (as in Fig. 8.9), the source, and detector-to-source configuration. The
simulation should be tested with at least one measurement each at low and high
gamma-ray energy to confirm the utilized information.
Calibration errors can occur if the instrument is incorrectly adjusted, settings
are accidentally changed during operation, or components fail. Care must be taken
that the computer code is appropriate to the geometric relation of sample and de-
tector, type of sample, radionuclide, and output that is to be calculated. Utilized
constants such as type of radiation, energy, half life, and decay or ingrowth fraction
must be checked (see Section 9.2). Other pertinent information that must be con-
firmed relates to the sample dimensions, density, weight, times of measurement,
counting period, and radiation background value. Detection system effects must
be addressed such as resolving time losses and coincidence summing, as discussed
in Section 8.2.1.
The problems listed above for calibration also apply to analysis of routine sam-
ples. Consideration must be given in these calculations to the consistency of the
periodic test sources, the sample collection and processing dates, and informa-
tion that relates amount measured to mass or volume collected. Also considered
must be numerical adjustments for drying or ashing the sample, concentrating the
radionuclide, and procedure yield.
Problems that may affect proportional counters relate to control settings, the
flowing gas, the detector window, and the automatic sample changer associated
with many detector systems. Wrong control settings can affect the operating volt-
age, pulse height discrimination, anticoincidence operation, and amplification. The
count rate is affected when the wrong gas is supplied, the gas flow rate deviates
from specifications, or the gas tank is shut off or empty. The detector window
258 Bernd Kahn
may be the wrong thickness or damaged. The sample changer may have the wrong
sample location information or otherwise may malfunction.
For LS counters, successful operation depends on functional dual PMTs and their
electronic circuits. Controls must be at appropriate settings. Some systems have
spectrometer and pulse shape and timing recognition circuits to restrict background
and evaluate the extent of quenching. Luminescence from ambient light sources
can be eliminated during a period for dark-adapting the sample before counting
it. Luminescence from contaminants must be quantified with associated detector
circuits and computer software or by comparison with prepared test samples.
Alpha-particle spectrometer systems often have multiple Si detectors operated
in parallel with a single spectrometer. A vacuum must be maintained in each de-
tector cell because attenuation in air causes low-energy tailing in each peak. The
adjustable detector-to-source distance affects the counting efficiency and resolu-
tion. Power supply stability is required when low-activity sources are measured
for long counting periods, e.g., several hundred thousand seconds. Results may be
based on only a few counts collected at a peak energy region over this extended
time. At higher count rates for which actual peaks can be viewed, knowledge of
the energies and intensities of multiple peaks for each radionuclide is required to
avoid misattribution of radionuclide activity.
A computer controls gamma-ray detector and spectrometer systems because of
their complexity. The Ge detectors require extremely stable voltage and cooling
by liquid nitrogen during operation to support high resolution. Failure to replace
coolant and freezing of coolant lines are occasional problems. Settings must be
determined initially and then maintained as long as QC measurements show ac-
ceptable ranges of the background, comparison source count rate at characteristic
peaks, energy calibration, and peak resolution. Error in data interpretation can
arise when a radionuclide that is counted at high efficiency emits gamma rays
in coincidence (see Section 9.4.5). Significant differences in density between the
standard and the unknown source can cause error.
1
Contribution by Douglas Van Cleef, ORTEC, Oak Ridge, TN 37830.
12. Methods Diagnostics 259
A standard deviation smaller than needed suggests that a briefer counting period,
a less sensitive detector, or a smaller sample can be applied. A standard deviation
smaller than expected indicates a detector problem.
To improve, i.e., reduce, the standard deviation of an analytical result, the com-
ponents of the standard deviation value (see Section 10.3.1) must be evaluated to
identify for reduction the major contributors to the uncertainty of the sample mea-
surement result. Multiple measurements of the gross count and the background
count provide direct measurements of the counting uncertainty. The measurements
may indicate that longer counting periods are needed or that the background must
be reduced or stabilized. Other contributors to analysis uncertainty, such as sample
260 Bernd Kahn
Management must support these activities to avoid analyst error, method break-
down, and loss of laboratory control. A methods evaluation and development group
in a large laboratory, or an individual in a small laboratory, should be responsi-
ble for systematically preparing methods for laboratory application, investigating
method breakdown, and evaluating replacement methods. A group member should
be available for discussing concerns by the analyst and supervisor. The group
should review the literature to identify alternative methods of consideration and to
consider potential problems discussed by others, such as MARLAP (EPA 2004).
An independent QA group or individual is needed to assure that a QA Plan is
prepared to include all methods and training requirements, among other items.
The group prepares QC samples, inserts them in the routine sample flow, reports
results, and assures that every instance of loss of laboratory control is evaluated
and remedied.
A data review group or individual (who may be the supervisor) must consider
every value reported by the laboratory in terms of internal consistency and the
pattern of past and nearby sample values and the uncertainty of these values. The
group or individual must consider missing and questionable values, and investigate
analysis and measurement problems to identify their causes.
13
Laboratory Design and Management
Principles
CHARLES PORTER1,3 AND GLENN MURPHY2
13.1. Introduction
The primary responsibilities in managing a radioanalytical chemistry laboratory
are to perform accurate analytical measurements and report the results in a timely
manner. Fine-tuning the design elements and management practices of the labora-
tory will invariably help a laboratory to meet those responsibilities. This chapter is
designed to give students an overview of what a modern radioanalytical laboratory
looks like and how it functions. The laboratory features discussed in this chap-
ter apply directly to laboratories processing environmental and bioassay samples
with low radionuclide content, but can be extrapolated to laboratory environments
where higher level samples are processed.
The early part of this chapter discusses the design and operating practices that
support analytical processes in an environment favorable for efficient work. The
design incorporates state of the art technologies in sample flow during processing,
hood design, ventilation systems, and waste disposal. The latter part of the chapter
addresses the staffing, costs, and attitudes appropriate for a reputable laboratory.
Management and operating considerations include personnel, operating costs, and
service orientation.
1
ELI Group, Inc., 3619 Wiley Rd., Montgomery, AL 36106
2
The Matrix Group & Associates, Inc., 118 Hidden Lake Dr., Hull, GA 20646
3
Charles R. Porter, ELI Group, Inc., 3619 Wiley Rd., Montgomery, AL 36106; email:
radlab@charter.net
261
262 Charles Porter and Glenn Murphy
Note: These hazard levels are not to be confused with the DOE classification of
nuclear waste into high-level, low-level, mixed low-level, transuranic and 11e(2)
byproduct material categories. These nuclear waste categories are established by
DOE Order 5820.2A, which can be viewed online at http://www.directives.doe.gov
(Dec. 2005). See DOE/EM (1997) for more information on nuclear waste. To
reiterate, waste hazard levels are different than laboratory hazard levels, although
the defining terminology is similar.
Recent experience at DOE sites has shown that most of the environmental sam-
ples collected today are levels C and D. Hence, the laboratory under consideration
in this chapter is designed for the analysis of levels C and D samples. These are
environmental or bioassay samples that contain radionuclides at low concentra-
tions, i.e., approximating levels of naturally occurring radionuclides. Samples at
levels A and B generally will be analyzed in on-site government laboratories for a
variety of reasons, i.e. transportation restriction, sample assay limitations, sample
security, and national security. In Table 13.1, the authors provide their suggested
activity levels to match the four categories identified by the DOE.
Table 13.1 is merely a guide. Each laboratory should develop specific quantity
limits. In some cases, the license under which the laboratory operates will specify
the quantity limits. For instance, the NRC issues specific radioactive material
licenses to facilities, and each license specifies the maximum quantity limit for a
given radionuclide. At government owned and operated sites, the DOE facilities do
13. Laboratory Design and Management Principles 263
not have quantity limits imposed by any regulatory condition or license, although
specific bases for operation may establish operational limits.
Whether of a regulatory or operational genesis, the quantities established are
based on several factors. These include the radionuclides to be processed; their
physical and chemical form; their decay scheme and half life; the education, skill
and training level of the analysts; the physical design of the laboratory; and the
radiological monitoring and controls imposed. Other conditions and factors also
may justify raising or lowering the quantity limits.
Regardless of the hazard level associated with the laboratory, several physi-
cal design features are held in common. All radioanalytical laboratory facilities
consist of a receiving and initial processing area (Area A), a set of individual
analytical laboratories and radiation detection (“counting room”) areas (Area B),
and the administrative and support facilities (Area C). These core elements con-
stitute the radioanalytical chemistry arm of the facility shown in Figs. 13.1–13.3.
The three figures fit together like a puzzle; they are displayed separately to facil-
itate easier viewing. This particular facility has an overall footprint of 261 ft. by
161 ft (80 × 49 m), or about 42,000 square feet (3,900 m2 ). A few aspects of the
complete mixed-waste laboratory are included in these figures while others are
omitted.
Numbers have been added to the layout in Figs. 13.1–13.3 to indicate sample
flow through the laboratory and help identify specific laboratory components.
Samples are received at the loading dock (1) and temporarily held in a storage area
(2) until ready for sample acceptance processing. The samples are moved to the
sorting area (3) where chain of custody is verified, sample containers are opened,
and the samples are inspected for acceptance. They are logged into the Laboratory
Information Management System (LIMS) (4) and bar codes are applied. One of
the key attributes of LIMS software is the ability to track the status of an individual
sample as it moves through the facility. Every sample bar code should be scanned
when the sample leaves one area and moves into another area. Each sample thus
can be readily located during the analytical process; a technician that moves the
sample to the wrong area is notified immediately by the LIMS.
The samples are held in the storage area (5) to await sample processing. Initial
preparations begin in area (6). Chemical purification and counting source prepara-
tion are performed in the laboratory areas (7) through (11) and (19) through (25);
the laboratory spaces have different sizes to handle various procedures. Prepara-
tion and counting of tritium samples are isolated in area (12), due to its unique
264 Charles Porter and Glenn Murphy
Loading
dock
Prescreening
counting room
FIGURE 13.1. Sample receiving and pre-screening portion of the laboratory (Area A).
preparation and counting requirements. The prepared samples are moved into
area (13) for counting. Areas (14) through (18) contain laboratory space to con-
duct analysis by various instrumental techniques, such as mass spectrometers (see
Chapter 17) and atomic absorption spectrometers. Areas (26) through (29) identify
administrative and support facilities.
The results of these analyses are transferred electronically (via the LIMS) to
workstations, where the data are validated and verified, and reports are prepared.
After the data are reported and accepted by the client, sample residues are moved to
the radioactive waste storage facility for disposal (see Section 13.4.5) or returned
to the client. The design features of individual laboratory areas are described in
Sections 13.3 to 13.6.
13. Laboratory Design and Management Principles
265
After placing the package in the fume hood, it is opened and the interior should be
checked for signs of internal damage, leakage, or removable surface contamination.
If a package does not meet acceptance criteria of the laboratory or Department of
Transportation, the package must be moved to an area designated for damaged or
leaking containers and processed to avoid further leakage. The laboratory director
must decide whether the package is rejected, repackaged and returned to sender,
or transferred to the waste storage building for disposal. Samples are scanned with
a high purity germanium (HpGe) detector plus spectrometer to check whether the
sample contents that emit gamma rays agree with the sample description.
This area also serves as the preliminary sample processing facility. It may contain
muffle furnaces, drying ovens, grinding mills, evaporation bays, and fume hoods
with dedicated exhaust lines. All exhaust from this area must be treated by pre-
filters, HEPA filters, and wet scrubbing prior to discharge to the environment. Each
muffle furnace and drying oven must be covered with a canopy exhaust system to
remove the heat and fumes generated from the heated samples.
The facility should have a separate room for storing radioactive standard and
stock solutions. This room usually is located near the sample receiving and pro-
cessing area. Radioactive standards and solutions must be kept separate from other
laboratory operations to prevent cross-contamination. The room should have cold
storage capabilities and lockable cabinets. It should be designed to the same spec-
ifications as other sample preparation rooms, with a fume hood and computer
access to permit dilution and other processing of radioactive standard reference
materials and stock solutions.
Finally, samples that have been pre-screened and are prepared for processing
can be moved through the air-lock doors into the analytical laboratory. The air-
lock doors separate the analytical laboratories and counting rooms from the other
parts of the facility. This minimizes contamination inadvertently transported to or
released in the sample preparation areas.
not be interchanged with other laboratories. The dishwasher internals and sinks
should be stainless steel to avoid deterioration from harsh chemicals left on the
glassware.
The row of laboratories should be backed up to a physical/mechanical support
chase that supplies water, electricity, vacuum, gas, air, and drain lines. The support
chase should be large enough to accommodate maintenance personnel and equip-
ment, but access should be restricted for other facility personnel. Some designs
incorporate an overhead mechanical room or eliminate the limited-access support
chase. All air supplies and ductwork should be overhead and directly accessible
from a maintenance penthouse that holds filter banks and scrubbers.
Individual laboratories should be designed to be shut down individually and
isolated from every other room in the event of an emergency. Valves for gas, water,
and vacuum lines should be accessible through the support chase or the mechanical
room penthouse. Electrical cut-off switches should be built-in at the door. These
design features prevent a spill or contamination incident from affecting operations
in other rooms or laboratories. Ideally, the contaminated laboratory is taken off-
line, decontaminated, serviced, and brought back on line with only minor impact
on other facility operations.
The laboratory design described in the following sub-sections considers effluent
and waste control. Its aims are minimizing cross-contamination and release of
radionuclides to the environment.
13.4.2. Airflow
The primary airflow design feature for the laboratory is single pass air. The air
is treated after it circulates through the laboratory and prior to its discharge to
the environment. Air is not re-circulated through the laboratory after treatment.
270 Charles Porter and Glenn Murphy
Each laboratory should have an average air turnover rate of 6–8 room volumes per
hour.
All air must be treated by filtration and/or scrubbing before it reaches the dis-
charge plenum. This treatment may occur in several stages. First, inexpensive
roughing filters (general collection efficiency of 50 percent or greater) are used
to remove large particles. Medium filters (collection efficiency of 80 percent or
greater) then collect smaller particle sizes, at a higher cost than the roughing fil-
ter. High efficiency particulate air (HEPA) filters (collection efficiency of 99.999
percent or greater) significantly reduce the amount of particulate materials above
1-micron particle size, but they cost considerably more. Without the roughing and
intermediate pre-filters, frequent HEPA filter changes would incur a prohibitive
cost.
Activated charcoal beds are used to capture volatile materials and delay the
discharge of radioactive noble gases. Wet-scrubbers are used to capture acids and
entrained liquids. In-line filters and wet-scrubbers should be located in the pent-
house of the facility. Scrubbers should have an average flow of 20–30 L/min and
should run continuously in any hood that evaporates mineral acids. The individual
hood washdown system should have a separate switch to wash down the entire
ductwork at the end of the day or at the end of a process.
Specific treatment combinations can be designed to match airborne contami-
nants from the physical or chemical operations, the type of sample (solid, liq-
uid, gas), and the quantity that is processed. For example, sample preparation
rooms where soil and vegetation are dried, ashed, ground and sieved require par-
ticle filter combinations but not charcoal beds and scrubbers. For treating lab-
oratory air, the multiple stage filter system should be based on the expected
maximum radionuclide concentration and airborne fraction of the processed
samples. Typical combinations include pre-filters, HEPA filters and charcoal
beds.
The individual laboratory is maintained under negative pressure relative to the
hallways and adjacent accessible areas. Negative pressure must be maintained
in each room to ensure that air is not drawn out of the fume hoods through the
room and into the corridor when a door is opened. The appropriate negative room
pressure is supported by sealing all wall joints and penetrations and by supplying
make-up air for hoods.
Each laboratory should have the capacity for four radiochemical fume hoods,
with one hood rated for perchloric acid. In Figs. 13.1 and 13.2, hoods appear as a
large X. Each fume hood must be provided with external make-up air to assist in
balancing the facility airflow. The make-up air should be at the same temperature
as the room air and provide from 50 to 75 percent of the airflow across the face of
the hood to maintain the hood face velocity at the required flow. This flow normally
is 85 to 100 linear feet per minute (26–31 meters per minute). Newer designs may
require different flow rates that should be verified by the facility industrial hygienist
or safety officer. Proper ductwork and blower motor sizes will keep the noise level
at the hood opening in the acceptable range of 85 to 92 decibels. The air discharged
from each fume hood should be wet-scrubbed and filtered, and then vented to the
13. Laboratory Design and Management Principles 271
environment through a single stack (see Fig. 13.4). Normal stack heights should
be roughly six times the stack diameter. In rooms without fume hoods, the room
air should be ducted into one of the existing plenums where filtration is in-line.
New designs or major upgrades should incorporate energy conservation tech-
niques. Design and construction of “green buildings” is gaining the support of
government agencies and can reduce utility costs. The planning architect should
include these techniques in the laboratory design. Examples of energy-conserving
design elements are:
r Heat exchangers on exhaust air streams
r Low-flow fume hoods (under development)
r Night set-back controls on hoods with closed sashes
r Improved control systems to regulate air volumes and temperature on HVAC
units
r A single air discharge plenum and fan for multiple hoods
r Water reuse and rainwater collection
r Task lighting to replace general lighting.
13.4.3. Liquids
The laboratory should be designed to discharge minimal amounts of radionuclides
to the sewer. Releases must meet the limits specified in the Code of Federal Reg-
ulations Title 10, Part 20, Table 3 (see Section 14.8 for an explanation of the CFR
titles) or DOE Order 5400.5. All liquids from the facility (including laboratories,
272 Charles Porter and Glenn Murphy
change and shower facilities, rest rooms, lunch rooms, and offices) are directly
piped into one of three holding tanks. Each tank should hold at least a 5-day vol-
ume of liquid waste, or 10,000 gallons (3.8 × 104 L) for the designed laboratory.
All fume hoods should have a water wash-down system. Each hood should be
washed at the end of the day to prevent acid build-up and the resulting degradation
of air ducts. The wash-down water is collected and re-circulated several times
prior to discharge to the holding tanks. The wash-down water usually is acidic
(pH <2) and needs to be neutralized daily. Eventually, solids will accumulate in
the wash-down water container and should be flushed to the holding tanks.
The holding tanks are located in a separate bermed area that can retain 110%
of the tank volume. The discharge lines from each laboratory (i.e., sinks, hoods,
dish washers) and the holding tanks are constructed of acid- and solvent-resistant
materials. The Teflon joints are heat-sealed to prevent leaks or separation. The
immediate vicinity of the tanks is surveyed at regular intervals to check for leaks
and associated radioactivity.
The holding tanks are monitored and agitated 24 hours a day, seven days a week.
One tank is actively receiving liquid waste; the second tank holds the liquid waste
off line for agitation, sampling, treatment, and discharge; and, the third tank is held
in reserve, in the event problems occur with the other two tanks. Tanks should be
discharged daily to prevent excessive accumulation; in no case should tanks be
held for more than 7 days prior to treatment and discharge.
The liquid waste in the second tank is treated by adjusting the pH to >5.2 and
then sampled for radionuclide analysis. If the concentration is acceptable, then
the tank may be discharged to the sanitary sewage system. If the concentration
exceeds release limits, several options may be considered. The three simplest are:
r Increase the dilution in the tank to reduce the concentration to an accept-
able level for discharge (NOTE: Dilution as a disposal option is not permitted
everywhere).
r Filter the tank contents to remove particulate matter for disposal as solid waste,
then re-analyze and, if the concentration is acceptable, discharge the liquid.
r Treat the liquid by flocculation to remove dissolved materials and then treat the
liquid as described in the second bullet.
The second column describes a smaller laboratory that can process up to 125
samples per month. Although some laboratories start by offering limited services
and then try to grow into a full-service facility, a minimum number of samples
must be processed monthly to produce a revenue stream that will sustain operation.
This minimum must be calculated for each type and quantity of sample, number
and types of analytes requested, and facility overhead.
The counting room houses the radiation detectors and counting systems indi-
cated in Table 13.2. The counting room depicted in Figure 13.5 is a blueprint for
an actual counting room. It is approximately 6 m × 9 m in the main body, with an
extra leg of about 1.5 m × 3 m for additional detectors. A “dutch” door (a door
divided horizontally, so that the upper portion can be opened without opening the
lower portion) is used to pass samples from the sample preparation room (num-
bered 6) into the counting room (numbered 13). This door, located to the left in
Figure 13.5, allows NO foot traffic. Authorized personnel may enter through the
other doors in Fig. 13.5.
This counting room has sufficient detector capacity to analyze 1000 samples
per month. The counting room should be located near the liquid nitrogen delivery
system, which is generally installed in the gas storage room. A separate loading
dock for the gas bottle storage room can be located directly behind the counting
room, as shown in Figures 13.2 and 13.5. The counting room requires liquid
nitrogen and P-10 counting gas, and possibly other gases that are used for chemical
analyses.
Germanium detectors require liquid nitrogen for cooling, and a minimum foot-
print of 1 m by 1 m. Each Dewar should be directly connected to the liquid nitrogen
feed line through a ball-cock on/off valve and special low-temperature piping. The
Dewar is mounted on a scale to monitor liquid nitrogen content. Detectors are
housed in a steel-clad lead shield to minimize the radiation background, which
13. Laboratory Design and Management Principles 275
includes ambient and nearby sources of radiation. The high-voltage and signal
cables can run overhead to each multi-channel analyzer. The low-level counting
room should have additional shielding in the walls that separate it from the second
room that is devoted to counting higher-level samples in the same area.
Because of the heavy weight of lead shields and steel stands, the counting room
should always be on the ground floor. Even with this stipulation, the floor should be
built of high density concrete with imbedded I-beams for increased support. Some
laboratories use a thick steel plate under the equipment footprint to distribute the
weight over a larger area.
Radiation detector shields must be of low-background materials (e.g., lead or
steel). In regions where radon gas intrusion and accumulation present a high-
background problem, the incoming air to the counting room may have to be passed
through charcoal beds to reduce the radon concentration, and room’s walls and floor
interfaces should be sealed to avoid radon emanation from the ground. The intended
location of the counting room should be checked to avoid elevated external radia-
tion levels of natural or man-made origin. Once the counting room is built, nearby
placement of facilities that will elevate external radiation levels must be prevented.
The counting room also houses alpha-particle spectrometry systems, which
require a vacuum line or pumps. These detectors can be placed anywhere within
the counting room and operated without interference from any elevated external
radiation background.
Low-background alpha/beta particle gas proportional counters can be placed
anywhere if a dedicated counting-gas cylinder is provided for each unit. If a central
supply of counting gas is used, the units should be placed in direct access to the
gas supply manifold. Liquid scintillation counters can be placed anywhere in the
counting room because they can be operated as stand-alone instruments.
All counting systems are linked directly to the laboratory computer system to
permit centralized control and for direct data transfer at the completion of counting.
The counting room requires one computer workstation to control each type of
counter, and additional terminals to support multiple users. Counting systems must
be connected to a stable uninterruptible power supply (UPS) that will provide
at least six hours of power in the event of power failure to the laboratory (see
Section 13.7.6). The UPS system should be mounted in the penthouse to minimize
equipment located on the counting room floor.
13.6. Storage
A busy laboratory utilizes large quantities of chemicals for sample preparation
and processing. The chemical storage facility should stand alone from the main
facility and be used to store and segregate hazardous materials such as flammables,
acids and bases (see Fig. 13.6). Chemicals in these different categories should
never be stored in the same room because inadvertent mixing could result in an
explosion or fire.
Radioactive waste also should be stored separately from the main facility to
avoid cross-contaminating other waste categories or supplies. Waste minimization
13. Laboratory Design and Management Principles 277
Organic Radioactive
Acid Solvent waste waste
place and direct competent scientists, engineers, and technicians in analytical and
technical positions, and support personnel in management, maintenance, secu-
rity, emergency response, health, and safety. A second task is to establish the
framework—facilities, equipment and supplies, budgets, and schedules—of labo-
ratory operation. A third task is to direct laboratory services to assure reliable results
that meet all client requirements and are consistent with government regulations.
month. In comparison, the smallest viable laboratory processes about 125 samples
per month.
Assurance of competent staff begins by preparing position descriptions that
specify the educational requirements, operational experience, and duties of the
laboratory personnel. The qualifications of the supervisors, analysts, and instru-
ment operators who are hired must match these descriptions. Each staff position
should have a set of core training specifications (e.g., radiation safety, chemical
safety, waste minimization and disposal, quality assurance) plus specific training in
its area of responsibility. Table 13.3 provides suggested position titles, educational
requirements, and general responsibilities for the optimal laboratory staff.
The radioanalytical chemistry laboratory has some specific needs, notably that
the chemists must have specific training and experience in radioanalytical chem-
istry. The skills of the radiation safety officer must match the duties described in
Section 14.3.2. The project managers should be fully versed in the operations and
capabilities of the laboratory to serve as client’s advocate with the analytical staff
and the laboratory’s liaison with the client.
features and (3) equipment. The HP staff oversees maintenance activities to min-
imize radiation exposure to personnel and spread of contamination. The HP staff
must first review the proposed activity, then survey the tasks while in progress, per-
form a final survey before resuming normal operations, and finally record radiation
exposure and radionuclide contamination levels.
Laboratories may employ either in-house or contractor janitorial staff. The
housekeeping staff needs initial training and routine re-training for areas that are
off-limits for routine cleaning or trash disposal. An individual laboratory may
have several trash receptacles to segregate normal trash, radioactive waste, and
hazardous chemical waste. Housekeeping staff should be trained to recognize the
various hazard symbols and placards and the appropriate response to each.
instruments can be obtained with maintenance and repair contracts at the time
of purchase. Operation of a small laboratory may at times be at the mercy of
a repair service for rapid turn-around in detector repair. A large laboratory may
consider comparing the cost and turn-around time of an internal repair staff with
contract repair. Rapid obsolescence of certain equipment may suggest a lease to
be more cost-effective than a purchase.
team, and the city or county governments. The EMP must itemize the potential
hazards at the facility and specific instructions for emergency response person-
nel. For example, the Fire Department may be advised not to enter a building
without respiratory protection or to perform fire suppression from a distance.
The Police Department may be advised to enter a building only when accompa-
nied by designated site personnel. The hospital may be advised that an incoming
radionuclide-contaminated patient should be treated in a designated radiation area
of the hospital. Communications must be concise, clear and rapid during an emer-
gency, and should be limited to selected personnel. Sections 14.5 and 14.6 provide
more information on emergency planning and accident response.
The EPA provides the computer code COMPLY for calculating the radiation
dose from release of radionuclides to air. The operator must use measurements of
stack velocity/volume flow rate and radionuclide amounts to calculate radiation
doses by COMPLY. A laboratory may qualify for an exemption from monitoring
requirements if specific radionuclide amounts do not exceed limiting values.
Clients need a copy of the laboratory license to verify permission to receive
samples that contain radionuclides. In turn, the laboratory should request a copy
of the client’s license. Samples not covered by the client’s license become the
responsibility of the laboratory that accepts them.
286 Charles Porter and Glenn Murphy
Facility design was discussed earlier in this chapter while the other topics are
addressed below.
with improved analytical methods. The laboratory cost list should be updated pe-
riodically to reflect such changes.
The basic cost of a radioanalytical chemical method is estimated from the person-
hours devoted to analysis from sample receipt to result submission. The associated
cost of supplies, supporting efforts, instrument use, and overhead is apportioned
among the various analyses that are performed during the same period. Personnel
cost is affected by the extent of processing needed to preserve and dissolve the sam-
ple and to remove interfering radionuclides. In essence, the simpler the process, the
lower the cost. The degree of precision and magnitude of detection level are other
cost determinants to the extent that they affect analytical skill, instrumentation,
and time requirements.
The quality assurance program, whether the normal effort or with special items
added by the client, is a major ancillary cost. The usual overhead costs of items
such as management, structures, maintenance, employee benefits, and taxes always
must be considered. Special costs may be invoked by studies to test or improve
methods of purification and measurement, explain unanticipated results, or answer
questions that arise during a project.
The above-cited factors suggest the potential benefits in achieving the most ap-
propriate cost by interacting with the client in preparing the Quality Assurance
Project Plan (see Section 11.3.2). Once the information needs are understood, an-
alytical efforts may be expanded or reduced with associated cost changes. Results
of initial screening analyses may suggest cost-reducing specification changes.
Scheduling the sample flow to the laboratory and the turnaround time for data
reporting may reduce cost by optimizing assignment of laboratory staff and
instruments.
13. Laboratory Design and Management Principles 289
A fine line separates operating the laboratory at capacity and being over-
extended; unanticipated events often cause a laboratory to cross the line between
full capacity processing and missed sample deadlines. Each project manager must
coordinate with the laboratory director and technical staff to ensure that no one
project or event detrimentally affects another client or overall laboratory operation.
If these disruptions occur consistently, one or more of the following changes can
be made:
r Increase staff levels.
r Add off-shift work.
r Install more counting instruments.
r Reduce the project load.
r Reorganize the sample-processing schedule with selected priorities.
13.9.5. Integrity
A reputation for integrity is vital to the successful operation of the radioanalytical
chemistry laboratory, as for any analytical facility. The client depends on the
validity of the reported analytical results in operating a program or performing
a task. In turn, the client provides information based on the analytical results to
various stakeholders, including workers, the public, and regulators, to demonstrate
safe operation or absence of excessive risk. Even a minor instance of unreliable
data can call into question entire reports of results and thus, the operation of the
client’s program.
Laboratory management must institute elaborate controls to guarantee data re-
liability (See Sections 10.5 and 10.6). Supervisors’ must overview the work of
analysts and instrument operators. A quality assurance program that typically in-
cludes comparisons with other laboratories, calibrations, quality control analyses
and measurements, and numerical comparisons of results should be instituted to
ensure accurate results. Periodic audits, questioning staff, and searching records
should confirm that all personnel are following written instructions. These reviews
have greatly expanded over the past several years in response to falsified results;
the cause of falsification ranged from minor errors to major incompetence and
even deceit.
13. Laboratory Design and Management Principles 291
Institutional controls can reassure the client and other users of data that any
problems will be found and corrected, but the main quality assurance effort must
be devoted to instilling integrity in laboratory staff. Managers and supervisors
must preach integrity at every opportunity; more importantly they must demon-
strate integrity in their practices and orders. While emphasizing production and
timeliness, they must place greater value on following specified procedures and
achieving correct analytical results. Chemists, instrument operators, and support
staff must be trained to follow procedures, report any deviations from protocol or
suspicions of unreliable instrumentation, and consider a questionable result to be
unacceptable. Integrity can be defined as always doing the right thing, especially
when no one appears to be watching. The laboratory director’s responsibility is
equating a correct result to “the right thing.”
A final aspect of integrity involves the interaction of the facility with the com-
munity in which it operates. The laboratory needs to maintain good will within the
community by being open, honest, and responsive to concerns as they are raised.
Most commonly, the concerns will center on laboratory waste and effluent, but may
also involve questions about the security of hazardous material and the soundness
of the facility’s emergency planning. Unless the facility is viewed as a good neigh-
bor and partner in the community, it risks becoming a target of animosity that will
inevitably damage its operations and morale.
14
Laboratory Safety∗
ARTHUR WICKMAN1,5 , PAUL SCHLUMPER2 , GLENN MURPHY,3
and LIZ THOMPSON4
14.1. Introduction
Laboratory instruments and the chemicals used in preparing samples can create
conditions that range from relatively benign to highly hazardous. If not identified
and controlled or eliminated, these hazards can expose the laboratory worker to
injuries and illnesses and cause damage to property and the environment. When
even an experienced radioanalytical chemist begins to work in a laboratory, it
behooves the supervisor to present clearly and with emphasis on safety the docu-
mented procedures of the laboratory and the behavioral practices that are expected
of the employee. The chemist should respond with attention and concern. Each
individual in the laboratory must be given and accept the primary responsibility
for safety. The safety culture must flow from management to laboratory worker
and must be embraced by each individual.
The reader was introduced to concern for a safe working environment and re-
sponsible behavior in the laboratory at his/her first experience with laboratory
activities. This chapter is intended to build on that initial training, by providing the
reader with a comprehensive discussion of safe laboratory practice from preven-
tion measures to emergency protocols. The discussions in this chapter focus on the
radioanalytical chemistry laboratory, whether it is a government, contractor, or aca-
demic research laboratory. Safety planning is discussed first to lay the groundwork
for more specific discussions to follow. Safety officers are identified and their duties
are described, together with the safety plans they administer. Workplace hazards,
∗
This text owes the genesis of its content to Dr. Isabel Fisenne, Department of Homeland
Security. The authors would like to thank Dr. Alena Paulenova of Oregon State University
for her careful review of the text.
1
Health Sciences Branch, Georgia Tech Research Institute, Georgia Institute of Technology,
Atlanta, GA 30332
2
Safety Engineering Branch, Georgia Tech Research Institute, Georgia Institute of Tech-
nology, Atlanta, GA 30332
3
The Matrix Group, Inc., 188 Hidden Lake Dr., Hull, GA 20646
4
Environmental Radiation Branch, Georgia Tech Research Institute, Georgia Institute of
Technology, Atlanta, GA 30332
5
Arthur Wickman, Georgia Institute of Technology, Atlanta, GA 30332-0837; email:
art.wickman@gtri.gatech.edu
292
14. Laboratory Safety 293
r making the immediate supervisor aware of the hazards and their possible controls
for proposed methods; and
r communicating constructively with safety staff by identifying concerns and re-
sponding to guidance.
Hazard identification and control are important aspects of safety in a laboratory.
Most hazards in a laboratory environment involve either unsafe conditions or
behavior. Conditions can be controlled through proper analysis and inspection of
the work environment, and implementation of controls to reduce or eliminate the
exposure to these hazards. A formal job hazard analysis, where individual tasks
are observed, broken down into their individual components, and analyzed for
existing and potential hazards is necessary for hazard identification and corrective
action. This activity must be followed by periodic formal inspections and hazard
assessments.
Unsafe behavior may be difficult to predict and control. Laboratory worker train-
ing is thus an important component of a safety and health management system.
Before they even set foot in the laboratory, workers must be aware of the potential
hazards to which they might be exposed, the protective measures they should take
regarding those hazards, and the management procedures and policies regarding
safety and health. Training is usually conducted for new workers as an orienta-
tion and then periodically in refresher courses. Training also is required whenever
a change occurs in the laboratory environment related to procedures, materials,
or equipment. Finally, training is an important aspect of corrective action, should
observation indicate that a laboratory worker is performing a task unsafely. Individ-
uals who demonstrate unsafe work practices require additional training to ensure
they have proper knowledge of safe work practices. If this remedy is ineffective,
disciplinary action must be instituted.
To close the loop in the safety and health management system, periodic assess-
ment and feedback are necessary. Indicators should be chosen that can assess the
overall performance of the laboratory with respect to safety and health. Whenever
possible, “leading” indicators such as behavioral observations should be measured
and reviewed, as well as “trailing” indicators such as the type and number of in-
juries and illnesses and loss of working time. The purpose of this assessment is to
determine the overall effectiveness of the safety and health management system
and to correct any areas of deficiency.
This safety and health management system constitutes a safety framework with
elements that should be implemented in all laboratories. More specific tools to
delegate the responsibility of safety management in the radioanalytical chemistry
laboratory are described in the following section.
Additionally, the CHP may contain the emergency plans for the laboratory (see
Section 14.5).
The CHP must be administered by a CHO, an individual qualified by training
and experience to handle safely the chemicals and processes in the laboratory who
also has the authority to take corrective actions when needed. The CHO must be
familiar with all safety and environmental regulations that apply to the laboratory.
The CHO must understand the health and safety hazards of the chemicals in use.
The CHO must be aware of any biological monitoring required for employees
or students who use regulated chemicals. To be effective, the CHO must have
the support of managers and administrators, and must have a clear mandate to
296 Arthur Wickman et al.
Each worker in the laboratory must undergo initial radiation safety train-
ing that prepares him/her to work with radioactive materials. Refresher training
must brief the employee about changes to operating procedures and regulatory
requirements.
Radiation safety staffing levels may vary widely among facilities on the basis
of facility size, radioactivity levels, and number of samples processed. Each facil-
ity is required to have an RSO that meets the education, training and experience
requirements by the NRC or DOE. The DOE protocol is for government labora-
tories while NRC and Agreement State licenses control other radionuclide-using
entities, including commercial and academic laboratories.
The NRC and Agreement States have specific license requirements for RSO
education, training and experience. To practice this specialty, the RSO must meet
the requirements of 10 CFR 35.900 series Subpart J—Training and Experience Re-
quirements. Three different sets of experience can qualify an individual to become
an RSO:
14. Laboratory Safety 297
The DOE does not use the license concept, so there are no formal require-
ments to become a Radiological Control Manager (RCM)—the equivalent of
an RSO. Regardless of title and regulatory mechanism, the RSO or RCM is re-
sponsible for implementing a safety program for the use and control of radioac-
tive materials in the laboratory. The RSO/RCM is responsible for the following
activities:
r maintaining the facility license (for the NRC or the Agreement State) or the
Radiological Protection Program Plan (for the DOE);
r procuring, receiving, and delivering the radioactive material to the individual
user;
r training the user in routine handling and emergency response procedures;
r implementing an internal and external dosimetry program;
r performing routine surveys in radioactive materials use areas;
r disposing of radioactive waste; and
r controlling any other item pertinent to radiation protection (for example, a posi-
tion description may specify “5% of time spent on other duties as assigned”).
The RSO is responsible for working with the radiation laboratory worker in
planning any program that may result in radiation exposure to persons and ra-
dionuclide contamination of the work place or the environment. The radioanalyt-
ical chemist should be able to depend on the RSO for guidance in minimizing
the potential for such exposure and contamination. In brief, radioactive material
must be stored, handled, and discarded separately from and differently than the
usual laboratory reagents. The RSO must inform the regulator, management, and
the worker of personnel radiation exposure and any unusual radiation exposure
conditions.
compatibility, and stored in a secure area equipped with local exhaust ventilation.
The storage area should be continuously ventilated and under slightly negative
air pressure, well lighted, supervised, and capable of being locked to prevent un-
authorized access. Chemicals should be stored at or below eye level. Large bottles
should be no more than two feet from ground level, and not stored on the aisles
or floor. Mineral acids should be stored on acid-resistant trays, separately from
flammable and combustible materials. Acid-sensitive materials, such as cyanides
and sulfides, should be protected from contact with acids. At least annually, the
inventory of stored chemicals should be examined to identify evidence of con-
tainer deterioration or damage and to identify any chemicals to be replaced based
on length of service. Chemicals that are no longer needed should be disposed of
properly.
Chemicals taken from the storage area for use at the laboratory bench should be
transported in durable secondary containers of sufficient size and composition to
retain any spill of the chemical. Preparation or repackaging of chemicals should
take place in a designated area of the laboratory, outside of the main storage
area. Chemicals stored at the laboratory bench should be limited to the amounts
necessary for current application. Bench chemicals should be stored in an orderly
manner, away from sunlight and external sources of heat. Containers should be
protected from falling by placing them well away from the edges of counter tops.
Containers may not be stored on the floor. Chemicals may not be removed from
the laboratory without permission. Unnecessary items should not be stored at the
laboratory bench.
Material Safety Data Sheets (MSDS) are compiled and available for all chem-
icals. These documents explain the inhalation and contact health risks, as well
as the flammability and toxicity of the chemical they describe. They also outline
special handling and storage considerations. When handling a new chemical, it is
prudent to examine its accompanying MSDS.
With or without an MSDS, however, caution in chemical handling should be
observed. As a general guideline, worker and student exposures to laboratory
chemicals should be kept to a minimum. Because so many laboratory chemicals
are hazardous to humans in some way, conservative risk assessment should be
employed. Persons in the laboratory should assume that personal protection is re-
quired whenever they are working with chemicals. Chemicals of unknown toxicity
initially should be treated as toxic with respect to exposure during work performed
in the laboratory. For work with chemicals of known toxicity, appropriate precau-
tions should be taken. When working with mixtures of chemicals, the risk of the
mixture as a whole should be assumed at least to equal the risk for the most toxic
component of the mixture.
Skin contact with chemicals should be avoided. Personnel must wash all areas
of exposed skin prior to leaving the laboratory. Mouth suction for pipettes or for
starting siphons is not allowed. Prohibitions on eating, drinking, smoking, gum
chewing, or the application of cosmetics in the laboratory should be posted, and the
rules concerning these activities should be enforced by the supervisor. Hands and
face should be thoroughly washed prior to eating or drinking. Protective aprons,
14. Laboratory Safety 301
jackets, or coats should be removed prior to entry into areas where food may be
consumed.
r Lack of a permanent and continuous ground: This hazard can result from a
broken or removed ground plug or a wiring problem at an electrical receptacle.
With an open ground, a fault in equipment can result in the path of electricity
through a laboratory worker instead of the desired ground path. Laboratory
inspections must ensure that ground pins are correctly in place for equipment
designed to have them and that receptacles are tested periodically for correct
wiring.
r Unused openings or missing covers in electrical equipment or boxes: Electri-
cal equipment can become damaged or neglected over time, with exposure to
energized electrical parts through unused openings such as knock-out closures
or missing covers. Exposure to energized electrical parts, especially parts op-
erating at over 50 Volts, must be prevented. Electrical parts should be serviced
or maintained only after de-energizing the circuit and following the established
lockout/tagout procedure. When de-energizing is not feasible, only qualified
electricians operating after training and with personal protective equipment and
insulating tools should attempt to work in the vicinity of energized electrical
parts.
r Damaged insulation: Damaged insulation must be repaired immediately by a
qualified electrician. Simply covering insulation with electrical tape will most
likely not be sufficient to protect individuals from contact. The repaired section
must have the same mechanical strength and insulating properties as the original
wiring.
r Wet or damp environments: Water can be a good conductor of electricity, hence
working with electrical equipment in a wet environment can be a hazardous
activity. Electrical equipment should be isolated from moisture. Electrical re-
ceptacles in wet or damp environments must be designed for this type of envi-
ronment, with ground-fault-circuit-interrupter (GFCI) protection. This type of
receptacle should be tested periodically in accordance with the manufacturer’s
recommendations.
r Circuit overload: Electrical equipment should not be utilized in a manner not
intended by its manufacturer. Overloaded wiring and receptacles can damage
the equipment and result in electrical shock, electrocution, and fire. Equipment
installers and users should always understand the intended use of the equipment
and follow manufacturer’s recommendations.
306 Arthur Wickman et al.
14.4.10. Monitoring
In the radioanalytical chemistry laboratory, higher- and lower-level radiation areas
must be clearly designated. When specifying the radiation level of an area, signs
akin to those in Figure 14.1 would be posted. There are many variations of this
type of sign, as they are placed anywhere radiation is used or found—including
hospitals, laboratories, chemical plants and waste clean-up sites. Each of these
locations has different radiation concerns, and a different audience viewing the
signage. What all signs will have in common, however, is the tri-foil. Depicted in
Figure 14.2, the tri-foil is the international symbol for radiation. The symbol can
be magenta or black, on a yellow (or sometimes white) background.
Access between the two appropriately labeled radiation areas should be re-
stricted, in order to avoid contaminating the low-level area. In laboratories where
both “hot” and “cold” operations are being performed in a small amount of space,
it is important to set up well-defined and clearly labeled radioactive material work
stations. In this way, radionuclide analysis is confined so that its impact on the
surrounding laboratory is minimal. The presence of radiation throughout the lab-
oratory should be monitored as discussed below.
Because the five senses are useless for detecting radiation, each facility must
have readily available portable radiation detection instruments. These instruments
should be selected to detect and quantify the three basic types of radiation: alpha
particles, beta particles, and gamma rays, as discussed in Chapter 2. In some
cases, neutron detectors may be required. The RSO/RCM generally is responsible
for calibrating the instruments at selected intervals, typically six months. The
individual user is responsible for daily operational and source checks prior to each
use.
Each radiation safety program should have an established survey program to
monitor the radiation and contamination levels in the laboratory so that the in-
dividual user can maintain a safe working environment. The measurements can
identify and measure individual radionuclides or measure radiation dose. The
dosimetry program should monitor exposures to individuals from external radi-
ation or internal—ingested or inhaled—radionuclides, and also in areas of the
laboratory.
External exposure levels generally are monitored with a thermoluminescent
dosimeters (earlier, with film badges) or electronic dosimeters. Either method
provides individual exposure information for a worker or a location in terms of ex-
posure over a pre-set monitoring period, after which the dosimeters are exchanged
for new ones. The used dosimeter is read to determine the radiation exposure dur-
ing the interval. In laboratory areas that process samples with higher radionuclide
concentrations, wrist and ring badges or direct-reading radiation dosimeters also
may be worn, but they usually are unnecessary when processing environmental
or bioassay samples. An exception may be made for persons who receive mixed
samples at the dock or handle radioactivity standards before dilution.
The internal dosimetry program generally relies on in vitro (outside) or in
vivo (inside) monitoring capabilities. In vitro programs use an external radiation
308 Arthur Wickman et al.
detection system to monitor radiation being emitted from inside the body and de-
tected on the outside of the body. In vivo programs use a sample from the body
(usually urine, but sometimes feces, blood, hair, or breath) to monitor contamina-
tion within the body.
Urine samples for bioassay are collected from workers who may have inhaled
or ingested radionuclides or hazardous chemicals, either while routinely handling
samples with higher radionuclide or chemical concentrations, or after accidental
exposure. The samples are analyzed for the radionuclides or chemicals that may
have been taken in by the worker and the measured concentrations are used to
calculate radiation exposures or to compare to radionuclide or chemical concen-
tration limits. The frequency of sampling for bioassay depends on the occurrences
of exposure and the rate of turnover of the radionuclides or chemicals in the body.
A large facility usually has an infirmary with medical staff for treating injuries
and performing physical examinations at the beginning and end of employment,
and at regular intervals. The type and frequency of routine physicals may depend on
the probability of adverse effects associated with the work environment. Employee
medical records should be maintained to track the occurrence or absence of health
effects due to the work environment.
Management must promptly report to the worker and the regulatory agency
any unusually elevated radiation exposures—both external and internal—or con-
centrations of hazardous chemicals, and institute changes in the work environ-
ment to prevent continuing exposures that reach exposure limits. The IH/CHO or
RSO/RCM staff must be instrumental in achieving such changes. The exposure
must be recorded with a description of the causes and instituted remedies.
must be free of any device or alarm that could restrict emergency use of the exit
route if the device or alarm fails.
r Exit routes must support the maximum permitted occupant load of each floor
and the capacity (width) of an exit must not decrease in the direction of travel.
r The minimum exit width is 28 inches (0.7 m), but should be wider in most
circumstances.
r Exit routes must be arranged so that occupants will not have to travel toward a
high-hazard area (such as a flammable liquid storage room), unless the path of
travel is effectively shielded from the high-hazard area by physical barriers.
r Exit routes must be free and unobstructed.
r Exit routes must be adequately lighted so that an occupant with normal vision
can see along the exit route. The lighting must be impervious to power outage, so
backup lighting must be on an independent power supply (generator or batteries).
Automatic illumination by these lights should be tested periodically.
r Each exit must be clearly visible and marked by a sign reading “Exit”.
r Laboratories must install and maintain an operable alarm system that has a
distinctive signal to warn occupants of fire or other emergencies.
extinguisher that contains one of several added chemical components that in-
crease permeation of the stream into the fire), foam or multipurpose dry chemi-
cal. The numerical rating refers to the amount of water or other fire extinguishing
agent that the extinguisher holds.
r Class B fires are fueled by flammable liquids and gases, or grease. Class B
extinguishers may be Halon 1301, Halon 1211, carbon dioxide, dry chemicals,
or foam. Here, the numerical rating indicates the approximate number of square
feet of Class B fire the average user can be expected to extinguish using that
extinguisher.
r Class C fires are electrical fires, and may be extinguished with Halon 1301, Halon
1211, carbon dioxide, or dry chemical fire extinguishers. Class C extinguishers
have no additional numerical rating.
r Class D fires involve flammable metals, such as magnesium, sodium, potassium,
titanium and zirconium. Extinguishers that contain water, gas, or certain dry
chemicals cannot extinguish or control this type of fire and may actually fuel it.
Class D fire extinguishers utilize agents including Foundry flux, Lith-X powder,
TMB liquid, pyromet powder, TEC powder, dry talc, dry graphite powder, dry
sand, dry sodium chloride, dry soda ash, lithium chloride, zirconium silicate,
and dry dolomite. There is usually no numerical rating for these extinguishers.
An excellent resource for fire safety and fire extinguishers can be found at
http://www.hanford.gov/fire/safety/extingrs.htm. This website is maintained by
the Hanford Fire Department in Richland, WA, which oversees fire safety and
operations for the DOE’s Hanford facility.
The old symbols for fire extinguishers geared for each type of fire are shown in
Figure 14.3. Figure 14.4 contains the new symbols. Note that the new pictograms
have no symbol for Class D extinguishers, which are often specific to the type of
metal burning.
Among Classes A–C, crossover in the active firefighting agent results in some
extinguishers that are Class A/B (for example, foam) or Class A/B/C (some dry
chemical extinguishers). All fire extinguishers are labeled with their class hazard
rating to insure that the extinguishers are properly used. Before attempting to put
out a fire, check the label to confirm that the extinguisher is appropriate. In the old
format, the symbol for each Class of fire that the extinguisher is designed to address
will be on the extinguisher bottle. In the new format, all three pictograms are on
the extinguisher bottle, with a red strike through any Class that the extinguisher
A B C D
Ordinary Flammable Electrical Combustible
Combustibles Liquids Equipment Metals
FIGURE 14.3. The old symbols, representing Class A, B, C, and D fire extinguishers.
Online at: http://www.hanford.gov/fire/safety/extingrs.htm. (Dec. 2005).
312 Arthur Wickman et al.
Electrical
Equipment
FIGURE 14.4. The new pictograms, representing Class A, B, and C fire extinguisher. There
is no pictogram for Class D extinguishers. Online at:
http://www.hanford.gov/fire/safety/extingrs.htm. (Dec. 2005).
is NOT designed to address. One should remember that water is only useful in
putting out Class A fires. A mixed-use extinguisher will never contain water, and
the user should never improperly use a water-based extinguisher (Class A) on any
other type of fire.
Fire extinguishers must be inspected routinely to insure that they are in operating
condition. Employees must be trained in the use of fire extinguishers unless im-
mediate evacuation is the facility policy. New employees should check the CHP to
determine their responsibility (some facilities may designate only some employees
as fire officers), and inquire about the laboratory fire safety training program.
In addition to fire extinguishers, a sprinkler system in the ceiling is almost always
available. These are, of course, useful in extinguishing Class A fires. However,
these sprinklers activate automatically, in response to the presence of smoke and/or
debris in the laboratory air. The automatic addition of water to a class B, C or D fire
would be counter-productive and should be avoided. Therefore, the fire-producing
potential of material in each room should be evaluated, prior to installation of the
sprinkler system.
1. Stop a fire, accidental spill, or gas leak at its source, if possible. This means ex-
tinguishing or removing the flammable source of the fire, stopping the chemical
leak, or turning off the gas. If these activities are too dangerous, do not attempt
them; isolate the room and proceed to the next step.
2. Raise the alarm for those in the area to evacuate.
3. Assess injuries and assist the injured, removing them from the area for medical
attention, if possible.
4. Notify the designated response group promptly and provide as much informa-
tion as can be obtained in a brief period. The designated responder will contact
the fire department, police, and HazMat, based on the information provided. Be
as calm and accurate as possible when relaying the details of the accident.
HazMat (see Section 14.9) publishes a guide to aid first responders in (1) quickly
identifying the specific or generic classification of the material(s) involved in the
incident, and (2) protecting themselves and the general public during this initial
response phase of the incident. This publication is called the Emergency Response
Handbook, and can be downloaded online. The handbook is updated every three
to four years to accommodate new products and technology. The next version is
scheduled for 2008.
314 Arthur Wickman et al.
15.1. Introduction
Many radiochemical analyses consist of a series of identical steps with little or
no variation from sample to sample. Operator fatigue in the execution of repeti-
tive steps leads to increased process variability and execution errors. Such tasks
are attractive candidates for automation because they can be made more efficient,
consistent, and cost-effective. This trend comes to radiochemistry at a time when
schoolchildren use computers regularly, typing or dictating homework assignments
to a word processor, with other tools such as “spell check” and “word count” to
complete the task. This digital age has also ushered in complex computers predict-
ing molecular, biological, and other highly intricate processes. While an automated
device can perform in a moment the manual computations of a scientist’s lifetime,
however, the same computer would continue computing in error the equivalent of
a thousand lifetimes, while a scientist would perceive the error and stop! Thus,
despite these tremendous information age advances, the automation of physical
and chemical processes must be done thoughtfully to avoid repetition of errors in
quality, lapses in safety, and other pitfalls that a scientist would instantly perceive
and halt.
Automated systems have several potential advantages over classical technical
labor. Analytical functions are reproduced on a programmed schedule, reducing the
turnaround time of the samples and enabling a more reliable estimate of analytical
costs. In addition to the accuracy of the equipment, costs are readily estimated,
the process is easier to scale up or down, and process parameters can be more
easily varied in an understandable way. Automated systems can produce data
from worldwide locations and may be readily standardized and compared. Manual
operation remains preferable for procedures that require considerable judgment in
the selection of alternatives.
Aspects of chemistry for which automation has been a popular and effective im-
provement over manual processes include continuous-flow separation processes,
instrumental analysis, and continuous monitoring. Radiometric measurements are
a good example, as numerous samples are counted, one after the other, each for a
318
15. Automated Laboratory and Field Radionuclide Analysis Systems 319
specified time period. Automation can replace manual placement of the samples
for measurement, with a large number of sequential measurements possible with-
out operator intervention. In fact, the sample collection process can be automated
in some cases and then mated to an automated measurement system.
Most radioanalytical chemistry processes may be automated to some degree,
depending on the type of physical manipulation the sample requires and the com-
plexity of the chemical process. In some cases, the essential mechanical manipula-
tion can be achieved with commercially available components originally designed
to automate manufacturing operations. For instance, the pick-and-place automa-
tion developed for the semiconductor and other industries can be readily adapted
to sample measurement. In other cases, automation relies on specially designed
commercial hardware and software tools. Examples of special computer-controlled
automated systems developed for atom-at-a-time detection and analysis of actinide
and transactinide element isotopes with half-lives as short as a few seconds are
shown in Figs. 16.4–16.6. Such techniques have been used to perform the very
first studies of the chemical properties of 104 Rf through 108 Hs.
Comprehensive knowledge of the radioanalytical chemistry procedure is needed
to devise successful mechanical substitutes for the manual steps of the process and
to ensure that the controlling software responds appropriately and predictably. As
might be expected, handling special cases and identifying problematic samples are
key elements. Examples of automated sampling, analytical processing, and mea-
surement are presented in this chapter to illustrate what may be achieved by com-
bining measurement and engineering expertise. These are illustrative of the prin-
ciples of automation and are arranged in order of increasing complexity. While the
engineering aspects of automation are beyond the scope of a radioanalytical chem-
istry text, it is useful to explore some of the ideas needed for a successful experience.
Automation adds value to the results recorded. The IMS network provides a vast
database of signals recorded under essentially identical conditions and reflecting
regional variations in backgrounds. This allows temporal filtering to eliminate
false alerts based on the frequency of similar signatures in the past and detailed
characterization of system state-of-health, thus enabling higher confidence in ra-
dionuclide spectral data. The remainder of Section 15.1 examines aspects of control
and data handling for automated systems in more detail.
graphical feedback. The most favorable human–computer interface for highly in-
teractive operations is a graphical user interface. These may require larger commu-
nications bandwidth but provide much greater speed of user interaction. Existing
schemes for supporting such interfaces over low-bandwidth connections include
methods for compressing X-Windows protocol traffic. Another solution is to op-
erate a graphical tool on the operator’s end of the connection, with only sparse
command-response transmissions with the instrument. Well-designed graphical
user interfaces require more software development, balanced by greatly reduced
user training.
The broad categories of remote control activities are
1. system inspection commands, or data requests,
2. parameter changes and commands, including Stop and Reboot,
3. interactive sessions, such as energy calibration, and
4. software uploads.
These categories are listed in order of increasing intrusiveness of the remote con-
trol. Reckless system modification can easily result in a disabled instrument con-
figuration with, worse yet, no remote way to recover normal operation. Even in the
early days of automation, it is surprising how often this circumstance resulted in a
technician making travel arrangements. The first and second categories can be han-
dled via email commands or a low-bandwidth graphical user interface command
system. The third category requires a higher bandwidth interface and effectively
must either transmit graphics or use interactive graphical tools located on the oper-
ator’s console. The last category can be implemented most easily with a standard
and secure file transfer protocol.
10 000
140B a
1000
100
10
dpm/kSCM
0.1
DL
0.01
DETONATIONS
0.001 USA
USSR
> 10 Mt CHINA
1-10 Mt Chernobyl
0 1-1 Mt
0 02-0.1 Mt
61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89
Many versions of this basic approach exist, the most significant variation being
whether matrix inversion is used to connect a library of radionuclides to observed
peak intensities or whether a list of energies of interest is used to make key cal-
culations. Thus, the two main types today are matrix inversion and list directed.
Automated spectral analysis software is available from commercial and academic
sources with a mix of national and international quality certifications, special-
ized capabilities, and user control. Programs of this type can handle thousands of
automated analyses per day and run on most types of computers.
324 Harry S. Miley and Craig Edward Aalseth
FIGURE 15.2. The R-TARAC, configured for under-wing pod use (top) and for fuselage- or
automobile-based use (bottom). U.S. Patent 6,184,531. (By permission of Pacific Northwest
National Laboratory)
15. Automated Laboratory and Field Radionuclide Analysis Systems 325
The crew can use such instant data availability to maneuver the aircraft for locating
and monitoring radioactive plumes.
As with any radiation detection system, background is a concern. Ambient
airborne radionuclides (mostly radon daughters) accumulate on the filter and lessen
the detection sensitivity for the radionuclides of interest as compared to a decayed
sample. This accumulation is addressed with a filter carousel, which rotates a fresh
filter into place either at regular intervals or when the background activity level on
the filter becomes excessive.
Another potential background source is contamination of internal and external
surfaces of and near the instrument from exposure to ambient airborne radionu-
clides. These can be deposited from the air stream onto surfaces near the filter
and detector. This background can be expected to increase while the R-TARAC
is exposed to an air stream until the surface contamination reaches mechanical or
decay equilibrium. Tracer tests showed that because of the geometry of surfaces
susceptible to contamination relative to the detector, these deposits contribute
generally less than 1% of the measured radiation. A larger contribution should be
expected after the aircraft has moved through a high-radiation plume into a region
of low-radionuclide concentration.
FIGURE 15.3. Automated gamma-ray spectral analysis system. (By permission of Georgia
Institute of Technology)
fluorescence, and then moved, one at a time, to an elevator that brings the sample
between two photomultiplier tubes for counting. After counting, the sample is
returned to the train.
Similarly, planchets for automated gas-flow proportional-counter systems are
stacked in plastic holders. The bottom holder is moved beneath the detector for
counting and then removed to a second stack for storage or recycling.
For gamma-ray spectral analysis, a set of bulk samples can be placed on a
rotating tray that moves each sample in turn next to the massive shield that encases
the detector (see Fig. 15.3). The door in the shield opens and a mechanical arm
places the sample on top of the detector. After counting, the sample is lifted and
returned to the tray. Alpha-particle spectral analysis generally uses no automation
because the samples are counted for a long time.
with 0.2 mol/l nitric acid, 1 mol/l sodium hydroxide, 0.2 mol/l nitric acid, 0.5 mol/l
oxalic acid, and 2 mol/l nitric acid is required prior to pertechnetate elution.
A flow-through scintillation detector equipped with a lithium glass solid scin-
tillator flow cell is used to detect the eluted 99 Tc. The glass scintillator enables
an absolute detection efficiency of ∼55% and is stable in the 8 mol/l nitric acid
medium for pertechnetate elution.
An automated standard addition technique is part of the analytical protocol. An
aliquot of a 99 Tc standard solution is added to a duplicate of the sample during
acidification. The 99 Tc standard is in a nitric acid solution of the same concentration
used for sample acidification. To perform the standard addition measurement, the
sample acidification monitor instrument automatically substitutes a given volume
of the 99 Tc standard solution for an equal volume of the nitric acid. The volume
of the standard solution is selected by the software to yield an estimated threefold
higher signal relative to the analysis of an unspiked sample.
The total effective analytical efficiency (product of the recovery efficiency and
the detection efficiency) is calculated based on the difference in analytical response
obtained by the analysis of the spiked and unspiked samples. This approach pro-
vides a reliable method for remote, matrix matched, instrument calibration. Au-
tomated standard addition can be used for each sample or batch of samples. The
automated radiochemical analysis procedure is rapid, with a total analysis time of
12.5 min per sample. The total analysis time for the standard addition measurement
is 22 min (including analysis of both unspiked and spiked samples). For low-level
samples, a much longer counting time can be expected.
The analyzer instrument was successfully tested with various waste solution
samples from the US DOE Hanford site, including those with high organic content.
Quantification was verified by independent sample analysis with ICP-MS.
Increasing the size and hence the bulk efficiency of the detector is one possible but
expensive avenue. Increasing the volume of air sample is another avenue. But since
the air contains both the radionuclides of interest and obscuring background ra-
dionuclides, the improvement factor due to volume increase is only approximately
proportional to the square root of the sample volume increase. The approach ap-
plied for substantial improvement in detector–source geometry and thus counting
efficiency was automatic folding or layering of the filter material.
Several sensitivity-enhancing techniques were applied simultaneously in the
Radionuclide Aerosol Sampler Analyzer (RASA) (Miley et al., 1998). This system
employs a Ge detector with about twice the efficiency of contemporary manual
systems (90 vs. 40% relative efficiency) and twice the airflow of manual systems
(24,000 vs. 12,000 m3 per day). A simple layering mechanism provides a large
filter (0.25 m2 ) during sampling and a moderately small filter volume (∼400 cm2
and 0.5-cm thick) for measurement.
Filter volume is minimized for efficient radionuclide measurement with a seg-
mented sampling head and six independent, continuous filter rolls, as shown in
Fig.15.5. These rolls store more than a 1-year supply of daily filter changes for
drawing to and through the sample head. The six simultaneously exposed filters
(each 10 cm × 40 cm) are brought together after exposure and sealed between two
layers of sticky polyester tape that is fed from two rolls. This single filter bundle
then rests in a decay position for 24 h to eliminate gamma rays from the 222 Rn
daughters, 214 Pb and 214 Bi, and reduce those from the 220 Rn daughters (mainly
212
Pb, 212 Bi, and 208 Tl). The filter package is then pulled into a loop around a
rotating drum that circles the germanium detector. The filters remain stationary for
about 24 h for gamma-ray spectrometric analysis and are then advanced by drive
rollers forward 50 cm into the next position. Thus, in normal operation the follow-
ing processes occur: (1) today’s 0.25 m2 filter collects aerosol, (2) the previous
day’s filter is held for decay, and (3) the 2-day-old filter is being measured by the
detector within a small lead cave.
The system is automated with software modules that (1) collect gamma-ray
spectra from the sensors, (2) control relays to activate motors for advancing the
FIGURE 15.5. Segmented sample head, showing filter baffles, rolls of encapsulating
polyester strips, and the wraparound path between the detector and lead shield. U.S. Patent
5,614,724. (By permission of Pacific Northwest National Laboratory)
15. Automated Laboratory and Field Radionuclide Analysis Systems 331
Start up Advance Filter Sample/Count Get Spectrum Save Spectrum Insert Source Count Source
0 1 2 E
4 3 5 6
A C I H J
D Abnormal Event
B G K
7
L M N
8 9 10
0 a state
Critical Error Detector Bad User Shutdown
A
a transition
Stop
FIGURE 15.6. State Machine showing states and transitions. (By permission of Pacific North-
west National Laboratory)
filters through the aerosol collection, storage, and analysis process, and (3) control
the associated processes. In all, a dozen such software devices can completely
control, monitor, and manipulate all the features of the RASA. Because the RASA
performs a sequential set of steps, from startup to shutdown, a State Machine—a
software construct that is the equivalent of a process flow diagram—was chosen
as control, as shown in Fig. 15.6.
Each decision point in the diagram corresponds to a transition from one state
to another. The State Machine most favorably is written such that a noncomputer
programmer can adjust the operation of the device by editing an English language
file that describes the transition from each state to the others. For example, the
RASA always recognizes its current state via electronic signals like pressure, tem-
perature, voltage, and timers. Upon a state change such as a loss of electrical power
or the time of sample measurement, the state is shifted and new functions, such as
filter advance and calibration, or data retrieval, are automatically performed.
An important feature of any automated system is its behavior at the application
and loss of power. The initial and final conditions need to be known to prevent, say,
loss of a sample or sample analysis data. This is easily accomplished by saving
the final condition at loss of power, but battery backup is needed to monitor and
record this state. A catastrophic condition could result if the automated system
fails to shut down or restore properly, such as system damage or bystander harm
from unexpected mechanical actions.
The spectral data collected daily by the system can confirm successful opera-
tional functioning of the RASA, including important features such as start and stop
time, detector resolution, and gain. To assure that the results are correct and can be
332 Harry S. Miley and Craig Edward Aalseth
used as legal evidence, a robust quality assurance (QA) program also is required.
This begins with system certification at the factory or production laboratory and
includes site-specific documentation of local operating procedures. The system
performs daily wide-range energy calibration, which also serves as a check on the
stability of efficiency values.
Regional laboratories that conform to national standards practices perform ex-
ternal QA measures of the network of RASA systems. An international testing
procedure assures that the laboratories remain proficient. Randomly selected fil-
ters from the network of monitoring stations, automatic and manual, are sent to
regional laboratories to determine if the station results are in control. This can be
partially accomplished by measuring the level of 7 Be (t1/2 = 53.28d), a cosmic-
ray spallation product of atmospheric nitrogen and oxygen that is always in the
atmosphere and is easily measurable on air filters.
One of the advantages of this automatic system is that the state-of-health data
recorded for the numerous subsystems, including blowers, component states, tem-
peratures, and other critical information, allow remote failure detection, diagnosis,
and possibly prevention. As an example, variation in detector temperature may
show the onset of failure of a mechanical cooler. Remote diagnostics are used to
schedule repair trips and minimize down time.
FIGURE 15.7. ARSA process diagram. Some duplicate process lines have been omitted for
clarity. (By permission of Pacific Northwest National Laboratory)
condensation with dryers, molecular sieve traps, and charcoal beds, as shown in
Fig. 15.7.
In Area 1 of Fig. 15.7, both trapping and regeneration occur; regeneration steps
are shown by the dashed line. Air is first forced though a pair of pressure-swing
dryers that consist of powdered alumina. While one is drying the incoming sample
air, the second dryer is being regenerated with dry waste air from elsewhere in the
process. These dryers switch from drying to regeneration every few minutes on a
timer, or more frequently, on detection of break-through moisture. The dried air is
then chilled.
The total flow rate is controlled by a commercial mass flow controller (MFC),
which contains an internal servo mechanism that links a mechanical valve to a
resistance thermal device (RTD). The RTD measures mass flow (rather than gas
velocity) by the change of electrical resistance in a sensing wire heated by an
adjacent hot wire. Because this measurement is affected by the specific heat of the
gas, the MFC must be calibrated for each individual gas. The desired MFC flow is
set by applying a voltage to the MFC that corresponds to the voltage generated by
the RTD at that flow. A comparator in the MFC opens or closes the internal valve
to balance the RTD and the applied voltage.
The cold sample gas flows at 100 l/min into a 0.2-l charcoal trap that is cooled
to –90◦ C to adsorb radon. Xenon flows through the trap and is then collected on
the 2-l main charcoal trap at –120◦ C. Nitrogen, oxygen, and the lighter rare gases
helium, neon, argon, and krypton pass through this trap.
334 Harry S. Miley and Craig Edward Aalseth
Gas inlet/outlet
PMT PMT
Source transfer tube
Scintillation cell
Optical window
PMT PMT
Nal Nal
PMT PMT
Optical isolation
FIGURE 15.8. Xenon detector and internal gas cell. (By permission of Pacific Northwest
National Laboratory)
Xenon is next eluted in Area 2. At the end of the sampling period, the main trap
is valved off from the system and heated to 200◦ C to release the xenon. The outflow
from the main trap is carried by ultrapure nitrogen (no CO2 ) through a MFC at
0.24 l/min. Desorption is timed to release most xenon; flow is diverted when the
remaining radon is expected to desorb. This slow flow takes the desorbed xenon gas
with N2 carrier gas though one of a series of disposable chemical traps (NaOH +
Al) to remove most CO2 . The now purified gas is collected on a 0.2-l 5 Å molecular
sieve trap at –40◦ C, while the carrier N2 escapes. The trap is then heated to 200◦ C
to desorb xenon and transferred to a tiny cold trap that is cooled to –120◦ C to sorb
xenon in preparation for loading the counting cell. This step is necessary to transfer
the gas into a scintillation cell, as shown in Area 3 of Fig. 15.7 and also in Fig. 15.8.
The tiny final trap, loaded with the equivalent of 10 cc (at STP) Xe/CO2 product,
is heated to 200◦ C and the product gas is allowed to expand into one of four counting
cells. Transfer efficiency is boosted by use of a syringe pump, as shown in Area 3
of Fig. 15.7. The exact ratio of the Xe and CO2 mix is determined with a thermal
conductivity device (TDC). This is an important measurement because the xenon
separation efficiency depends on various factors, e.g., the conditioning of the traps
and the ambient temperature. The ARSA typically produces a product gas that is
about 50% xenon.
The ARSA system is designed to operate continuously in remote areas with
minimal maintenance and consumables limited to replacement bottles of nitrogen
carrier gas and CO2 removal traps. Dryers and gas traps were chosen that could
be automatically regenerated in the field. The system has many components and a
complex path for the analyte gas that traverses the system. The numerous valves
(∼100) for controlling the gas flow and trap regeneration processes require a control
system about one order of magnitude more complex than that for the RASA system
described in the preceding section, if the number of parts is taken as a measure of
15. Automated Laboratory and Field Radionuclide Analysis Systems 335
10000
1000 30 keV
Gamma Singles
81 keV
133Xe (32 mBq/m3)
Counts
100
250 keV
135Xe (3.7 mBq/m3)
10
Beta Gated
1
0 50 100 150 200 250 300 350 400 450 500
Channel Number
FIGURE 15.10. Conceptual diagram of FDTAS, showing process components from sampling
to analysis. (By permission of Pacific Northwest National Laboratory)
channels, and pulse shape can be recorded to correct for the presence of other
radionuclides and the effects of quenching or luminescence (see Section 8.3.2). The
counting period is selected to achieve the required detection limit. Measurements
may be at the rate of 1 per 100 min or less frequent if longer counting times are
required.
The FDTAS design adds water purification and improved detection sensitivity.
Laboratory tests of the various methods of purifying environmental samples appli-
cable to unattended field deployment led to selection of a single use, commercially
available, mixed bed resin column for water purification. By employing active
background suppression and pulse shape discrimination with bismuth germanate
(BGO) guard scintillators, lead shielding, and low-background cell components,
the FDTAS achieves tritium detection sensitivity that almost rivals laboratory LS
instruments. The combination of a shield, an active BGO guard detector, and low-
background materials attains a routine background of ∼1.5 count/min in the tritium
energy channel. A special low background quartz counting cell, containing 11 ml
of a 50:50 mixture of the sample and the liquid scintillation counting fluid, yields
a tritium detection efficiency of ∼25%.
The combined low background, high detection efficiency, and 5.5-ml sample
volume lead to a detection limit of 10 Bq (600 dpm) per liter for a 100-min count
(95% confidence limit). These results are achieved routinely in field tests of the
FDTAS at monitoring wells, surface streams, ground water remediation facilities,
sewage treatment plants, and the Savannah River. Results have been confirmed by
parallel sampling and laboratory analyses.
The automation of the FDTAS is of a classic and cost effective type that mostly
uses commercial hardware and support from manufacturers to create sophisticated
custom radiation detection apparatus. QC capabilities are built-in to validate the
results. Nearly real-time reporting allows operational authorities a cushion of time
to end a major tritium release and mitigate its effects.
15.5. Summary
A variety of radiochemical measurements have been successfully automated, both
in the laboratory and in the field. It might appear to the student that the effort
involved in automating a process is more trouble than the manual measurement of
the activity. To an experienced practitioner, it is exactly this economic and qual-
ity decision point that makes the selection of automation or manual approaches
interesting. When the expected variations in sample type and process parameters
allow, automation should be evaluated for ongoing capabilities and field measure-
ments. But it should also be said that the interactions of scientists, technicians, and
engineers on an automation project can be truly rewarding in many ways, as the
fusion of disciplines tends to accelerate the productivity of all.
16
Chemistry Beyond the Actinides
DARLEANE C. HOFFMAN
16.1. Introduction
Humans have been fascinated ever since ancient times with trying to understand
the composition of the terrestrial world around them and even the stars beyond
their reach. In the 4th century B.C., Aristotle proposed that all matter could be
described in terms of varying proportions of four “elements”—air, earth, fire, and
water. Elements, including gold, silver, and tin, that are found relatively pure in
nature were isolated and used over the period of the next several hundred years.
The alchemists of medieval times isolated and discovered additional elements and
used secret formulas and rituals in the attempt to find the “philosopher’s stone” and
attain their goal of transmuting lead into gold. The development of experimental
science and the scientific method in the 18th century accelerated the pace of the
discovery of new elements, but uranium, discovered in 1789 by Klaproth in pitch-
blende from Saxony, Germany, remained the heaviest known chemical element for
150 years.
After the discovery by Becquerel in 1896 that uranium salts blackened photo-
graphic plates due to radiations from the natural decay chains of uranium, Marie
Curie and Pierre Curie (Curie and Curie, 1898) began extensive studies of the new
phenomenon that they dubbed “radioactivity.” They were successful in isolating
and identifying the first radioactive elements Po (84) and Ra (88) in pitchblende.
They shared the Nobel Prize in Physics with Becquerel in 1903 for their studies
of radioactivity. Marie Curie continued her investigations of radioactive species
and isolated macro quantities of Ra to be used for medical purposes. She received
the Nobel Prize in Chemistry in 1911 for the isolation of pure radium and the
determination of its molecular weight. She is recognized as the founder of the sub-
discipline of “radiochemistry,” which can be defined as the study of radioactivity
and radioactive elements; she could rightfully be acknowledged as the founder of
nuclear medicine as well.
In the mid-1930s, the new breed of “nuclear” scientists, including both chemists
and physicists, became intrigued with the possibility of actually synthesizing new
“artificial” elements not found in nature. The discovery of “artificial” radioactivity
by Joliot–Curies in 1934, the invention of the cyclotron by E. O. Lawrence in
338
16. Chemistry Beyond the Actinides 339
FIGURE 16.1. 2004 periodic table showing placement of transactinides through element
154.
1999), this transition series ends at element 112 with the filling of the 6d shell.
The p shells are filled in elements 113 through 118, which are expected to be the
heaviest members of the noble gas group. The 8s shell is expected to be filled
in elements 119 and 120, making them homologs of groups 1 and 2. Based on
relativistic calculations, element 121 should have an 8p electron in its ground-
state configuration in contrast to the 7d electron expected by simple extrapola-
tion from the elements in group 3 of the periodic table. A 7d electron would
then be added in element 122, giving it the configuration [118]8s2 7d8p, differ-
ent from that expected from its homolog thorium, which has the configuration
[Rn]7s2 6d2 . The situation becomes even more complicated beyond element 122
because the energy spacings of the 7d, 6f, and 5g levels as well as those of
the 8p, 9s, and 9p levels are so close that the chemical properties of these ele-
ments will be nearly impossible to characterize on the basis of currently known
properties.
Research (JINR) in Dubna, the USSR. In 1974, the International Union of Pure
and Applied Chemistry (IUPAC) and the International Union of Pure and Applied
Physics (IUPAP) appointed an ad hoc committee of neutral experts to consider
the competing claims and to facilitate cooperation between the groups in reaching
agreement on these issues. However, the committee never met as a whole, and it
was finally disbanded in 1984, although a few smaller meetings with some of the
Berkeley and Dubna researchers were sponsored, observers were exchanged, and
some reports were issued (Hyde et al., 1987).
Pending settlement of the competing claims to discovery and approval of names,
the Commission on Nomenclature of Inorganic Chemistry (CNIC) of the IUPAC
mandated (Chatt, 1979) the use of three-letter “systematic names” based on 0=nil,
1=un, 2=bi, 3=tri, 4=quad, 5=pent, 6=hex, 7=sept, 8=oct, and 9=enn for el-
ements with Z > 100. Thus, elements 104 and 105 officially became unnilqua-
dium (unq) and unnilpentium (unp), although these names never found common
usage among heavy element researchers. In the meantime, the names kurcha-
tovium (Ku, 104) and nielsbohrium (Ns, 105) and rutherfordium (Rf, 104) and
hahnium (Ha, 105) continued to be used by the Dubna and Berkeley groups, re-
spectively.
Among the reasons for the controversies over discovery of elements 104 and
105 are their short half-lives and small production rates that made it necessary to
study and identify them “online” at the accelerators where they were produced on
the basis of their decay properties rather than on “conventional” radioanalytical
methods. New radioanalytical techniques had to be developed for unequivocally
establishing that a new element had, indeed, been produced.
The Berkeley group developed the α–α correlation technique in which the un-
known element’s α-decay is correlated with that of the α-decay of known daughter
and/or granddaughters. In some cases, the daughters can be identified by radio-
chemical separations. The measurement of the characteristic X-rays of a new ele-
ment is another definitive method. It was used at Oak Ridge National Laboratory
by Bemis et al. (1973, 1977) to confirm the Berkeley group’s discoveries of ele-
ments 104 and 105. However, this method requires detection of relatively “large”
numbers of events in order to obtain statistically significant measurements of the
X-ray energies. In contrast to these methods, the Dubna group relied primarily
on the detection of spontaneous fission (SF) decay. This is an extremely sensitive
technique, but detection of only the fragments from the fission process makes it
extremely difficult to deduce the identity of the fissioning species, especially based
on detection of only a few events.
Indirect methods such as half-life systematics, excitation functions for the pro-
duction reactions, and cross bombardments have been used to reinforce this in-
formation. In order to positively identify the atomic number of a spontaneously
fissioning nuclide from detection of the fragments, the atomic numbers of both
primary fragments from the same SF event must be determined in coincidence
and added together to determine the Z of the new, unknown fissioning nuclide.
Detection of only SF decay has resulted in much controversy concerning discovery
and identification of the transactinide elements.
342 Darleane C. Hoffman
Transactinides
104 Rutherfordium Rf 1969
105 Dubnium (Hahnium)a Db (Ha)a 1970
106 Seaborgium Sg 1974
107 Bohrium Bh 1981
108 Hassium Hs 1984
109 Meitnerium Mt 1982
a Many publications of chemical studies prior to 1997 use hahnium (Ha) for element 105.
suggested by the discoverers for the elements they claimed to have discovered and
which had been in common use.
A worldwide storm of protest and criticism resulted and the IUPAC con-
vened a series of meetings to try to agree upon a compromise set of names.
The claims to discovery and the subsequent naming controversies surrounding
elements 104–109 are discussed in detail in Chapters 9, 10, and 13 of Hoff-
man et al. (2000). Finally, after several unacceptable naming proposals, the
IUPAC backed down on their edict disqualifying living persons, and the so-
called compromise list of names and symbols shown in Table 16.1, including
“seaborgium, Sg” for element 106, was approved in August 1997 (CNIC, 1997).
This was only a year before Glenn Seaborg suffered the stroke and fall at the
1998 Boston ACS meeting, which ultimately resulted in his death on February 25,
1999.
a Silva
(1970); b Gregorich (1988); c Schädel (1997a); d Eichler (2000); e Düllmann (2002a,b); Kirbach
(2002).
faster than manual operations, they usually give more reproducible results and are
nearly essential for around-the-clock experiments lasting weeks at a time.
To avoid dissolution of the highly radioactive and rare targets listed in Table 16.2
that must often be used to produce the isotopes of interest, foils sometimes are
placed directly behind the target to catch the activities recoiling from the target
after irradiation and then they are processed. Now it is more common to attach
the recoiling activities to aerosols in a flowing stream of inert carrier gas such as
He in which they can be rapidly transported outside the irradiation site within the
accelerator shielding to the site of the chemical system to be used.
Cm
Mg beam
targets
He/O2 Quartz
column Thermostat
LN
Gas Oven
chamber
IVO COLD System Detector Pairs
FIGURE 16.4. Schematic of IVO-COLD system for study of gas-phase properties of HsO4
and lighter homologs. Adapted from Düllmann et al. (2002b).
There also is the usual problem of positively determining the atomic number of
the fissioning nuclide whose chemistry was being studied.
Recently, these problems were solved in a very imaginative manner by forming
the chromatographic column from opposing (Si) photodiode detectors upon which
the volatile radioactive species were directly deposited. In this way, both the radi-
ations (α and SF), and their deposition positions as a function of the temperature
gradient along the Si column, were determined simultaneously and continuously
recorded and stored via a computer system. A gradient cryogenic version of this
system, the Cryo On-Line Detector (COLD), was used to perform the first success-
ful chemical studies of Hs (108) with the α-emitting isotopes 269,270 Hs, which were
identified by their decay to known α-emitting daughters. The COLD system was
used together with the In situ Volatilization and On-line (IVO) detection system
to study the volatile oxides of Hs and Os. A schematic diagram of the IVO-COLD
system (Düllmann et al., 2002a,b) is shown in Fig. 16.4. The study indicated that
Hs formed a volatile oxide similar to that of the tetroxide of Os, its lighter group 8
homolog.
Online isothermal chromatographic systems, such as the On-Line Gas Analyzer
(OLGA) developed by Gäggeler and coworkers (Gäggeler, 1994) and the Heavy
Element Volatility Instrument (HEVI) developed by Kadkhodayan et al. (1996),
have been used to study volatile halides and oxyhalides of Rf, Db (Ha), Sg, and
Bh. The α-emitting isotopes shown in Table 16.2 were used in these studies. In
these systems, the entire length of the chromatographic column (usually quartz)
is kept at a constant temperature and the volatile species pass through the column
in a carrier gas such as He and undergo numerous sorption/desorption steps. The
yield through the column is determined by measuring the α-radiations from the
exiting species that are either reattached to aerosols and transported to a detection
device or condensed directly on a detector system.
Experiments are conducted at a series of isothermal temperatures to study the
chemical yields of the volatile species as a function of temperature. Retention
time is indicative of the volatility at the given isothermal temperature. It is equal
348 Darleane C. Hoffman
to the half-life of the isotope at the temperature at which 50% of the plateau
or steady-state yield is obtained. This T50% retention time can then be used as
a relative measure of volatility. A Monte Carlo program taking account of all
the experimental parameters, such as gas flow rate, length of column, and the
half-lives of the isotopes in the volatile species, is used to deduce the adsorption
enthalpies for the volatile species. A schematic diagram of the HEVI system, used
for isothermal gas-phase studies, is shown in Fig. 16.5. The recoiling products from
the reaction are attached to either KCl, KBr, or MoO3 aerosols, and transported
in He gas to the entrance to HEVI where they are deposited on a quartz wool
plug and halogenated at 900◦ C. The volatile products are transported through the
quartz column in flowing He gas and are again attached to aerosols in the recluster
chamber and transported via another gas-jet to the Merry Go-around (MG) system.
They are then deposited on thin polypropylene films placed around the periphery
of its horizontal wheel, which is rotated to position the foils successively between
pairs of surface barrier detectors for α and SF spectroscopy.
Photos of HEVI, OLGA, and the MG rotating wheel and detection system are
shown in Fig. 16.6(a)–(c); parts (d) and (e) of Fig. 16.6 will be discussed in the
following section.
FIGURE 16.6. (a) B. Kadkhodayan with HEVI (1992); (b) J. Kovacs and H. W. Gäggeler with
OLGA (1988); (c) D. C. Hoffman and D. M. Lee with MG rotating wheel system; (d) J. V.
Kratz and M. Schädel with Automated Rapid Chemistry Apparatus (1988); (e) Schematic
diagram of a typical SISAK liquid–liquid extraction configuration with Berkeley Gas-filled
Separator as a preseparator.
350 Darleane C. Hoffman
procedures, longer lived daughter products can be detected and used to deduce
the properties of the parent element. The ARCA shown in 16.6(d) has been used
successfully in separations of Rf through Sg. Several reviews of the results have
been published (Hoffman, 1994; Schädel, 1995; Hoffman and Lee, 1999; Kratz,
1999a; Kratz, 1999b).
The SISAK system, which had previously been used to perform liquid–liquid
extraction studies of γ-emitting nuclides with half-lives as short as a second,
was adapted for use with α-emitters by coupling it to a flowing liquid scintillation
system. This provides continuous separation and measurement of α–α correlations
and detection of SF decay of nuclides with half-lives of only a few seconds. This
permits chemical studies of short-lived nuclides but the energy resolution is not as
good as with solid-state detectors. Nevertheless, a recent experiment by Omtvedt
et al. (2002) showed that rapid preseparation in the Berkeley Gas-filled Separator
(BGS) achieved sufficient decontamination from the extremely high background
of unwanted activities. Detailed information was obtained about the chemical
properties of Rf for 4.7-s 257 Rf produced in the 208 Pb(50 Ti,n) reaction. Similarly,
the chemistry of Db (Ha) can be studied with 4.4-s 258 Db and still heavier elements
can be investigated although the production rates are steadily decreasing. Another
limitation is the requirement to choose extraction systems that have kinetics that are
rapid enough that equilibrium can be attained in the short extraction contact times
associated with the SISAK system. A schematic diagram of the SISAK system as
configured for these experiments is shown in Fig. 16.6(e).
and in the nuclear fuel cycle; (7) surveillance of clandestine nuclear activities and
evaluation of terrorist threats using gas-analysis techniques; (8) radioisotope pro-
duction and separation; (9) radiopharmaceutical synthesis, and detection systems
for both diagnostic and therapeutic nuclear medicine procedures; (10) biochemical
and agricultural research; and (11) geochemical dating studies.
-70
ZrCl4
RfCl4
-80
ΔH ads(kJ/mol)
RfBr4
Volatility
-90
HfCl4
o
-100
ZrBr 4
-110
HfBr4
-120
-130
30 40 50 60 70 80 90 100 110
Atomic number
FIGURE 16.7. Adsorption enthalpy values and relative volatilities on SiO2 for Zr, Hf, and
Rf chlorides and bromides.
actinide and Sc chlorides. They concluded that element 104 should be placed in
group 4 of the periodic table. The results were not considered conclusive (Kratz,
1999b) because only SF events were measured, making it impossible to identify
positively the element being studied.
After these initial studies, a rather long time elapsed before additional exper-
iments were conducted in the late 1980s. These were motivated by theoretical
calculations, suggesting that relativistic effects might alter the chemical properties
of Rf and Ha from those expected by simple extrapolation of properties within a
given group in the periodic table. Subsequent studies of the gas-phase chemistry
of Rf, Zr, and Hf chlorides (Kadkhodayan et al., 1996; Türler et al., 1996) and
bromides (Sylwester et al., 2000) were conducted with OLGA and HEVI at the 88-
Inch Cyclotron at LBNL. Contrary to expectations based on simple extrapolation
of group 4 volatilities, these studies showed (Türler et al., 1996) that the Rf halides
were more volatile than those of Hf. In agreement with relativistic calculations
(Pershina and Fricke, 1999), the bromides were all found to be less volatile than
their respective halides as shown in Fig. 16.7
Many additional studies to compare the solution chemistry of Rf with Zr, Hf,
Th(IV), and Pu(IV) have been conducted and include both manual column and ex-
traction studies, and automated studies using ARCA. Extractions from aqueous so-
lutions into triisooctylamine, tributylphosphate, and thenoyltrifluoroacetone have
been studied in considerable detail. They have shown that, although Rf behaves
generally like a group 4 element, its behavior varies with different complexing
agents.
16. Chemistry Beyond the Actinides 353
the same for Nb and Ta, but Db appeared to be significantly less volatile and it was
postulated that an oxybromide might have formed from traces of oxygen present.
It is predicted to be less volatile than the pentabromide (Pershina et al., 1992).
Later experiments in which the partial pressure of oxygen was varied showed that
both volatile and less volatile species were formed and additional measurements
are needed to try to produce the pure pentabromide.
mode was used to avoid interference from the ubiquitous 212 Pom 8.8–MeV α-
activity. Three events attributable to 265 Sg (7 s) were identified: two in daughter
mode and one triple correlation. One α–α correlation between 266 Sg (21 s) and
its 262 Rf daughter was detected. This gas-phase study of SgO2 Cl2 constituted the
first chemical study in which the detected Sg activities were positively identified
as belonging to Sg.
In a later experiment (Türler et al., 1999), the isothermal temperature was varied
to obtain the adsorption enthalpy for Sg. The yield of short-lived W isotopes from
reactions on Gd incorporated in the target was also measured at the same time. An
adsorption enthalphy of –96 ± 1 kJ mol−1 was obtained for the dioxydichloride of
W in agreement with the previous measurement. Based on 11 events attributable
to Sg, an adsorption enthalpy of –100 ± 4 kJ mol−1 or 96 ± 1 kJ mol−1 was
obtained depending on the Sg half-life used in the Monte Carlo calculations. Thus,
the experimental results confirmed the theoretically predicted volatility sequence
of Mo > W > Sg. From an analysis of the decay properties of the correlated decay
chains detected in these experiments, Türler et al. (1998) were able to recommend
better half-lives and cross sections of 7.4 + 3.3/ − 2.7 s and approximately 0.24
nb for 265 Sg and of 21 + 20/ − 12 s and approximately 0.025 nb for 266 Sg for 22 Ne
beam energies between 120 and 124 MeV.
Schädel et al. (1997a,b) also reported the first successful studies of the aqueous
chemistry of Sg in the same series of collaborative experiments conducted at the
UNILAC at GSI. In order to favor production of the longer lived 266 Sg formed by
the 4n out reaction, 121-MeV 22 Ne projectiles were used. A helium gas-transfer
system was used to transport the activities from the irradiation site to the ARCA
II system, where 3,900 collection and elution cycles were performed. A solution
of 0.1 M HNO3 /5 × 10−4 M HF was used to elute the activity sorbed initially
on columns (1.8 mm i.d. × 8 mm long) filled with the cation exchange resin
Aminex A6. This eluant was chosen because previous online tracer experiments
had shown that the formation of neutral and anionic complexes with F− ions is
characteristic of the group 4, 5, and 6 elements, but that there are distinct differ-
ences in behavior between the groups. The procedure was designed to provide
rapid decontamination from the interfering high-energy Bi and Po α-activities and
trivalent actinides also formed in the irradiation, and efficient separation of the Rf
and No daughter activities from their Sg precursors. The mean time for separation
of Rf and No from Sg was only approximately 5 s, but α-particle and SF measure-
ments were not begun until approximately 38 s after the end of collection because
of the time required to evaporate the eluted samples and manually transport them
to the PIPS detectors. The energies, times, and detector positions were recorded
in list mode on magnetic disk and tape for later data analysis. Collection times of
45 s were used for most of the cycles.
Three correlated α–α-events were identified as belonging to the 261 Rf
(78 s) →257 No (26 s) → decay sequence, thus indicating that the 7-s 265 Sg parent
had been present in the chemically separated Sg (W) fractions and decayed to Rf
and No during the time interval before measurements began. No evidence for the
longer lived 266 Sg was found, presumably because of its much smaller production
356 Darleane C. Hoffman
cross section. Later investigations were performed (Strub et al., 2000) to verify that
Rf, Th, Hf, and Zr would not elute from the cation columns under these conditions,
and it was concluded that 261 Rf could have been only in the Sg fraction as a result
of the decay of 265 Sg and that Sg, like its lighter group 6 homologs Mo and W,
formed neutral species of the type MO2 F2 .
Another similar series of more than 4,500 cycles was conducted (Schädel et
al., 1998) under the same conditions using 0.1 M HNO3 but without any flu-
oride ions to determine whether Sg would then behave in a manner different
from that of W as suggested by theoretical examination of the hydrolysis of
group-6 elements and U(VI) (Pershina and Kratz, 2001). Indeed, the Sg activ-
ity remained on the cation-exchange resin columns while W was eluted with
0.1 M HNO3 . This behavior was tentatively attributed to the weaker tendency
of Sg to hydrolyze in dilute HNO3 so that it remained as a 2+ or 1+ complex
while hydrolysis of W (and Mo) resulted in the neutral species MO2 (OH)2 . In
the earlier experiments containing fluoride ion, Sg may have been eluted from
the cation-exchange columns as neutral or anionic fluoride complexes of the type
SgO2 F2 or [SgO2 F3 ]− rather than as [SgO4 ]2− . Theoretical calculations of com-
plex formation in HF solutions (Pershina and Hoffman, 2003) indicate that com-
plex formation competes with hydrolysis in aqueous solutions and the dependence
on pH and HF concentration is extremely complicated and reversals in trends
may occur. The theoretical predictions provide experimenters with valuable guid-
ance for planning future experiments. These studies illustrate the fruitful syner-
gism that can result from close interactions and iterations between theory and
experiment.
60
50
40
Relative Yield [%]
30
20
10
0
–20 –40 –60 –80 –100 –120 –140 –160
Deposition Temperature [∞C]
FIGURE 16.8. Merged thermochromatogram of HsO4 and OsO4 . Adapted from Düllmann
et al. (2002b).
A negative temperature gradient from –20◦ C to –170◦ C was established along the
TC and measured at five positions.
Three α–α correlated decay chains detected in the 64-h experiment were as-
signed to the decay of 269 Hs. Two others possibly due to a new nuclide 270 Hs or an
isomer of 269 Hs were also detected. The merged thermochromatograms for OsO4
and HsO4 are shown in Fig. 16.8. Indicated are the relative yields of HsO4 and
OsO4 as a function of the deposition temperature. Measured values are represented
by bars—HsO4 : light grey; OsO4 : dark grey. The distributions of the seven decay
chains attributed to Hs isotopes are indicated by the pattern—269 Hs: blank; 270 Hs:
hatched; Hs (isotope unknown): cross-hatched. For Os, the distribution of 1·105
events of 172 OsO4 is given. The maxima of the deposition distributions were evalu-
ated as (−44 ± 6)◦ C for HsO4 and (−82 ± 7)◦ C for OsO4 where the uncertainties
indicate the temperature range covered by the detector that registered the maxi-
mum of the deposition distribution. Solid lines represent results of a Monte Carlo
simulation of the migration process of the species along the column with standard
adsorption enthalpies of −46.0 kJ mol−1 for 269 HsO4 and −39.0 kJ mol−1 for
172
OsO4 .
As noted above, the best Monte Carlo fits to the Os data gave an adsorption
enthalpy of −39 ± 1 kJ mol−1 for OsO4 on the Si3 N4 surface in good agreement
with previous measurements on SiO2 . Using only the three events assigned to
269
Hs and a half-life value of 11 s in the Monte Carlo analysis resulted in an
adsorption enthalpy of −46 ± 2 kJ mol−1 for HsO4 . This is considerably lower
than the calculated value of −36.7 ± 1.5 kJ mol−1 and suggests that it may be
less volatile than OsO4 . Additional experiments to measure more 269,270 Hs decay
16. Chemistry Beyond the Actinides 359
chains and to clarify the nuclear properties are clearly needed. However, the study
does show that Hs forms a volatile oxide similar to that of the tetroxide of Os, and
should be placed in group 8 of the periodic table.
No studies of Bh or Hs in solution have yet been attempted, but they should be
quite interesting and could furnish valuable information about most stable oxida-
tion states in aqueous solution and redox and complexing reactions in different
media. They might be expected to show oxidation states ranging from 3 to 7 for
Bh and 1 to 8 for Hs, as do their homologs in groups 7 and 8. A summary of the
chemical properties predicted for Bh and Hs was given in a review by Seaborg and
Keller (1986).
16.4. Future
16.4.1. Prospects for Chemical Studies of Mt (109)
Through Element 112
Very little attention has been paid to theoretical predictions of the chemical prop-
erties of Mt through element 120 since the 1986 summary of Seaborg and Keller.
Based on the positions of Mt and element 112 in the periodic table, they should be
noble metals like Pt and Au, and volatile hexafluorides and octafluorides might be
produced and used in chemical separation procedures. Early relativistic molecular
calculations (Waber and Averill, 1974; Rosen, 1998) suggested that 110F6 should
be similar to PtF6 .
Because the maximum of relativistic effects in the ns shell in group 11 is at
element 111, there has been considerable interest in the electronic structures of its
compounds. Theoretical studies (Seth et al., 1996; Liu and van Wüllen, 1999) of
the simplest molecule 111H show that the bonding is considerably increased due
to relativistic contraction of the 7s orbital. Theoretical studies (Seth et al., 1998a,b)
of the stability of higher oxidation states support earlier predictions that the 3+
and 5+ oxidation states will be more common for element 111 than they are for
Au and that the 1+ oxidation state may be difficult to prepare.
Element 112 is the most interesting of the heaviest elements from the chemical
point of view because the maxima of both the relativistic effects on the 7s shell in
the entire 7th row and within group 12 occur at element 112. The strong relativistic
contraction and stabilization of the 7s orbitals and its closed shell configuration
should make it rather inert, and the predicted rather small interatomic interaction
in the metallic state may even lead to high volatility as seen in the noble gases
(Pitzer, 1975). A simple extrapolation of the sublimation enthalpies of its lighter
group 12 homologs, Zn, Cd, and Hg, leads to a similar conclusion.
Recent relativistic molecular orbital calculations (Pershina et al., 2002) con-
firmed that element 112 can form rather strong metal–metal bonds 112M and
deposit on some transition metals as 112M in which M = Cu, Pd, Ag, Pt or Au.
The stability of the 2+ state is predicted to decrease from Hg to element 112,
and 112F2 may tend to decompose but 112F4 may be stable. The species 112F− 3
360 Darleane C. Hoffman
and/or 112F− 5 may form in solution in an appropriate polar solvent, but they will
probably quickly hydrolyze and bromides or iodides may be more stable and bet-
ter for experimental studies of solution chemistry. A recent review of the results
of theoretical studies of the properties of Mt through element 112 has been pub-
lished (Pershina and Hoffman, 2003) and should be consulted for more detailed
and up-to-date information and complete references.
Some preliminary chemical experiments on element 112 have been reported
(Yakushev et al., 2001) using the spontaneously fissioning nuclide 283 112 (∼3
min) reportedly formed in the 238 U(48 Ca, 3n) reaction with a cross section of about
5 pb. The experiments were designed to test whether element 112 behaved more
like Hg, which had been shown to deposit on Au- or Pd-coated silicon surface-
barrier detectors or was a noble gas like Rn and stayed in the gas phase. Eight
SF events were detected in the gas phase, which would indicate Rn-like behav-
ior. However, the results cannot be considered definitive since it was not shown
that the SFs belonged to element 112 nor has the original report of the approxi-
mately 3-min SF activity attributed to 283 112 been confirmed. Additional exper-
iments are planned, but use of an α-decaying isotope of 112, such as 284 112 (10
to 70 s) reported previously (Oganessian, 2001) to be the daughter of 288 114,
may be required to ascertain whether or not element 112 was actually being
observed.
Chemical studies of Mt depend on the discovery of longer lived isotopes than
42-ms 268 Mt, currently the longest known isotope of Mt, shown in Fig. 16.2. Longer
lived isotopes might be expected around the deformed nuclear shell at 162 neutrons.
The half-life of the 162-neutron nuclide 271 Mt can be estimated from interpolation
of Smolańczuk’s calculations (Smolańczuk, 1997) for even-proton, even-neutron
nuclides to be a few seconds and should decay primarily by α-emission. However,
270
Mt with only 161 neutrons may be only tens of microseconds, too short for
chemistry. Possible production reactions for 271,270 Mt include 238 U(37 Cl, 4n,5n)
and 249 Bk(26 Mg, 4n,5n). The cross sections for these reactions may be a few pb
and only tenths of pb, respectively. The BGS at LBNL is capable of separating
and positively identifying such a new nuclide based on measurement of the known
α-decay chain of its daughters even if the 271 Mt half-life is as short as tenths of
seconds. Rotating, multiple targets of 238 U could be used to increase the production
yields for the first reaction but use of multiple targets of the highly radioactive 249 Bk
would be extremely difficult.
The half-lives of 7.5 s and 1 min reported for 280,281 110 produced as granddaugh-
ters of the 288,289 114 decay chains are long enough for chemical studies. However,
the reports need to be confirmed and the yields via this decay route must be high
enough to permit statistically significant atom-at-a-time studies. Currently, there
appear to be no suitable reactions for direct production of these very neutron-rich
isotopes of element 110.
Some isotopes of element 111 should have half-lives of seconds or more. Promis-
ing production reactions need to be investigated first using an online separator and
detection system to measure half-lives and production cross sections before chem-
ical studies can be contemplated.
16. Chemistry Beyond the Actinides 361
17.1. Introduction
Like most analytical instrumentation, the mass spectrometer (MS) initially found
use as a research system. These complex laboratory-built research instruments
were subsequently refined for commercial sale and widespread use, and are now
available for routine sample assay.
In recent years, the MS has become a useful alternative to the radiation detector
for measuring longer-lived radionuclides. It provides a second and completely
independent method of measurement to confirm results; it also gives the analyst
a choice, because some measurements are easier or more reliable by one method
than the other. The MS has been used to measure radioactive atoms with half-
lives greater than 10 years because the number of these atoms relative to their
decay rate is proportional to the half life; for half lives greater than 109 y, even the
conventional measurements of analytical chemistry are applicable.
Development of new MS systems focuses on the ability to measure smaller
samples accurately, either for greater sensitivity or shorter-lived radionuclides.
Advances in both construction and design have been applied, with the effect that
the modern MS operator can measure ever-smaller numbers of atoms. A related de-
velopment is the ability to individuate those samples more thoroughly by increasing
MS resolution to decrease interference from isobars—atoms and molecules with
the same atomic mass number but with minute differences in mass.
The distinction in sensitivity between measurements of radiation and atoms (as
ions at a given mass-to-charge ratio by MS) can be demonstrated by comparing
intensity in terms of the radioactive decay Eqs. (2.4) and (2.7) for a radionuclide
with an atomic mass of 100 amu and a half-life of 10 y (3.16 × 108 s). If the
sample under assay emits beta particles at the relatively low rate of 1 × 10−2 d/s,
this corresponds to 4.6 × 106 atoms or, for a 100 amu radionuclide, 7.6 × 10−14 g.
Alpha particles can be measured at an approximately 1,000-fold lower decay rate,
for which the number of atoms and the mass correspondingly are 1,000-fold lower.
1
Pacific Northwest National Laboratory, Richland, WA 99352
2
To whom correspondence should be addressed. John F. Wacker, MS P8-01, 902 Battelle
Blvd., Richland, WA 99354; email: john.wacker@pnl.gov
362
17. Mass Spectrometric Radionuclide Analyses 363
The radiation of this test isotope can be measured at these levels but the number of
atoms (4.6 × 103 ) falls below the detection limits for quantification of most mass
spectrometers.
The efforts in applying mass spectrometry to the measurement of radionuclides
are devoted to achieving detection of small numbers of atoms. The crucial question
is whether an isotope that has, say, 106 atoms in a gram of solid or liquid can be
detected and quantified by MS in the presence of possibly 1021 atoms and molecules
of other isotopes, including many with nearly the same mass. For some elements
(e.g., Pu), detection limits below 106 atoms have been demonstrated for a variety of
sample matrices (soil, water) using several mass spectrometric techniques (Beasley
et al., 1998b; Oughton et al., 2004; Sahoo et al., 2002). Instrumental detection
limits, in terms of counts (i.e., detected ions) per atom, have exceeded 5 counts
per hundred atoms under optimum conditions.
This chapter provides an overview of mass spectrometer function and operation.
It describes specific instrument types with demonstrated or potential application
for measuring radionuclides and surveys the application of these instruments to
radionuclide detection. Finally, it discusses the circumstances under which use
of mass spectrometers is advantageous, the type of mass spectrometer used for
each purpose, and the conditions of sample preparation, introduction and anal-
ysis. Its perspective is from a national laboratory active in environmental and
non-proliferation monitoring. It emphasizes isotope ratio measurements, but mass
spectrometric measurements also provide isotope mass information. Several re-
cent books describe elemental and isotope ratio mass spectrometry in far greater
detail than is presented here (Barshick et al., 2000; De Laeter, 2001; Montaser,
1998; Nelms, 2005; Platzner, 1997; Tuniz et al., 1998). High-resolution mass spec-
trometry forms the basis of the mass scale used for elemental and isotopic masses
(Coplen, 2001), but this application of MS falls outside the scope of this chapter.
Each of the instrumental components has electronic control for operation and
data acquisition.
The following is a brief description of the various ionization sources used in mass
spectrometry. Those sources most commonly used in the analysis of radionuclides
will be described in greater detail later in the chapter.
Electron impact (EI) ionization is useful for elements that are either volatile or
form volatile compounds. Typical ionization efficiencies are in the range of 0.1 to
1 percent. Suitable elements include noble gases and light elements such as C, N,
O. Other elements include those that form volatile compounds (e.g., uranium in
the form of UF6 ). Except for specialized applications (e.g., noble gas analysis), EI
is no longer widely used for elemental and isotopic analysis.
Thermal ionization is useful for elements with ionization potentials less than
about 7 electron volts (eV). The thermal ionization process forms positive ions
when an analyte is evaporated from a hot metal filament. The filament can be a
single filament, where ionization and evaporation occur together, or a double or
triple filament, where the sample is evaporated from one filament and the vapor
is ionized on a separate filament. Common filament materials include rhenium,
tungsten, and tantalum. Whatever the arrangement of the filament, it is always
located above the case plate, shown at the top of Fig. 17.1.
Ions are generated at the filament and accelerated downward towards the exit
plate. In a typical mass spectrometer, the filament is at high voltage (typical voltages
are +5,000 V for positive ions and –5,000 V for negative ions; voltages from
1,000 to 15,000 can be used) and the exit slit plate is at ground. Typical ionization
efficiencies range from 0.1 to 10 percent. Suitable elements include alkali, alkaline
earth, many transition elements, the rare earths, and actinides, as well as lead and
boron. Negative thermal ionization has become important in the past 15 years and
it has increased the range of elements available for analysis by thermal ionization.
17. Mass Spectrometric Radionuclide Analyses 367
FIGURE 17.1. Thermal ionization source. Figure from Smith (2000), pg. 9.
As with the positive ion case, typical ionization efficiencies are 0.1 to 10 percent.
Suitable elements include the halogens and oxides of transition elements such as
Mo, Re, and Os.
Inductively–coupled plasma (ICP): An argon ICP will ionize all but a few ele-
ments (e.g., fluorine) and, therefore, is a nearly universal ion source. Figure 17.2
shows the ICP source; its ionizing coil and torch are on the right, while the vacuum
interface is shown on the left.
Samples are introduced into the plasma as an aerosol formed by nebulization of
a solution, by vaporization of solids in a furnace, or using laser ablation. Analyte
vapor or aerosol is carried by flowing argon in the central tube in the torch. Argon
also flows in the annular region of the torch. The two argon flows are balanced to
give a stable plasma, which is driven by radiofrequency induction from the load
coil. Analyte ions are formed by collisions with argon ions and other species in
the plasma; these are subsequently drawn into the mass spectrometer through a
differentially pumped gas inlet. Typical ionization efficiencies range upwards to
100%, but this figure is misleading as the transport of the ions from atmospheric
pressure to the vacuum of the mass spectrometer is inefficient. Overall efficien-
cies are typically 0.1% or less (ions detected compared to atoms introduced into
the plasma). Nominal voltages are near ground for a quadrupole mass analyzer,
whereas for a magnetic sector analyzer the entire ICP source (torch and vacuum
interface) is held at high voltage (thousands of volts).
Secondary ionization mass spectrometry (SIMS) is used mainly for the analysis
of solid compounds and materials. A focused ion beam (called the primary ion
beam) is directed against a surface containing the analyte of interest and ions are
formed by the sputtering action of the impacting ion beam. Figure 17.3 shows a
SIMS source, the actual dimensions of which are exaggerated for clarity. Note
that some of the ions from the incoming beam remain embedded in the sample.
Secondary ions are extracted by ion optics and accelerated into a mass analyzer
for measurement. The sample is normally at high voltage for efficient extraction
368 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
FIGURE 17.2. ICP source. Figure from Gray and Date (1983), Fig. 1.2.
and focusing of the secondary ions. Typical ionization efficiencies range from
<0.1 to about 1 percent. Virtually the entire periodic table can be analyzed with
this method. Both positive and negative ions can be produced, depending on the
composition and polarity of the primary ion beam.
Laser ablation can be used to volatilize samples. An intense laser pulse is focused
onto a solid at sufficient pulse power and energy (e.g., milliJoule energy in a pulse
of ∼10 nanosecond, or less, duration) to remove material from the surface. Typical
conditions result in crater formation after one or a few laser pulses. If performed
in flowing argon at atmospheric pressure, the ablated material can be fed into an
ICP for ionization. Ions may also be formed by an ablation pulse in a vacuum and
directly focused and extracted into a mass analyzer. Laser ablation/ionization is
used primarily for analysis of solid compounds and materials. Virtually the entire
periodic table can be analyzed with this method.
Resonance ionization mass spectrometry (RIMS) uses absorption of narrow
bandwidth laser light to ionize an element of interest. The wavelength is chosen
17. Mass Spectrometric Radionuclide Analyses 369
FIGURE 17.3. SIMS source. Figure courtesy of G. Gillen, National Institute of Standards
and Technology, http://www.simsworkshop.org/graphics.htm (Jan. 2006).
to coincide with an allowed electronic transition of the atom (or isotope) of in-
terest. Single photon transitions are exploited as well as “multi-photon” transi-
tions. Isotopic selectively can be achieved with sufficiently narrow wavelength
tuning. This method offers very high selectivity compared to other ionization
methods, but usually with relatively low ionization efficiencies (<1%). Under
specialized conditions, such as specific ion source geometry, narrowband ex-
citation by continuous-wave lasers, rapid heating, pulsed laser operation, and
well-matched laser diameter to sample size, much higher ionization efficien-
cies are possible. RIMS systems usually are found only in laser research lab-
oratories because the lasers needed for resonance ionization are complex and
expensive.
A notable exception regarding cost and complexity is the method of resonant
laser ablation (RLA). In RLA, a low-fluence pulsed laser beam is focused onto
the solid sample (Eiden et al., 1994). The laser pulse first ablates or desorbs a
small amount of sample. On the timescale of the laser pulse (typically a few
nanoseconds), the atoms liberated from the surface absorb laser photons and are
ionized. By using a laser wavelength resonant with an atomic transition in the
atoms of interest, ionization is highly selective. This method has been extensively
demonstrated with relatively low-cost YAG pumped dye lasers.
The ions produced by any of the preceding sources are then detected by the
mass analyzers described below.
370 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
FIGURE 17.5. Magnetic sector analyzer. Figure from Platzner (1997), pg. 5.
where:
R = radius
m = ion mass
e = electron charge
z = ion charge (in units of electron charge e)
V = ion accelerating voltage
B = sector magnetic field
A single sector can achieve abundance sensitivity (i.e., the ability to detect a
small abundance mass peak adjacent to a large abundance one) in the range of
1 ppm (i.e., an isotope ratio of 1 × 10−6 ). Abundance sensitivity can be critical for
rare isotopes; an example is 236 U, which occurs only at extremely low abundance
(ratio relative to 238 U < 10−10 ) in natural U, but is produced by irradiation of
uranium in a reactor (236 U abundances vary from 10−10 to 10−2 compared to the
abundances of all U isotopes present). Common sector geometries include 60◦ and
90◦ angles, and common radii include 15 and 30 cm.
A larger radius leads to higher dispersion of the masses (e.g., more separa-
tion between adjacent masses). High dispersion leads to better mass resolution
and abundance sensitivity and enables the use of larger detectors and ion sources,
17. Mass Spectrometric Radionuclide Analyses 373
which simplifies the design and construction of some component. High disper-
sion results in a larger overall instrument because the size of the instrument scales
directly with the sector radius. This is especially true of the magnet, which for
large instruments can weigh hundreds to thousands of kilograms. Large electro-
magnets are expensive and require large power supplies and control electronics
for stable operation. Modern commercial sector instruments are designed with so-
called extended geometries, which use specialized magnetic focusing to achieve
the equivalent dispersion of a large radius sector with a smaller magnet. Extended-
geometry instruments offer high dispersion at smaller size.
In its simplest form, a mass spectrum is created by scanning the magnetic field
and measuring the ion intensity. Alternatively, the magnetic field is held constant
and the ion source voltage (which controls the ion energy) is scanned. In modern
instruments, peak switching is performed by setting either the magnetic field or
the ion source voltage to values that correspond to the mass peaks of interest.
Sectors have the advantage that multiple ions can be detected simultaneously with
multiple detectors for each mass of interest or with an imaging detector. Early
mass spectrometers used a photoplate for ion detection to measure the entire mass
range at once, but photoplate ion images were difficult to measure quantitatively
for relative ion intensities. In modern instruments, electronic detectors (see Section
17.1.4) are used to measure ions directly.
Sector mass analyzers offer many advantages, including high precision iso-
topic analysis, simultaneous mass detection (with multiple detectors), and stable
operation. Disadvantages include large size, magnets that are heavy and require
sophisticated electronics to operate in a stable manner, and high voltages of more
than 10,000 V.
The quadrupole mass filter (QMF), shown in Fig. 17.6, represents an entirely
different approach to selecting ions by mass. In its typical configuration, the QMF
consists of four parallel cylindrical rods that are set on a square. The opposite
corners are connected electrically to both DC and RF voltages. Ions traverse along
the axis of the rods in a complicated motion and only ions of a selected ion mass-
to-charge ratio can successfully transit the quadrupole. The (a) portion of the
figure shows the conceptual layout, with electrodes with perfect hyperbolic cross-
sections. The (b) portion shows a schematic of the QMF in practice with electrodes
of circular cross-sections. Also shown are the ion source lens, collector, and a
generalized version of the drive electrical circuitry. The QMF uses a combination
of DC and RF electrical potentials to filter ions within a narrow range of m/z. In
effect, the QMF is operated as a bandpass filter.
The equations governing the operation of the QMF are derived from solutions
to Laplace’s equation, which for the geometry of the QMF are described by the
Mathieu equation (Duckworth et al., 1986; March and Hughes, 1989). The fol-
lowing discussion of the Mathieu equation is adapted from March and Hughes
(1989).
d 2u
+ (au − 2qu cos 2ξ )u = 0
dξ 2
374 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
FIGURE 17.6. Quadrupole mass filter (QMF). Figure from March and Hughes (1989),
pg. 3.
where
ωt
ξ=
2
8eU
au = ax = −a y = (17.2)
mω2r 2
4eV
qu = qx = −q y =
mω2r 2
17. Mass Spectrometric Radionuclide Analyses 375
FIGURE 17.7. Mathieu stability diagram. Figure from March and Hughes (1989), pg. 50.
where:
The solutions to the Mathieu equation translate into a series of regions of stable
motion for various values of a and q, which depend on the parameters r , m, V ,
U , and ω. A common approach to visualizing the operational parameters of the
QMF comes from the use of a Mathieu stability diagram (March and Hughes,
1989), which plots a versus q. Figure 17.7 shows a Mathieu stability diagram,
with the operating lines for various resolutions, R (which is M/M), overlayed
on the diagram. For the line marked “R = 10,” only mass M2 is stable and will be
transmitted through the QMF; the other masses will be ejected.
The apex of the stability curve for the primary stable region occurs at a y =
0.23699 and q y = 0.70600. For typical QMF analyzers, a given mass m is analyzed
by use of set values for the parameters r and ω (r and ω typically 1 cm and 1–3 MHz,
respectively) and then varying the parameters U and V (the DC and RF voltages,
respectively). Mass selection is achieved for the desired mass resolution, R, given
376 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
Electron beam
source
Endcap electrode
Sample pin
aperture
Ring electrode
Laser aperture
Channel electron
multiplier
FIGURE 17.8. Schematic diagram of an ion trap mass spectrometer analyzer. From Gill et al.
(1991), Fig. 1a.
in the usual form as m/m, where m is the desired mass range for transmission
through the QMF. A mass spectrum is created by scanning the DC and RF voltages
U and V , respectively, while maintaining a constant ratio of U/V . Resolution is
controlled by how close to the apex one scans the U/V ratio.
QMF analyzers offer advantages of light-weight, small size, and rapid response
(an entire elemental mass spectrum can be scanned in less than a second). They
find application in low resolution (∼1 amu), low abundance sensitivity (>1 ppm),
and low measurement precision (∼1%) applications (exception: precision of 0.1%
can be achieved with state-of-the-art ICP/MS instruments with QMF analyzers).
QMF analyzers have distinct advantages over magnetic sectors as the QMF does
not require a large magnet or high voltages, and the QMF is more compact. A
disadvantage of the QMF is that it operates as a bandpass filter and can essen-
tially measure one ion mass at a time, whereas a magnetic sector is capable of
simultaneous multi-mass detection.
The RF quadrupole ion trap mass spectrometer (ITMS) is a close relative of
the QMF and ideally can be thought of as a three-dimensional quadrupole (see
Fig. 17.8). The close relationship of these two devices is evident by the fact that
ion motion in the two devices is governed by essentially the same mathematical
equations. As with the QMF, the ITMS uses DC and RF electric fields and the
operation of the IT is described by solutions to the Mathieu equation. Unlike the
QMF, ITMS analyzers trap ions within the mass analyzer. Ions are trapped, ejected
to select the mass of interest, and then ejected in a controlled manner for detection.
17. Mass Spectrometric Radionuclide Analyses 377
FIGURE 17.9. Time of flight MS analyzer. Figure from Larsen and McEwen (1998), pg. 20.
Modern ITMS analyzers utilize milli-Torr pressures of helium to cool ions, which
enables higher resolution operation and long (millisecond to second) storage times.
The ITMS is applied in organic and biochemical analysis, and in a limited fashion
for isotopic and elemental analysis. The ITMS in Fig. 17.8 is configured to use a
laser pulse to vaporize and ionize metal atoms inside the mass analyzer. Unique
features of ion traps include the ability to store selectively specific m/z ratios (ion
species).
Many other kinds of ion traps are known, but apart from the ITMS, only the ion
cyclotron resonance (ICR) trap has been used for elemental or isotopic analysis. In
an ICR trap, the ions are trapped by a combination of RF and DC fields applied to
the walls of a box. The box is held inside a very strong magnetic field and the com-
bined RF and magnetic fields result in ions undergoing cyclotron motion whose
frequency depends on the ion’s mass-to-charge ratio. The mass spectrum is deter-
mined by measurement of these characteristic frequencies. First, the time domain
signal associated with ion motion is measured. The mass spectrum is generated by
inverting the time domain waveform to a frequency domain representation with a
Fourier Transform. The method is thus known as Fourier transform, ion cyclotron
resonance mass spectrometry or FT-ICR-MS. A key feature of FT-ICR-MS is
extremely high resolution.
The time-of-flight (TOF) mass spectrometer utilizes a pulsed ion beam for mass
analysis. In practice, a packet of ions is accelerated to the same energy and injected
into a field-free drift tube (Fig. 17.9). At constant energy, light ions move faster
than heavy ones, and mass separation is achieved by measuring the arrival times
of ions at the ion detector. Given a constant drift distance L (the length of the drift
tube from ion injection to ion detector), the time, t that an ion of m/z requires to
travel the distance L is given by:
t = L (m/z)/2eV (17.3)
378 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
where:
m = ion mass
e = electron charge
z = ion charge (in units of electron charge e)
V = ion accelerating voltage
L = drift tube length
v = ion velocity
Figure 17.9 shows a schematic diagram of a reflectron-type TOF used for laser
ablation mass spectrometry. Ions are generated in the source region, which consists
of a sample mount between two parallel plates (the plate on the right has a small
hole for transmitting ions into the drift tube on the right). The laser generates a pulse
of ions that are quickly accelerated into the larger drift region on the right. Lighter
ions move faster than heavier ions. The reflector (also known as a reflectron, see
discussion below) diverts the ions towards the ion detector at the bottom. A mass
spectrum is produced initially as a plot of ion arrival time versus ion intensity. The
laser ablation system is well-matched to the TOF since DC voltages can be used
for ion extraction and acceleration. In TOF systems that have a continuous source
of ions, pulsed electronics must be used to generate the rapid pulse of ions for the
TOF.
The TOF analyzer is capable of measuring a wide mass range of ions, but only
on a pulsed basis. In simple designs and operating modes, the spread in ion energies
must be small (<1%) and the ions must initially be tightly packed (within 1 mm or
less), otherwise the resolution of the TOF will be severely degraded. Methods are
applied for compensating for both initial ion energy spread and spatially extended
ion sources. The duty cycle (the fraction of time that ions are being generated
and analyzed) is usually short, hence the overall throughput of a TOF is low. For
longer duty cycles, specialized high speed pulse generators and high throughput
data acquisition systems are used (such systems were used for a brief time in a com-
mercial ICP-TOFMS instrument). Because data are acquired with high-speed dig-
itizers (typically 1 × 109 samples/sec) with deep memories (>1 × 106 samples),
digitizing resolution is usually limited to 8 bits. If signal averaging can be used,
the effective resolution can be improved, but the usual 8-bit data severely limits
abundance sensitivity of the overall instrument. The abundance sensitivity is thus
much less than other mass analyzers, typically in the 100–1000 ppm range (i.e.,
isotope ratios less than 10−4 are unresolvable).
Mass resolution is dependent upon drift tube length, ion energy, extractor design,
and timing resolution. High resolution is easily achieved, but TOF analyzers are
not often used for applications that require high mass resolution. A variant of the
TOF analyzer, called the reflectron, uses a gradient electric field to reflect the ions
back to an ion detector located near the ion source. The reflectron has energy-
focusing properties that correct for the energy spread (i.e., velocity distribution)
of the ions to result in higher mass resolution.
In a TOF analyzer, a mass spectrum is created by measuring the ion intensity
as a function of time of arrival after the ions are injected into the TOF analyzer.
17. Mass Spectrometric Radionuclide Analyses 379
Ions may be pulsed either by pulsing the ion source itself (e.g., a pulsed laser)
while the TOF-MS ion extraction electrodes are held at constant DC potential, or
by pulsing the ion extraction electrodes. Pulsing the electrodes is difficult because
the pulse rise time must be short (typically <10 nanoseconds) and the potentials
are high (typically 1–5 KV). Depending on whether the ions to be analyzed are
formed in a continuous ion source (hot filament, continuous plasma) or in a pulsed
ion source (laser, pulsed glow discharge), the TOF analyzer may detect all of
the ions formed or just a portion of the ions formed. If all the ions formed are
contained in the extraction region of the TOF-MS, as in a typical laser source,
then they can be accelerated and delivered to the detector with high efficiency. The
TOF-MS analyzer is capable, in principle, of simultaneous detection of different
isotopes. In practice, isotope ratio precision is limited by the coupling between the
ion extraction process and the continuous ion source and by the poor digitizing
resolution and analog detection schemes (i.e., not ion counting).
TOF analyzers offer advantages of compactness (they can be made very small—
cm-length analyzers are possible), simple construction (the TOF is basically a
tube), and their wide mass range (they obtain the full mass spectrum, usually
1–300 amu for elemental and isotopic analysis) for every measurement. Draw-
backs include the difficulty of efficiently coupling to various ion sources, limited
mass resolution, relatively low precision (1–10 % are typical) and the need to
use expensive high precision and high speed electronics for pulse generation, ion
detection, and data acquisition.
FIGURE 17.10. Faraday Cup ion detector. Figure A from Platzner (1997), pg. 34.
FIGURE 17.11. Electron Multiplier ion detector. Figure from (ETP 2005).
operational amplifier (op amp) that ultimately produces an “analog” signal. The
remainder of the electron beam in the dynode chain continues to the end of the
chain to produce a pulse counting signal. When the ion beam becomes too intense
to operate the detector in pulse counting mode (the anode current must be kept
below a critical value to avoid damaging the detector), the pulse counting part of
the detector is switched off (by switching the bias to one of the dynodes) and the
analog signal is used. The response of these two sections of the detector can be
cross-calibrated to yield an overall dynamic range in excess of 108 .
A variant of the EM is the Daly detector (Daly, 1960) in which ions are accel-
erated by ∼30 KV (this is called post-acceleration because it occurs after mass
analysis) into a conversion dynode, which generates a significant number (∼10)
of secondary electrons. These electrons are accelerated into a scintillator and con-
verted into light, which is detected with a photomultiplier. The Daly detector offers
high gain, low noise, and excellent stability. Other variants of the post-acceleration
detector exist; a simple configuration uses a metal plate to convert the ions into
electrons for detection with an EM. Another variant of the EM is the microchannel
plate (Coplan et al., 1984; Odom et al., 1990; Wiza, 1979). Microchannel plate
EM detectors have excellent sensitivity but poor gain stability. When operated in
382 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
an analog mode (measuring amplified ion current), the gain of some detectors can
drift by several percent during an analysis. Ion counting eliminates most gain-drift
problems. EM detectors have limited lifetimes due to the accumulated damage at
the ion conversion dynode from the impact of detected ions.
Besides the ability to count ions, the EM detector responds to changes in the
ion beam intensity instantaneously (on a nanosecond timescale), whereas the high
feedback resistance associated with FC detectors results in a long time constant for
the measurement (typically 1 to 10 seconds). Fast scanning mass analyzers such
as QMF, ITMS, and TOF analyzers almost always use EM type detectors because
the FC response is too slow.
17.2.5. Vacuum
Mass spectrometers require various degrees of vacuum in different parts of the
instrument. A vacuum is produced by a range of vacuum pumps, including rotary
vane, diaphragm, scroll, turbo-molecular, diffusion, ion, and getter pumps. Most
ion sources operate at low pressure, some at high vacuum, while the ICP operates
at atmospheric pressure. The mass analyzer and ion detectors of a mass spectrom-
eter almost always operate in high or ultrahigh vacuum. Because gases flow and
behave differently in these various vacuum/pressure regimes, understanding vac-
uum system design and operation requires some familiarity with the properties of
gases in such systems. An excellent reference for vacuum systems is Dushman
and Lafferty (1962).
The various mass spectrometer performance requirements drive vacuum sys-
tem design. These include correct operation of the ion source (arcing can occur
in the high voltage components if the operating pressure is too high), and good
transmission of ions through the ion optics and mass analyzer (scattering by back-
ground gases reduces transmission). Rough vacuum is required on the foreline of
high-vacuum pumps such as turbo-molecular pumps and diffusion pumps and in
the interface between atmospheric pressure and the vacuum of an ICP/MS. Most
modern instruments use oil-sealed rotary vane pumps for these applications, al-
though oil-free or so-called “dry” pumps (also known as scroll and diaphragm
pumps) are used because of their inherent cleanliness. Dry pumps do not insert
oil vapor into the vacuum system, a high-cost component of a high-performance
mass spectrometer system.
ion beam changes) and instrument tuning (changing the potential of an optical
element affects the total ion flux and its spatial distribution). These effects have
been recently discussed by Praphairaksit & Houk (2000).
The ICP ion source has been utilized in combination with most types of mass
analyzer (RF quadrupole ion filters and traps, time-of-flight, and magnetic sector
field), although commercial instruments utilize either quadrupoles or magnetic
sectors as mass analyzers. The observed mass bias is comparable in these differ-
ent instruments (Encinar et al., 2001). The precision attainable does not depend
(strongly) on the mass analyzer employed per se, but rather on whether the instru-
ment is a multi-collector/detector instrument or not. The attainable precision with
single detector instruments is much more limited than what can be achieved with
multi-collectors (MC). Subtle corrections are, therefore, of much greater interest in
the latter instrument type. Mass bias corrections are derived by measuring the bias
with an isotopic abundance standard material and fitting measured bias to a func-
tion. In most quadrupole instruments, a linear fit suffices. In higher-accuracy work
(MC-ICP/MS), the functional forms are non-linear (power law or exponential).
The isotopes used for mass bias correction can be measured in the same solution
as the sample (for internal bias correction) or in a separate solution (for external
bias correction) when correction isotopes and the sample isotopes of interest may
overlap.
Extrapolation of the bias function from the masses of the reference isotopes
to the masses of the analytes of interest is typically kept as small as is practical.
Ideally, the analytes of interest are bracketed by the mass bias calibration isotopes,
i.e., the bias correction is interpolated. For example, a Tl isotopic standard is used
to correct unknown Pb ratios, but Tl would not be a good choice for correction
of Pu isotopes. If appropriate isotopic abundance reference materials are available
and appropriate care is used, MC-ICP/MS is among the most versatile (applies
to most of the Periodic Table) and accurate methods for determining unknown
isotopic composition.
counts are required; this level or better is the precision/counts regime of most
modern mass spectrometry. External precision is determined by running duplicate
samples. External precision is a measure of internal precision combined with how
reproducibly sampling and sample preparation are performed.
Accuracy of a method of analysis generally is determined by analyzing standards
with matrices the same or similar to the samples of interest. These standards have
known isotopic abundances and/or concentrations. The limit to this approach often
is the accuracy of the reference technique that was used to characterize the standard.
where:
Ns = the number of atoms of analyte in the sample
Nt = the number of atoms of the tracer added to the sample
Rm = the measured ratio of isotopes A to B for the traced sample
At , Bt = the atom fractions of isotopes A and B in the tracer
As , Bs = the atom fractions of isotopes A and B in the untraced sample
All relevant isotopes of the analyte element are monitored and the amounts of the
unknown isotopes are referenced to the known amount of isotope tracer. If the tracer
isotope is an isotope of interest in the sample, both traced and untraced sample
analyses are performed, the former analysis providing elemental concentrations
and the latter providing isotope ratios.
Equation (17.4) is a general form that is applicable to any element and tracer
combination. For quantitation of radionuclides in which isotopically pure tracers
are available that are not present in the sample at significant abundance, a simplified
version of Eq. (17.5) will suffice:
N s = N t × Rm (17.5)
where:
Ns = the number of atoms of analyte isotope in the sample
Nt = the number of atoms of the tracer isotope added to the sample
Rm = the measured ratio of analyte to tracer isotope A for the traced sample
For the case in which the tracer also contains the radionuclide of interest,
Ns = Nt × (Rm − Rt ) (17.6)
where:
Ns = the number of atoms of analyte isotope in the sample
Nt = the number of atoms of the tracer isotope added to the sample
Rm = the measured ratio of analyte to tracer isotope for the traced sample
Rt = the ratio of analyte to tracer isotope in the tracer
Equations (17.5) and (17.6) are commonly used to quantify radionuclides by
isotope dilution mass spectrometry.
multiplier (EM), Faraday cup (FC), and Daly-type (post-acceleration and scin-
tillation). These systems are used alone or in combinations of “multi-collector”
instruments. In addition, researchers have coupled these and other mass analyzers,
such as Fourier transform-ion cyclotron resonance, and multiple analyzers as in
quadrupole-quadrupole configurations, to the ICP.
The power of this MS technique has driven the development of methods to
interface ICP/MS instruments with various sample introduction systems. Special-
ized sample introduction systems include ion chromatography (Seubert, 2001), gas
chromatography (Vonderheide et al., 2002), and capillary electrophoresis (Costa-
Fernandez et al., 2000). Other techniques are hydride generation (used to volatilize
selected species and obtain some matrix/elemental separation) (Reyes et al., 2003)
(Bings et al., 2002) laser ablation (Gonzalez et al., 2002; Heinrich et al., 2003;
Russo et al., 2002), and electrothermal vaporization (Richardson, 2001; Vanhaecke
and Moens, 1999).
on factors such as the ionization potential of the atom and the temperature of the
plasma. This source has an ionization efficiency of around 95% with high excita-
tion temperatures for inducing optical emission for the majority of the elements in
the periodic table. Because of the high efficiency of ionization, researchers such
as Houk and Fassel (1980) developed the ICP as an ion source for mass spectrom-
etry. Early attempts were unsuccessful due to formation of a boundary layer that
obstructed the ion-sampling aperture. Progress in overcoming this complication,
among other key developments, led to the powerful technique in use today.
The key to the development of ICP/MS was a technique to transfer ions from the
plasma into the high vacuum of the MS. This was accomplished with a differen-
tially pumped atmospheric-pressure-to-vacuum interface shown in Fig. 17.12. An
aerosol is injected into the central argon flow, where aerosol particles are desolvated
and vaporized. The ions created in the plasma are drawn into the MS at the vacuum
interface. The plasma density and temperature permit isotropic flow through the
first orifice (typically ∼1 mm diameter aperture), and the plasma undergoes free
jet expansion (Scoles, 1988) in the first vacuum chamber.
This free jet structure is “skimmed” as it passes through a second aperture
into the first high vacuum stage. The interface contains a skimmer cone (made of
temperature-resistant alloy) with a small orifice through which the plasma (consist-
ing of positive ions, electrons, and neutral gas) is drawn into the vacuum interface.
High capacity vacuum pumps remove the gas, while ions are electrically injected
into the mass spectrometer using ion optical elements and accelerating voltages.
The skimmed beam is accelerated and focused by ion optical elements (typ-
ically, cylindrically symmetric lenses). The strong electric field in these lenses
(∼ 1 keV/cm) leads to rapid charge separation in the skimmed beam. That is, the
17. Mass Spectrometric Radionuclide Analyses 389
electrons are rapidly repelled by this field and the positive ions are accelerated and
focused. The intensity of the beam is such that once the electrons are removed,
the self-repulsion of the positive ions due to space charge effects (the so-called
space charge explosion) leads to significant loss of ion beam intensity, an important
source of analyte loss. Loss of ion current in this atmospheric-pressure-to-vacuum
interface is the main contributor to low efficiency in ICP/MS.
TABLE 17.4. ICP/MS precision for various instruments and sample amounts
Isotope ratio Method∗ Precision Sample size Efficiency Reference
∗∗235 U/238 U ∼1 ICP/QMS 0.028% 255 pg Not Avail Platzner
∗∗235 U/238 U ∼1 DF-ICP/MS 0.026% 255 pg Not Avail Becker
238 U/235 U = 137.9 ICP/QMS 0.07% 663 pg 0.13% PNNL
238 U/235 U = 31.8 ICP/QMS 1.4% 307 fg 0.15% PNNL
235 U/234 U = 160.5 ICP/QMS 3.8% 9.42 fg 0.15% PNNL
236 U/234 U = 1.05 ICP/QMS 7.4% 324 ag 0.16% PNNL
∗ QMS: QMF mass analyzer, DF-ICP/MS: double-focusing sector-field mass analyzer
∗∗ From (Becker and Dietze, 2000).
ions and the element A molecular ions. In this case, interference by A in B can
be corrected because an AO+ ion exists that does not overlap any of the B isotope
masses. From the intensity of the lightest isotope and the abundance of each of the
three isotopes of A, one can calculate the contribution of the other AO+ isotopes
to the signals observed at the B isotope masses.
Each of the major types of MS instruments for radionuclide measurements has
unique mechanisms by which isobaric interference are formed within them. In an
ICP/MS, the most common interference is from atomic isobars and the diatomic
molecular oxide, hydride, nitride, and argide ions. In the typical argon support gas
for the ICP, argide-based interferences abound, from 40 Ar interfering with 40 Ca
to 195 Pt40 Ar interfering with 235 U. The usual aqueous aerosol introduced into the
ICP provides oxygen and hydrogen, which form oxides and hydrides. Examples
of these interfering ions are listed in Table 17.5 with the analyte(s) of interest and
the mass resolving power needed to separate them.
Considerable effort has been expended to reduce interference without losing
analyte signal. These efforts include chemical removal of interference-causing
atoms prior to mass analysis, with sample introduction by hydride formation at
elevated temperature, and with selective removal of water vapor from the sample
aerosol (Montaser, 1998). Reduction of molecular ion interferences involving high
ionization potential (IP) atoms, such as argides, has been achieved by reducing the
plasma temperature (via lower RF power). Reduction of plasma temperature re-
duces the intensity of high-IP species much more than that of lower-IP species. This
“cold” plasma has been shown to reduce Ar+ and ArO+ faster than the reduction
in intensity of the species with which they interfere (Ca and Fe, respectively).
ion source and the mass analyzer. The reacting ions undergo a change in mass-
to-charge ratio in the cell. If the interfering ions react at a much greater rate
than the ions of interest, the analyte of interest is detected with reduced back-
ground. Typically, interferences can be reduced by three to nine orders of magni-
tude; there is some accompanying loss of analyte, from a few percent to 10-fold
loss.
Collision/reaction cells used in ICP/MS consist of an enclosure to contain the
gas, apertures at each end to allow ions in and out of the cell, and a set of electrodes
(e.g., octopole, hexapole, or quadrupole rod sets) to guide ions through the cell.
The main considerations in the use of the method are the reagent gas (whether to
use a “buffer” gas), gas number density, ion kinetic energy, cell length, multipole
design, use of auxiliary fields (in addition to the ion guiding field), and reaction
time. Other considerations include reagent gas purity, side reactions, and chemical
resistance of the sample ions (including desired analyte, interferences, and matrix
ions) to the reagent gases. The operating principles and design of RF multipole
ion guides and gas collision cells have been described in detail (Gerlich, 1992).
The dramatically different reactivities of gas phase ions provide an opportunity to
reduce or eliminate many isobaric interferences. Identification of suitable gaseous
reagents is aided by the advances in the understanding of gas phase ion-molecule
chemistry achieved in recent years (Armentrout, 2004). Hydrogen gas is the most
selective reagent gas currently in use. Other reagent gases, including NH3 , H2 O,
CH4 , and O2 , have been found to be much less selective, and usually require that
selective ion storage/reaction methods be employed to avoid generation of new
interferences.
Collision/reaction cells are typically operated at the lowest possible kinetic en-
ergy for several reasons. First, the benefits of collisional focusing are realized at low
17. Mass Spectrometric Radionuclide Analyses 393
kinetic energy. Second, ions at lower energy remain longer in the cell and have more
opportunity to react. Last, lower energy enables control over endothermic reactions.
Most manufacturers of ICP/MS instruments now offer a model that incorporates
a collision cell. Interference reduction by sample purification, use of a collision cell
(also called chemical resolution), and mass separation—are said to be orthogonal,
i.e., gains in each method are independent and multiplicative with gains in the
others. The ultimate instrument in this regard—a collision cell equipped, high
resolution, magnetic sector MS—has not yet been built, but would enable the
nearest approach yet to detection without interference of any isotope at ultra-trace
levels.
material made of high work function metals and for analyte species with low ion-
ization potentials. Most metals have work functions in the range of 4 to 5 eV,
which effectively limits TIMS to elements (or molecular species) with ionization
potentials less than 8 eV. In practice, TIMS is generally used for elements with
ionization potentials of 6 eV or less. Rhenium is the filament material of choice
because of its high work function (4.98 eV) and high melting point (3180 C).
Commercial rhenium is available in ultra-high purity (99.999% or better), which
minimizes unwanted ion emission. Other metals commonly used for filaments in-
clude tantalum and tungsten. Although platinum has the highest work function
(5.13 eV), its relatively low melting point (1772 C) limits its utility for TIMS use.
Although primarily associated with positive ions, negative TIMS is also possible.
The governing equation (17.8) is similar:
N− E A−W
= Ae kT (17.8)
N0
where:
Unlike the positive ion case, negative TIMS is most efficient with low work
function surfaces. Such surfaces are created by coating the ionizing filament with a
low work-function material such as LaB6 or Ba(NO3 )2 . Elements with high electron
affinities (hence easily ionized) include the halogens, selenium, and tellurium.
Oxides of some metals (e.g., Mo, Os, Re) also have high electron affinities and
also have high volatilities, especially when compared to their reduced, metallic
forms. Since the mid 1980s, negative TIMS has become the method of choice for
high precision and high sensitivity isotopic analysis of refractory Group VIb, VIIb
and VIII elements. For instance, Heumann (1995) developed and demonstrated a
viable method for analyzing the geochronometer Re-Os using negative TIMS. The
Re-Os system has also recently been analyzed by MC-ICPMS (Malinovsky et al.,
2002; Yin et al., 2001).
Whether the ionization is positive or negative, TIMS requires careful sample
preparation, often involving considerable chemical processing to separate and pu-
rify the element of interest. TIMS finds applications in geoscience, environmental
analysis, cosmochemistry, biosciences, medicine, material science, and physics.
Samples generally include soil, minerals, meteorites, and biological tissue. More
information on the specifics of the TIMS technique and its applications is in the
monograph by De Laeter (2001).
17. Mass Spectrometric Radionuclide Analyses 395
17.5.1. Instrumentation
At the heart of the TIMS ion source are one or more hot filaments that serve
to vaporize and ionize atoms or molecules of interest. Once generated, the ions
are accelerated, focused, and directed into the mass analyzer for measurement.
The classic TIMS instrument consists of an ion source, a single magnetic sector
mass separator, and an ion detector. Such an instrument is capable of measuring
isotope ratios as small as 1 × 10−6 , sufficient for the isotopic analysis of most
elements. For radionuclide analysis, smaller isotope ratios are often encountered.
Specialized mass spectrometers include multiple magnetic and electric sectors
and sector instruments with retarding quadrupole lenses (Smith, 2000) to measure
down to the 10−9 range.
TIMS instruments tend to be table-top-sized, with total path-lengths from ion
source to detector in the 1 to 2 m range. Popular sector geometries include 60◦ and
90◦ deflection, which is the nominal deflection produced by the magnetic field.
The deflection radius is normally in the 10 to 30 cm range. Typical operating pa-
rameters for actinide analysis are 10 kV acceleration voltage and ∼0.5 to 1 Tesla
magnetic field to produce ion beams that are separated by mm-scale distances (1
amu mass difference at 238 amu ion beam mass). Ions are detected by either a
Faraday cup or an electron multiplier. Because of the small mass separation, de-
tector size can be a critical issue, especially for simultaneous detection of multiple
ion masses.
Historically, most TIMS instruments were equipped with a single ion detector.
In operation, the ion mass region of interest is either scanned or peak-stepped, by
changing either the ion accelerating voltage or the magnetic field. Peak-stepping
requires very stable electronics and magnet control (if an electromagnet is used) to
select accurately and with good repeatability a suite of ion masses. The ion beam
must also be of high quality and reproducibility.
A common requirement for TIMS instruments is to have flat-topped ion peaks.
Flat-top peaks are produced when the ion beam is focused to a fine line and
the analyzer slit (usually the slit in front of the detector) is wider than the ion
beam; therefore, the entire ion beam is captured by the ion detector. Flat-top
peaks are desired for high precision isotopic analysis because the measured ion
beam signal is insensitive to slight instabilities in the instrument magnetic field,
which causes the beam to shift position. Flat-top peaks are achieved at the cost
of reduced mass resolution, but such reductions are usually insignificant com-
pared to the gain in isotopic measurement precision. Figure 17.14 shows the
TIMS peak shape, acquired by high-resolution mass scans for several neodymium
isotopes. Note that several lines are slightly shifted from each other. The data
were acquired with an Isoprobe-T instrument and multi-collector FC detection
system.
Modern TIMS instruments are equipped with multiple ion detectors (see Fig.
17.15). In the 1980s, Faraday cup arrays became commercially available, and
these provided significant improvements in isotopic precision and sample utiliza-
tion. In the 1990s, arrays of pulse counting ion detectors with very compact EM
396 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
or microchannel plate technology (De Laeter, 1996; De Laeter, 2001) were devel-
oped for high sensitivity and high precision isotopic analysis. These are becoming
commercially available.
TABLE 17.6. Performance Parameters for uranium and plutonium analysis by TIMS
Element Measured isotope ratios Detection limit Comment
U 234 U/238 U ∼1pg Detection limit is usually limited by
235 U/238 U background U
236 U/238 U
17.6.1. Instrumentation
The AMS consists of an ion source, an input mass or momentum selector (usually
a magnetic sector), a high–energy accelerator, a “stripper” (a low-pressure gas
or thin foil placed in the middle of the high voltage region), post accelerator
ion optics and mass selectors, and one or more detector systems. Figure 17.16
shows a generalized diagram. The high-energy accelerator is usually a tandem
accelerator, which consists of two back-to-back high-energy accelerators with a
stripper between the accelerators. The voltage increases from zero at one end
to the middle and decreases to zero at the other end. Low-energy negative ions
injected into one end of the accelerator emerge as high-energy positive ions at
the other end. This arrangement allows the ion detector region to be operated at
or near ground potential to simplify construction and electrical operation of the
instrument.
Ions typically are produced by a secondary ion or sputter source (the usual choice
for the AMS), often with a beam of Cs ions to sputter the sample to produce negative
ions. These ions are extracted, focused, and accelerated to energies of 10–100 keV.
17. Mass Spectrometric Radionuclide Analyses 399
FIGURE 17.16. Schematic diagram of an AMS instrument. Figure from De Laeter (2001),
pg. 62.
The input mass selector directs a specific ion mass into the tandem accelerator,
which then accelerates the negative ions to MeV energies. A stripper in the central
region of the accelerator dissociates molecular ions and removes multiple electrons
from the ions to convert ingoing negative ions (elemental or molecular) into out-
going multi–charged positive ions. These positive ions are further accelerated by
the same potential and emerge from the back-end of the accelerator with energies
that can reach above 30 MeV, depending on the accelerator base voltage (or poten-
tial) and the degree of ionization. Charge states of +10 or more can be produced,
especially for heavy ions such as the actinides. After acceleration, additional ion
optics and switching magnets are used to select the ion mass and charge state and
to direct these ions into detectors for measurement.
By its nature, an AMS instrument is large. Path-lengths from ion source to
detector can be tens of meters. The high ion energy results in large differences in
ion path angles and trajectories. After selection by mass, the ions are separated
at cm-scale distances, in contrast to mm-scale separations produced in sector-
based mass spectrometers. A Faraday cup detector is a common choice for the
major isotope mass, whereas nuclear particle detectors (which include electron
multipliers) count individual ions of the rare isotope. The high ion energies allow
the use of detectors that can measure ion time-of-flight, ion energy, ion energy loss
rate, and ion charge. Measurement of the ion charge/mass ratio provides additional
selectivity for identifying the rare isotope.
400 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
In operation, the major and rare isotopes are measured sequentially by peak
switching from one isotope to the other, as with other mass spectrometers, to de-
termine atom ratios. The major isotope may produce a current in the microampere
range, whereas the rare isotope produces a signal as low as a few counts per minute.
The raw isotope ratio consists of counted pulses divided by integrated current. Cal-
ibration by measurement standards with known isotope ratios provides the factor
to convert the raw ratio into an atom ratio.
Unique features of the AMS include ion energy in the MeV range compared
to keV range or lower for other mass spectrometers, conversion of negative ions
into positive ions, and generation of multi–charged ions. The rare and major iso-
topes may be measured in different charge states, depending on the element under
analysis, the need to eliminate isobaric interference, and the configuration of the
instrument. On the other hand, the high energy poses safety and health consider-
ations. The ions in an AMS have sufficient energy to produce radionuclides and
radiation by nuclear reactions. Ions that strike various system parts (e.g., slits,
deflection plates, the inner walls of the vacuum system) can produce neutrons and
x-rays. High voltages are associated with the ion source, deflection plates, and
other parts of the instrument. These features require strict attention to radiological
and electrical safety protocols and protection measures.
TABLE 17.7. Experimental Parameters for AMS, after Tuniz et al. (1998)
Measured Detection
Isotope Sample form isotope ratio limit Carrier Comments
14 C Graphite 14 C/12 C 5 × 10−16 None —
10 Be BeO 10 Be/9 Be 3 × 10−15 Sometimes —
26 Al Al2 O3 26 Al/27 Al 3 × 10−15 None —
36 Cl AgCl 36 Cl/35 Cl, 1 × 10−15 Sometimes, Both ratios measured
36 Cl/37 Cl Chloride
41 Ca CaH2 or CaF2 41 Ca/40 Ca 6 × 10−16 None —
129 I AgI 129 I/127 I 3 × 10−14 KI —
U U + Nb metal 236 U/238 U 1 × 10−12 Nb metal (Berkovits, 2000)
Pu Pu + Fe–oxide 240 Pu/239 Pu, <1 fg Fe–oxide Detection limit in
241 Pu/239 Pu, atoms; multiple
242 Pu/239 Pu ratios analyzed
tritium, 59 Ni, 90 Sr, and 205 Pb. Detection limits for AMS are typically given as
limiting isotope ratios, which is the measured quantity in AMS. The overall ion
transmission of a typical AMS is such that the absolute atom detection efficiencies
are lower than either TIMS or ICPMS—on the order of 1 count per 100,000 atoms
injected into the instrument. Nevertheless, the extremely low background of the
AMS technique permits the detection and quantification at the sub-femtogram-
level, which is comparable to ICP/MS and TIMS analyses. Some experimental
parameters are given in Table 17.7 for the most commonly analyzed isotopes.
For some radionuclides, only the direct measurement is needed, based on cali-
bration with a radionuclide standard and reference to the measured sample mass.
For other radionuclides, the isotope ratio to its stable element is needed. An exam-
ple of a more complex situation is measurement of the 14 C/12 C isotope ratio of an
environmental or archeological sample in comparison to the modern atmospheric
CO2 value. This value must be adjusted for anthropogenic 14 C produced by atmo-
spheric testing of nuclear weapons and by other nuclear operations, and also for
changes in atmospheric CO2 with cosmic-ray flux fluctuations over time. For an
element such as Pu, which has no stable isotope, the total quantity is measured by
isotope dilution mass spectrometry in which the sample is traced (or spiked) with
242
Pu or 244 Pu (see Section 17.3.3).
FIGURE 17.17. The Cameca IMS-6F secondary ion mass spectrometer. Figure from
De Laeter (2001), pg. 49.
and magnetic sectors to perform the mass separation, and TOF-SIMS, which uses
a time-of-flight mass analyzer. Both variants are available as commercial instru-
ments, as in the conventional SIMS instrument shown in Fig. 17.17.
The primary ion beam is formed on the left and accelerated and focused onto a
sample held at high voltage. Secondary ions are extracted and accelerated into the
analyzing mass spectrometer, which consists of an electrostatic sector (for energy
focusing) and a magnetic sector (for mass separation). Among the detectors used
to analyze the ions is an imaging detector. The ion optics on the Cameca instrument
are specifically designed to permit imaging of the sample with ions.
Production of secondary ions depends on many factors, but the composition of
the primary ion beam is vital to the efficient production of ions. Cesium or oxygen
ions (positive or negative, depending upon whether negative or positive secondary
ions are desired) are used (Cristy, 2000; De Laeter, 2001). Other primary ions
include noble gases, especially Ar, and more exotic species such as rhenium oxide
(Groenewold et al., 1997). The primary ion beam usually is mass selected with a
magnetic sector or quadrupole mass analyzer, and focused to a micrometer-sized
spot on the sample under analysis. Normally, the beam can be raster-scanned across
the sample with magnetic and electrostatic deflection in a manner similar to the
electron optics used in an electron microscope.
An advantage of SIMS is that the method provides micro-scale analysis of
the sample. The primary ion beam can be set to analyze a small area or raster-
scan across the sample. Some instruments are designed to produce ion images
analogous to images obtained in an electron microscope. The imaging capability
allows for high-resolution isotopic (and elemental) analysis of the sample, with
spatial resolution of 1 μm or better that depends on the instrument. The sputtering
17. Mass Spectrometric Radionuclide Analyses 403
action of the primary ion beam can be controlled to permit calculation of the
depth of the material removed during the analysis. Depth resolution to a few
nanometers is possible. Because of these micro-scale and imaging capabilities, a
SIMS instrument is sometimes referred to as an ion microprobe or ion microscope.
Production of secondary ions is non-selective in that all elements can be ionized
by this method. Thus, the entire periodic table is accessible using SIMS. The
relative and absolute efficiencies for the production of ions for a given element are
highly dependent upon the composition of the sample, as well as the configuration
and operating parameters of the instrument. Generally, isotopic compositions are
easily measured, whereas elemental compositions, which are derived from the
analysis of isotopes of a given element, are more difficult to determine. Good
quantitative analysis requires the use of calibration standards that closely resemble
the sample under analysis.
Preparation of samples for analysis resembles that used to prepare samples for
electron microscopy. In general, preparation consists of polishing small specimens
and mounting them on a planchet. A common planchet consists of a cylinder
of vitreous carbon with polished ends. The planchet must provide a conductive
pathway to drain the charge imparted by the impact of the primary ion beam.
Without some means to dissipate the charge, the sample will become charged,
which can deflect the primary ion beam and can interfere with the extraction of
secondary ions.
Operation of a SIMS instrument resembles both that of an isotope ratio mass
spectrometer and an electron microscope. Most SIMS instruments include an op-
tical microscope so that the sample can be directly viewed during analysis, which
allows for accurate positioning of the area of interest on the sample. Data can be
in the standard mode used for other types of mass spectrometers in which ions are
produced and the mass spectrum is analyzed by scanning or peak-hopping. This
mode is sometimes called the microprobe mode in SIMS. Another application for
SIMS is the acquisition of ion-images. This mode is called the microscope mode
because the SIMS is operated as an ion microscope.
from the ICP or by analyzing the ions produced in the ICP. This latter technique is
referred to as laser ablation ICP/MS. Numerous mass spectrometer analyzer types
have been used for LA-ICPMS including time-of-flight (TOF), quadrupole, sector
field, and ion trap. The analyzer type is determined by the application.
One of the problems that has been noted in laser ablation is fractionation between
the sample surface and the material removed. For any given sample type, certain
elements will be preferentially removed, giving these elements a higher relative
abundance in the material analyzed by the mass spectrometer. If fractionated, this
material is not representative of the sample, and biased data results. Research has
shown that LA-ICPMS with deeper UV wavelengths (e.g., 157 nm, 193 or 213 nm)
results in less fractionation. A pulsed laser with short pulse widths—on the order
of picoseconds to femtoseconds—provides even less fractionation.
The most common LA-ICP/MS work to date is with flash-lamp pumped Nd:YAG
lasers that produce a light pulse of 3–10 nanoseconds in width and are relatively
inexpensive. Shorter pulse width lasers (picosecond to femtosecond) are widely
available but more expensive. A wavelength other than IR or UV is used only for
resonant laser ionization or ablation. In this case, the goal is to excite or ionize a
specific element preferentially, and the laser wavelength is chosen to correspond
to a particular transition of an element, as discussed in Section 17.7.4.
The laser typically used for laser ablation is the frequency quadrupled Nd:YAG
(266 nm) with a pulse width of a few nanoseconds. With this laser, some frac-
tionation will occur. Unless a good external standard is used, quantification of the
elements in the sample is difficult if not impossible. This is not an impediment
to use of laser ablation in the field of radiochemistry or radionuclides. One of the
most important aspects of radionuclide analysis is determining isotope ratios of
elements of interest. All isotopes of a particular element behave the same when
removed from a solid sample by a laser pulse. Ionization of those isotopes can
depend on the laser bandwidth and polarization so care must be used when using
lasers for direct ionization of atomic species.
The purpose of LA-ICP/MS use is to remove material from the surface and trans-
port it into the ICP for ionization to obtain isotope ratios quickly with little sample
preparation. Inter-element isotope ratios are also important in certain radionuclide
applications, but are compromised by fractionation issues. Laser ablation provides
an overview of which analytes and isotopes are present and approximates the
concentration of each.
Literature reviews and primary literature are available for a more thorough
analysis (Durrant, 1999; Gunther et al., 2000; Russo et al., 2002; Winefordner
et al., 2000). (Gastel et al., 1997) measured long-lived radionuclides in a concrete
matrix using LA-ICPMS. The radionuclides investigated were 99 Tc, 232 Th, 233 U
and 237 Np. With a quadrupole mass spectrometer, detection limits on the order
of 10 ng/g were achieved while a double-focusing sector instrument was able to
deliver sub ng/g detection limits. (Gibson, 1998) used resonant laser ablation mass
spectrometry to investigate actinide oxides. Oxides of Th, U, Np, and Am were
imbedded in a copper matrix and analyzed by resonant LA-MS. Since actinides
are present as oxides in many common forms such as glass, ceramics, soils, and
17. Mass Spectrometric Radionuclide Analyses 405
others, the study was designed to determined whether resonant laser ablation would
provide any benefit over normal LA in the analysis of these elements. Due to the
many closely spaced energy levels of these elements, a significant advantage of
resonant laser ablation over normal laser ablation was not observed, but the study
showed the ability to use laser ablation for these types of samples.
Becker and Dietze (2000) used laser ablation to measure long-lived radionu-
clides in geological samples, high-purity graphite, and concrete. An effort was
made to improve quantification by using geologic standards and synthetic stan-
dards of the graphite and concrete. Solution nebulization was also used as a cali-
bration method. The isotopes studied were 99 Tc, 232 Th, 233 U, 235 U, 237 Np and 238 U.
The detection limits that resulted were in the low pg/g range. Boulyga et al. (2003)
used LA-ICP/MS to determine plutonium and americium in mosses. To improve
quantification, isotope dilution was used. Detection limits in the single fg/g range
were demonstrated for both elements.
Pajo et al. (2001a) used GD-MS to measure impurities in uranium dioxide fuel
and showed that these impurities could be used to identify the original source of
confiscated, vagabond nuclear materials. De las Heras et al. (2000) used GD-MS to
determine neptunium in Irish Sea sediment samples. The sediment samples were
compacted into a disk that was used with a tantalum secondary cathode in the
glow discharge. Using a doped marine sediment standard for calibration, detection
limits down to the mid pg/g level were determined.
17.8. Applications
Mass spectrometry finds applications in many fields. A partial list of fields includes:
r archeology
r bioassay
r biosciences
r cosmochemistry
r geochemistry
r health physics
r environmental monitoring
r nuclear non-proliferation treaty monitoring
r nuclear science
r radiochemistry
17.8.1. Uranium
ICP/MS and TIMS are the methods of choice for analyzing uranium isotopes. TIMS
can measure all of the long-lived isotopes of uranium, including 236 U, with high
precision (0.01 percent or better) and high sensitivity (10−12 g or less) with mul-
tiple collectors (Adriaens et al., 1992; Becker and Dietze, 1998; De Laeter, 2001;
Delanghe et al., 2002; Efurd et al., 1995; Pajo et al., 2001b; Platzner, 1997; Sahoo
et al., 2002; Smith, 2000; Stoffels et al., 1994; Taylor et al., 1998; Yokoyama
et al., 2001). The TIMS requires extensive chemical processing to isolate ura-
nium from the sample. In contrast, ICP/MS, typically with a quadrupole mass
analyzer—although multi-collector, sector-based ICP/MS instruments are becom-
ing increasingly popular—is used for various samples, often without chemically
separating the uranium. The ICP/MS can provide isotopic information at the 0.01
to 1 percent precision level for the major isotopes, 235 U and 238 U (Aldstadt et al.,
1996; Becker et al., 2002; Becker et al., 2004a; Becker and Dietze, 2000; Bellis
et al., 2001; Boulyga and Becker, 2001; Boulyga et al., 2000; Haldimann et al.,
408 John F. Wacker, Gregory C. Eiden, and Scott A. Lehn
2001; Kerl et al., 1997; Magara et al., 2002; Manninen, 1995; Platzner et al., 1999;
Schaumloffel et al., 2005; Uchida et al., 2000; Wyse et al., 1998). Recently, AMS
has been used for uranium analysis, especially for measuring rare isotopes such
as 236 U (Brown et al., 2004; Danesi et al., 2003; Fifield, 2000; Tuniz, 2001; Zhao
et al., 1997).
quadrupole MS (Wendt et al., 1997). Detection limits are in the 105 atom range at
90
Sr/88 Sr isotope ratios as small as 10−10 .
410
Appendix 411
7500 Ra Radium
A. Introduction
B. Precipitation Method
C. Emanation Method
D. Sequential Precipitation Method
E. Gamma Spectroscopy Method
7500 Rn Radon
A. Introduction
B. Liquid Scintillation Method
7500 Sr Total Radioactive Strontium and Strontium 90
A. Introduction
B. Precipitation Method
7500 3 H Tritium
A. Introduction
B. Liquid Scintillation Spectrometric Method
Regulation ID Description
ISO 9696:1992 Water quality: Measurement of gross alpha activity in non-
saline water (Thick source method)
ISO 9697:1992 Water quality: Measurement of gross beta activity in nonsaline
water
ISO 9698:1989 Water quality: Determination of tritium activity concentration
(Liquid scintillation counting method)
ISO 10703:1997 Water quality: Determination of the activity concentration of
radionuclides by high resolution gamma-ray spectrometry
ISO 15366:1999 Nuclear energy: Chemical separation and purification of ura-
nium and plutonium in nitric acid solutions for isotopic and
dilution analysis by solvent chromatography
Mr. Doug Ashley, Architect, with Bullock Tice Associates, Pensacola, FL, provided the information
given in this appendix.
Appendix C
420
Hydrogen Helium
3 6.94 4 9.01
Si symbol: transitional metals 5 10.81 6 12.01 7 14.01 8 15.999 9 18.998 10 20.18
Na Mg noble gases AI Si P S CI Ar
Sodium Magnesium Aluminum Silicon Phosphorus Sulfur Chlorine Argon
19 39.10 20 40.08 21 44.96 22 47.90 23 50.94 24 51.996 25 54.94 26 55.85 27 58.93 28 58.70 29 63.55 30 65.37 31 69.72 32 72.59 33 74.92 34 78.96 35 79.90 36 83.80
K Ca Sc Ti
Potassium Calcium Scandium Titanium
V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
85.47 38 87.62 39 88.91 40 91.22 41 92.91 42 95.94 43 (98) 44 101.07 45 102.91 46 106.42 47 107.87 48 112.41 49 114.82 50 118.69 51 121.75 52 127.60 53 126.90 54 131.30
37
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium
I
Iodine
Xe Xenon
132.91 56 137.33 57 138.91 72 178.49 73 180.95 74 183.84 75 186.21 76 190.23 77 192.22 78 195.08 79 196.97 80 200.59 81 204.37 82 207.19 83 208.98 84 (209) 85 (210) 86 (222)
55
Cs Ba La Hf Ta W Re Os Ir Pt Au Hg TI Pb Bi Po At Rn
Cesium Barium Lanthanum Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Thallium Lead Bismuth Polonium Astatine Radon
Mercury
(223) 88 226.03 89 227.00 104 (261) 105 (262) 106 (266) 107 (264) 108 (269) 109 (268) 110 (271) 111 (272) (277) (284) (289) (288) (292)
87
58 140.12 59 140.91 60 144.24 61 (145) 62 150.40 63 151.96 64 157.25 65 158.93 66 162.50 67 164.93 68 167.26 69 168.93 70 173.04 71 174.97
Lanthanides Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium
90 232.04 91 231.04 92 238.03 93 237.05 94 (244) 95 (243) 96 (247) 97 (247) 98 (251) 99 (252) 100 (257) 101 (258) 102 (259) 103 (262)
Actinides Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr
Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium
Superactinides (122-153)
Glossary
421
422 Glossary
−1
becquerel (Bq): The SI unit of activity of a radionuclide, equal to 1 s or 1
disintegration per second. One Bq = 2.703 × 10−11 curies (Ci).
berm: Man-made geographical feature; a mound of earth. In this context, a berm
is usually an earthen wall that surrounds holding tanks to contain any liquid that
might leak from them.
beta particle (β particle): Term given to an electron (β − particle) or positron (β +
particle) emitted from the nucleus.
bioassay: Analysis of biological material.
“blind” sample: Jargon; a sample of known content, but unknown to the analyst.
branching ratios: The fractional radiation decays that result in two or more prod-
ucts.
Bremsstrahlung radiation: “Braking radiation” (German). Electromagnetic radi-
ation produced when one charged particle (say, an electron) is deflected or deceler-
ated by another charged particle (say, a nucleus). Bremsstrahlung has a continuous
spectrum that increases in intensity with decreasing energy.
carrier: A stable isotope in appreciable amount which, when thoroughly mixed
with a trace of a specified radioelement, will carry the trace with it through a
chemical or physical process. The stable and radioactive isotopes must behave
identically.
chain reaction: A self-sustaining series of nuclear fissions that is propa-
gated through the absorption of neutrons produced as the product of prior
fissions.
Cherenkov radiation: Electromagnetic radiation resulting from charged particles
traveling in a medium faster than light can travel in the same medium. The de-
celeration of the particle produces this radiation, which is sometimes seen as the
characteristic “blue glow” at a nuclear reactor.
clear melt: The transparent form of melted solids.
Cockcroft-Walton accelerator: A high-voltage device inside which charged ca-
pacitors discharge to add electrons to hydrogen atoms to create negatively charged
ions. The charged particles are then driven through an accelerating tube toward a
target.
cold: Jargon; characterized by an absence of radiation or radionuclides.
colligative: A property that depends on the quantity of atoms, not their form.
colorimetry: Spectrophotometric analysis of visible electromagnetic radiation.
combined standard uncertainty: A standard uncertainty calculated by propa-
gation of uncertainty. The combined standard uncertainty of a result y may be
denoted by u c (y).
combustible liquid: Liquid with a flashpoint at or above 100◦ F. Class II com-
bustible liquids have a flashpoint between 100◦ F and 140◦ F and Class III com-
bustible liquids have a flashpoint above 140◦ F.
424 Glossary
Compton scattering: Occurs when gamma rays interact with electrons in a mate-
rial. The photon donates some of its energy to the electron. The electron is ejected
from its atom and the reduction in energy of the photon results in an increase in
its wavelength. The photon with the remaining energy is deflected.
continuum (energy): Range of particle energies, distributed in a continuous fash-
ion from E = 0 to E = maximum value.
conversion electron (CE): The inner orbital electron that receives the excess
energy of a metastable nucleus during the course of an internal conversion. Because
of its relative proximity to the nucleus, the conversion electron usually comes from
the K shell.
coprecipitation: The simultaneous precipitation of a normally soluble component
with a macro-component from the same solution by the formation of mixed crystals,
by adsorption, occlusion, or mechanical entrapment.
coulombic barrier: Charge potential experienced when one electrically charged
projectile approaches another charged object. For example, electrical charge re-
pulsion when an alpha particle approaches the nucleus of an atom.
coverage factor: The factor k by which the combined standard uncertainty of a
result, u c (y), of a measurement is multiplied to obtain the expanded uncertainty,
U.
coverage probability: The approximate probability that the interval y ± U de-
scribed by a measured result, y, and its expanded uncertainty, U , will contain the
true value of the measurand.
critical value: Threshold value to which the result of a measurement is compared
to make a detection decision (also sometimes called “critical level” or “decision
level”). The critical value can be expressed as (for example) the critical gross count,
critical net count, critical net count rate, or the critical activity.
cross section: Expression of the interaction probability for a nuclear reaction per
unit of area, in cm2 per atom. The term describes the area the nucleus presents for
a particular projectile to strike.
cross-talk: Interfering signals among parallel measurements.
crystal lattice energy: The energy holding ions or atoms together in a crystal
structure.
curie (Ci): A former unit of radioactivity, corresponding roughly to the acitivity
of one gram of the radioactive isotope 226 Ra. 1 Ci = 3.7 × 1010 decays per second
= 37 gigabequerels (GBq).
cyclotron: Circular charged particle accelerator invented by E.O. Lawrence in
1929. Pole diameters range in size from a few inches up to 236 inches with energies
from several million electron volts (MeV) up to 700 MeV for protons. Heavier
projectiles can also be accelerated.
DAC: Derived air concentration of a radionuclide that is related to an intake
limit.
Glossary 425
Dalton: Named after English scientist John Dalton, the Dalton is defined as exactly
1/12 of the relative atomic mass assigned to 12 C. It is functionally the same as an
atomic mass unit (amu) and unified atomic mass unit (u).
daughter(s): Term given to the radioactive product of a radioactive parent, also
progeny.
decay: A term describing the emission of energy from a radionuclide during the
course of a nuclear transformation. Typically one decay mode, such as alpha par-
ticle emission, predominates for a particular nuclide. When more than one decay
mode is evident, branching ratios are used to describe the fraction of the decay
that occurs by each pathway.
decay chain: The succession of radioactive products (called daughters or progeny)
to which a radionuclide decays.
decay constant (λ): A constant (in units of reciprocal time) that is characteristic
of a particular radionuclide’s decay.
decay fraction: The fraction of a particular mode of decay when a radionuclide
decays by several modes; also termed branching ratio.
decontamination factor: The ratio of the proportion of contaminant to product
before treatment to the proportion after treatment.
de-excitation: Process in which an nucleus, atom, or molecule releases energy,
with the effect of reaching a less energetic state.
dewar: A metal or glass container designed to hold liquefied gases. The container
was invented by Scottish chemist and physicist Sir James Dewar, who produced
liquid hydrogen.
disintegration rate: Rate describing the number of nuclear transformations per
unit time, in units of bequerels (Bq). Also referred to as the activity of a radionuclide
sample.
doping: The process of adding a foreign material to an existing material in order
to create a composite material that contains either an excess of electrons (an n-type
material) or an excess of positive holes (a p-type material).
dosimetry: Relating to the practice of monitoring the degree of radiation exposure
of individuals, whether externally or internally.
electrodeposition: Transfer of ions from solution to the surface of an electrode as
part of an electric current.
electromagnetic radiation: Energy which has both an electrical and magnetic
component; when characterized as a wave, identified by its wavelength or fre-
quency; when characterized as an energy bundle, identified by its energy in electron
volts.
electron: The negatively charged subatomic particle that is found in orbital spaces
near the nucleus of the atom.
electron capture (EC): A decay mode characterized by the capture of an atomic
electron from, most probably, an inner electron shell by the nucleus. Electron
426 Glossary
capture parallels positron emission. If the energy difference between the parent
and the daughter is less than 1.022 MeV, electron capture is the sole decay mode.
electron multiplier: A device with multiple dynodes in series to increase an
electron pulse for measurement. Electrons are accelerated as they move toward
each dynode so that multiple electrons are freed at the dynode per arriving
electron.
electron rest energy: Mass of an electron at rest expressed in units of energy:
0.511 MeV.
electron volt (eV): The unit of energy commonly used in quantifying nuclear
processes, the electron volt is defined as the work done by an electron when it falls
through a potential difference of 1 V stated another way, the quantity of energy
that must be added to a standard electron (whose charge equals 1.602 × 10−19 C)
to accelerate it through 1 V; potential difference. Mathematically, 1 eV = (1.602 ×
10−19 C)(1 V) = 1.602 × 10−19 J.
elutriant: A solution that removes a previously retained substance from a sorbent,
such as an ion-exchange column.
Emax : The maximum beta-particle energy for a beta-particle group.
excitation: Term used to describe the process in which an atom, molecule, or
nucleus is elevated to an excited state.
exoergic: Releasing energy.
expanded uncertainty: An uncertainty, U , chosen so that the interval y ± U about
the measured result y is believed to have a high, numerically defined, probability
of containing the true value of the measurand.
extractant: A reagent that implements the transfer of a substance from one solvent
to another.
extraction: Transfer of a substance from one solvent to another.
face velocity: The velocity at which air flows from the laboratory into the interior
of the hood, measured at the face of a hood. This velocity must be high enough
to prevent backspill of inhalants from the hood to the laboratory atmosphere, but
not so high that it interferes with the stability of analytical reagent containers and
equipment.
Fano factor: The observed variance relative to the calculated Poisson distribution
variance, as observed in the peak width in spectral analysis.
Faraday cup: A sensitive device for collecting a stream of electrons or ions and
measuring it in terms of current.
Federal Facility Compliance Act of 1992: Written to amend the Solid Waste Dis-
posal Act, in order to clarify provisions concerning the application of requirements
and sanctions to federal facilities.
fissile material: Any nuclide which is capable of fissioning, either spontaneously
or induced, although the term mostly refers to heavy elements such as uranium or
plutonium. The term “fissionable” may also be used.
Glossary 427
fission: Process in which a nucleus is split into smaller nuclei. Fission may be
spontaneous or induced, symmetric or asymmetric.
flammable liquid: Liquids with a flashpoint below 100◦ F, also known as Class I
liquids. Class IA liquids have a flashpoint below 73◦ F and a boiling point below
100◦ F; Class IB liquids have a flashpoint below 73◦ F and a boiling point at or
above 100◦ F; Class IC liquids have a flashpoint at or above 73◦ F but below 100◦ F.
flashpoint: The minimum temperature at which a liquid gives off vapor within a
test vessel in sufficient concentration to form an ignitable mixture with air near
the surface of the liquid.
flux (as a rate): The rate of transfer of particles or energy across a given barrier.
flux (chemical): A substance added to the analyte, with the goal of facilitating its
dissolution.
footprint: The amount of floor space that an item requires.
Frisch grid: An ionization chamber used for alpha-particle spectroscopy. The
size of an electronic pulse from the chamber is proportional to the energy of
the alpha particle that produced the pulse. Because the energies of alpha particles
are characteristic of the radionuclides that emit them, the Frisch grid chamber is
able to distinguish between different radionuclides by measuring pulse height.
full width at half maximum (FWHM): Width of a peak at 1/2 of its maxi-
mum peak height. The FWHM measurement is a means of expressing the energy
resolution of a spectrometer; also full peak at half maximum (FPHM).
functional group: A specific group of atoms within a molecule that predictably
engages in characteristic chemical reactions.
fusion (nuclear): The process of two or more nuclei joining together to form a
heavier nucleus.
fusion (chemical) : The process of liquifying solid materials by heating.
gamma ray (γ ray): Electromagnetic radiation emitted as a result of a nuclear
transformation.
gas-filled detector: A chamber filled with gas that contains a cathode and an
anode connected by electric circuit to a recording device to measure count rate or
radiation exposure.
Gaussian: Normally distributed or shaped like the familiar “bell curve.”
Gaussian distribution: A normal distribution of a population that is described by
approximations of the Poisson distribution.
Gaussian peak: A spectral peak whose shape is Gaussian.
Geiger–Mueller (G-M) counter: A gas-filled ionization detector operating in the
Geiger–Mueller applied voltage region.
getter: Jargon; term for reactive metals used to scavenge for impurities.
grab sample: A sample collected by hand.
gravimetric: Measurement by weight.
428 Glossary
ground: A large conducting body such as the earth that is used as a common return
for an electric circuit.
half-life (t1/2 ): The time required for any given amount of a radionuclide to decay
to one-half its value. t1/2 = ln 2/decay constant.
HAZMAT: Term colloquially used to refer to hazardous materials and to functions
of agents of the Office of Hazardous Materials Safety. The stated mission of the
OHM is to “promulgate a national safety program that will minimize the risks to
life and property inherent in commercial transportation of hazardous materials.”
See website at http://hazmat.dot.gov (01/15/2006).
Henry’s Law: At contant temperature, the amount of a given gas dissolved in a
given liquid is directly proportional to the partial pressure of the gas when it is in
equilibrium with the liquid.
homogeneous: Uniform throughout.
homolog: In inorganic chemistry, an element in a similar position, such as a col-
umn, in the periodic table.
hoods: Devices designed to protect the analytical chemist from excess exposure
to inhalants by continuously removing laboratory air to the outside, after the air
has passed through a filtering system.
hopcalite: A catalyst for oxidizing hydrogen in hydrocarbons to water vapor.
hot particle: Jargon for an intensely radioactive particle.
hot swap: One of the features of a RAID system, whereby the drives are connected
to the controller. A “hot swappable” RAID drive is one that can be changed out if
it fails, without shutting down the system.
hot: Jargon; characterized by the presence of radiation or radionuclides.
hygroscopic: A characteristic of a material, wherein it absorbs water.
hyperpure: Extremely pure; describes purified germanium for a detector.
ICP/ MS: Inductively coupled plasma/mass spectrometer.
immiscible: Not able to be mixed together.
induced fission: The process in which neutrons are “fired” at a fissionable source
material to cause it to split.
ingrowth: Describes the accumulation of product atoms from radioactive parent
atoms as they decay.
input estimate: In a particular measurement, a measured or imported value of an
input quantity.
input quantity: In a mathematical model of measurement, any of the quantities
whose values are measured or imported and used to calculate the value of the
output quantity Y.
interlaboratory testing: Program in which a number of laboratories participate
to demonstrate competence.
Glossary 429
photoelectric effect: The emission of a free electron from an atom that interacts
with and totally absorbs a gamma ray.
photopeak: In gamma-ray spectral analysis, the photoelectric interaction peak,
but more commonly defined as the full-energy peak.
planchet: A metal disk on which a sample is deposited or fixed in preparation for
counting.
positron: Positively charged electron emitted in a nuclear transformation or pro-
cess.
precipitate: The solid product of a chemical separation, left over after the super-
natant liquid from which it came is removed.
principal quantum number (n): The energy of an electron in an atom is primarily
dependent on this quantum number, as is the size of the atomic orbital in which it
is contained. The greater the value of n for an electron, the higher the energy of
that electron. The principal quantum number may have any positive integer value.
private key: An algorithm that allows a group or individual to enter a key agree-
ment, wherein all communications will require that key before proceeding.
process stream: A part of an industrial process that transfers materials.
propagation of uncertainty: Mathematical operation of combining standard un-
certainties (and estimated covariances) of input estimates to obtain the combined
standard uncertainty of the output estimate.
proportional counter: A gas-filled ionization detector that operates in the propor-
tional region of the applied voltage, in which the output pulse energy is proportional
to the deposited energy.
proton: Fundamental subatomic particle of the atom that has a charge of +1 with
a mass of approximately one dalton.
public key: A form of encryption that allows users to communicate securely
without prior possession of a secret key.
p-type: Describes a material whose structure contains positive holes through which
charge can move.
quality assurance (QA): A program that systematically monitors and evaluates
laboratory operation to ensure that sample analysis meets standards of quality.
quality assurance plan (QAP): A planning tool that allows for the comprehensive
management of laboratory operation to ensure that standards of quality are met.
quality assurance project plan (QAPP): A planning tool that allows for the
comprehensive management of a data collection activity (DCA), from sample
collection through data reporting, to ensure that standards of quality are met.
quality control (QC): System for ensuring maintenance of proper standards in
the laboratory by use of standard and comparison samples.
quenching: In a scintillation detector, to extinguish light emanation from the
sample, to a degree that depends on the quenching agent, thus reducing the
Glossary 433
sector field analyzer: A spectrometer that separates ions by the curvature of their
flight path, for which the radius is inversely proportional to the magnetic field and
directly proportional to the square root of the mass to charge ratio.
secular equilibrium: A case where the activity (disintegration rate) of the parent
and progeny becomes equal after a time. For this type of equilibrium, the half-life
of the parent is at least 10 times greater than that of the daughter.
secure shell (ssh): A set of standards and accompanying network protocol that
allows for the establishment of a secure channel between a local computer and one
located remotely.
sequential analysis: Serial chemical analysis of several substances in the same
solution.
shell (as in K, L, etc): The letter designation of the principal quantum number, n.
Orbitals of the same n value are referred to as belonging to the same shell. Several
subshells may be contained in a given shell; subshell designation depends on other
quantum numbers
slurry: A watery mixture of insoluble material.
solubility product (Ksp ): Mathematically, the product of the thermodynamic ac-
tivities of the ionic components of a substance precipitated from a solution at
equilibrium.
source: In the context of radionuclide analysis, a sample prepared for radionuclide
measurement, i.e., the source of the radionuclides that are being measured.
specific activity: For a specified isotope, or mixture of isotopes, the activity (dis-
integration rate) of a material divided by the mass of the element.
spallation: To break off. In nuclear physics, the term refers to the emission of a
large number of neutrons by a heavy radionuclide after it has been bombarded by
charged particles.
spike: See tracer.
spiking: Jargon; addition to a sample of a tracer for the radionuclide of interest.
spontaneous fission: Natural decay process found in elements with Z greater than
or equal to 90 (although it is unclear whether Th (Z = 90) can fission sponta-
neously).
stable nuclide: A nuclide that is not radioactive.
standard method: A procedure that has been selected by testing and agreement
among skilled professionals.
standard reference material (SRM) or standard radioactive material (SRM):
Material, often in liquid form, that is certified to have a known concentration or
decay rate. SRM is used in instrument calibration and as primary standards for
laboratory measurements.
standard uncertainty: Uncertainty of the result of a measurement expressed as
a standard deviation (sometimes called a “one-sigma” uncertainty). The standard
uncertainty of a measured value x may be denoted by u(x).
Glossary 435
stakeholder: In the context presented herein, a person or group that has an interest
in the results produced by a laboratory or facility.
state-of-health: During the lifetime of an instrument, its performance or “health”
tends to deteriorate. The degree of this decline can be inferred from measurements,
specified for a particular system, that define its well being. Parameters may include
battery power, calibration outputs and various hardware checks.
stopping power: The differential energy loss per travel distance, (dE/dx) of a
particle.
STP (standard temperature and pressure): Zero degrees Celsius (273◦ Abso-
lute) and one atmosphere pressure.
subatomic particle: Any fundamental particle that is smaller than the atom.
symmetric fission: Fission process in which product yields decrease on both sides
of an intermediate mass that is produced with maximum yield.
Szilard–Chalmers effect: The rupture of the chemical bond between an atom
and the molecule of which the atom was a component, as a result of a nuclear
transformation of that atom.
thermoluminescent detector (TLD): Thermoluminescent material in powder
form or small chips is placed in a badge for radiation monitoring. After a specified
period of time, the badge is collected and “read” by heating it to emit light that is
proportional to the energy deposition of the detected radiation.
thin target: Term applied to materials that are irradiated but do not significantly
attenuate the projectile beam or flux.
traceability: Documentation for a standard that allows it to be historically linked
to the laboratory of its formulation and testing, notably at NIST.
tracer: A substance that is added to a material to observe the behavior of one of
its components.
transactinides: All elements with atomic number greater than lawrencium (103),
the last of the actinide series.
transient equilibrium: A case where the ratio of the daughter activity to the parent
activity becomes constant with a value that exceeds 1.0 after several half-lives. The
half life of the parent is longer than that of the daughter.
transport velocity: The velocity at which air is flushed from the hood to the outside
air. This velocity depends on fan speed and ductwork dimensions and is important
to assure that airborne particles are transported from the hood and not deposited
in the ductwork.
Type A evaluation: “Method of evaluation of uncertainty by the statistical analysis
of series of observations” (ISO 1995).
Type B evaluation: “Method of evaluation of uncertainty by means other than the
statistical analysis of series of observations” (ISO 1995).
uncertainty propagation: See propagation of uncertainty.
436 Glossary
ACGIH 2001. Air Sampling Instruments for Evaluation of Atmospheric Contaminants, 9th
ed. Cincinnati OH: American Conference of Government Industrial Hygienists, Inc.
ACS 1990. American Chemical Society. Informing Workers of Chemical Hazards: The
OSHA Hazard Communication Standard. 2nd ed. Washington, DC: ACS.
ACS 1998. American Chemical Society. Living with the Laboratory Standard: A Guide for
Chemical Hygiene Officers. Washington, DC: ACS.
ACS 2000. American Chemical Society. Safety Audit/Inspection Manual. Washington, DC:
ACS.
ACS 2003. American Chemical Society. Safety in Academic Chemistry Laboratories. 7th
ed. Washington, DC: ACS.
Adloff, J.-P. and Guillaumont, R. 1993. Fundamentals of Radiochemistry. pp. 327–352.
Boca Raton, FL: CRC Press.
Adriaens, A. G., Fassett, J. D., Kelly, W. R., Simons, D. S., and Adams, F. C. 1992. De-
termination of Uranium and Thorium Concentrations in Soils—Comparison of Isotope-
Dilution Secondary Ion Mass-Spectrometry and Isotope-Dilution Thermal Ionization
Mass-Spectrometry. Anal Chem 64(23), 2945–2950.
AIHA 1995. American Industrial Hygiene Association. Laboratory Chemical Hygiene: An
AIHA Protocol Guide. Fairfax, VA: AIHA Press.
Alaimo, R. J. and Fivizzani, K. P. November/December 1996. Qualifications and training
of chemical hygiene officers. Chemical Health and Safety 3(6), 10–13.
Aldstadt, J. H., Kuo, J. M., Smith, L. L., and Erickson, M. D. 1996. Determination of
uranium by flow injection inductively coupled plasma mass spectrometry. Anal Chim
Acta 319(1–2), 135–143.
Alonso, J. I. G., Babelot, J. F., Glatz, J. P., Cromboom, O., and Koch, L. 1993. Applications
of a glove-box ICP-MS for the analysis of nuclear- materials. Radioc Acta 62(1–2),
71–79.
Alonso, J. I. G., Sena, F., Arbore, P., Betti, M., and Koch, L. 1995. Determination of fission-
prroducts and actinides in spent nuclear-fuels by isotope-dilution ion chromatography
inductively-coupled plasma-sass spectrometry. J Anal Atom Spectromy 10(5), 381–393.
Altshuler, B. and Pasternack, B. 1963. Statistical measures of the lower limit of detection
of a radioactivity counter. Health Phys 9, 293–298.
Amphlett, C. B. 1964. Inorganic Ion Exchangers. New York, NY: Elsevier.
Anders, E. 1960. The Radiochemistry of Technetium. National Research Council Report
NAS-NS 3021. Washington, DC: National Research Council.
437
438 References
Beasley, T. M., Kelley, J. M., Maiti, T. C., and Bond, L. A. 1998b. Np-237/Pu-239 atom
ratios in integrated global fallout: a reassessment of the production of Np-237. J Environ
Radioactiv 38(2), 133–146.
Beasley, T. M., Kelley, J. M., Orlandini, K. A., Bond, L. A., Aarkrog, A., Trapeznikov,
A. P., and Pozolotina,V. N. 1998c. Isotopic Pu, U, and Np signatures in soils from
Semipalatinsk-21, Kazakh Republic and the Southern Urals, Russia. J Environ Radioactiv
39(2), 215–230.
Becker, J. S. 2002. State-of-the-art and progress in precise and accurate isotope ratio mea-
surements by ICP-MS and LA-ICP-MS—plenary lecture. J Anal Atom Spectrom 17(9),
1172–1185.
Becker, J. S. 2003. Mass spectrometry of long-lived radionuclides. Spectrochim Acta B
58(10), 1757–1784.
Becker, J. S. and Dietze, H. J. 1997. Double-focusing sector field inductively coupled
plasma mass spectrometry for highly sensitive multi-element and isotopic analysis—
Invited lecture. J Anal Atom Spectrom 12(9), 881–889.
Becker, J. S. and Dietze, H. J. 1998. Inorganic trace analysis by mass spectrometry. Spec-
trochim Acta B 53(11), 1475–1506.
Becker, J. S. and Dietze, H. J. 2000. Precise and accurate isotope ratio measurements by
ICP-MS. Fresen J Anal Chem 368(1), 23–30.
Becker, J. S., Pickhardt, C., and Dietze, H. J. 2000. Laser ablation inductively coupled
plasma mass spectrometry for the trace, ultratrace and isotope analysis of long-lived
radionuclides in solid samples. Int J Mass Spectrom 202(1–3), 283–297.
Becker, J. S., Burow, M., Boulyga, S. F., Pickhardt, C., Hille, R., and Ostapczuk, P. 2002.
ICP-MS determination of uranium and thorium concentrations and U-235/U-238 isotope
ratios at trace and ultratrace levels in urine. Atom Spectrosc 23(6), 177–182.
Becker, J. S., Burow, M., Zoriy, M. V., Pickhardt, C., Ostapczuk, P., and Hille, R. 2004a.
Determination of uranium and thorium at trace and ultratrace levels in urine by laser
ablation ICP-MS. Atom Spectrosc 25(5), 197–202.
Becker, J. S., Zoriy, M., Halicz, L., Teplyakov, N., Muller, C., Segal, I., Pickhardt, C.,
and Platzner, I. T. 2004b. Environmental monitoring of plutonium at ultratrace level
in natural water (Sea of Galilee-Israel) by ICP-SFMS and MC-ICP-MS. J Anal Atom
Spectrom 19(9), 1257–1261.
Bellis, D., Ma, R., Bramall, N., McLeod, C. W., Chapman, N., and Satake, K. 2001. Airborne
uranium contamination—as revealed through elemental and isotopic analysis of tree bark.
Environ Pollut 114(3), 383–387.
Belloni, J., Haissinsky, M., and Salama, H. N. 1959. On the adsorption of some fission
products on various surfaces. J Phys Chem 63, 881–887.
Bemis, C. E., Jr., Dittner, P. F., Silva, R. J., Hahn, R. L., Tarrant, J. R., Hunt, L. D., and
Hensley, D. C. 1977. Production, L x-ray identification and decay of the nuclide (260)105.
Phys Rev C 16, 1146–1157.
Bemis, C. E., Jr., Silva, R. J., Hensley, D. C., Keller, O. L., Jr., Tarrant, J. R., Hunt, L. D.,
Dittner, P. F., Hahn, R. L., and Goodman, C. D. 1973. X-Ray identification of element 104
Phys Rev Lett 31, 647–650.
Berkovits, D., Feldstein, H., Ghelberg, S., Hershkowitz, A., Navon, E., and Paul, M. 2000.
U-236 in uranium minerals and standards.Nuc Instrum Meth B 172, 372–376.
Betti, M. 1996. Use of a direct current glow discharge mass spectrometer for the chemical
characterization of samples of nuclear concern. J Anal Atom Spectrom 11(9), 855–860.
Bings, N. H., Bogaerts, A., and Broekaert, J. A. C. 2002. Atomic spectroscopy. Anal Chem
74(12), 2691–2711.
440 References
Blanchard, R. L., Kahn, B., and Birkhoff, R. D. 1960. The preparation of thin, uniform,
radioactive sources by surface adsorption and electrodeposition. Health Phys 2, 246–255.
Bland, C. J. 1984. Tables of the geometrical factor for various source-detector configura-
tions. Nuc Instrum Methods 223, 2–3, 602–606.
Blue, T. E. and Jarzemba, M. S. 1992. A model for radon gas adsorption on charcoal for
open-faced canisters in an active environment. Health Phys 63(2), 226–232.
Bock, R. 1979. A Handbook of Decomposition Methods in Analytical Chemistry. New York:
John Wiley.
Boni, A. L. 1966. Rapid ion exchange analysis of radiocesium in milk, urine, seawater and
environmental samples. Anal Chem 38, 89–92.
Boulyga, S. F. and Becker, J. S. 2001. Determination of uranium isotopic composition and
U-236 content of soil samples and hot particles using inductively coupled plasma mass
spectrometry. Fresen J Anal Chem 370(5), 612–617.
Boulyga, S. F. and Becker, J. S. 2002. Isotopic analysis of uranium and plutonium using
ICP-MS and estimation of burn-up of spent uranium in contaminated environmental
samples. J Anal Atom Spectrom 17(9), 1143–1147.
Boulyga, S. F., Becker, J. S., Matusevitch, J. L., and Dietze, H. J. 2000. Isotope ratio
measurements of spent reactor uranium in environmental samples by using inductively
coupled plasma mass spectrometry. Int J Mass Spectrom 203(1–3), 143–154.
Boulyga, S. F., Desideri, D., Meli, M. A., Testa, C., and Becker, J. S. 2003. Plutonium and
americium determination in mosses by laser ablation ICP-MS combined with isotope
dilution technique. Int J Mass Spectrom 226(3), 329–339.
Bowyer, T. W., Abel, K. H., Hubbard, C. W., Panisko, M. E., Reeder, P. L., Thompson,
R. C., and Warner, R. A. 1999. Field testing of collection and measurement of radioxenon
for the comprehensive test ban treaty. J Radioanal Nuc Ch 240, 109–122.
B riesmeister, J. F. 1990. MCNP-A general Monte Carlo code for neutron and photon
transport. Version 4.2. Technical Report LA-7396-M, Los Alamos National Laboratory.
Brown, T. A., Marchetti, A. A., Martinelli, R. E., Cox, C. C., Knezovich, J. P., and Hamilton,
T. F. 2004. Actinide measurements by accelerator mass spectrometry at Lawrence Liv-
ermore National Laboratory. Nuc Instrum Meth B 223–224, 788–795.
Brown, E., Skougstad, M. W., and Fishman, M. J. 1970. Methods for collection and analysis
of water samples for dissolved minerals and gases in Techniques of Water-Resources
Investigations, Book 5. Chapter A1. Reston, VA: US Geological Survey.
Brown, R. M., Workman, W. J., and Kotzer, T. G. 1993. Tritium program at chalk river
laboratories. Tritium monitoring in the environment in Transactions of the Ameri-
can Nuclear Society. Paper 1, pp. 20. Vol. 69. LaGrange Park, IL: American Nuclear
Society.
Budnitz, R. J., Nero, A. V., Murphy, D. J., and Graven, R. 1983. Instrumentation for Envi-
ronmental Monitoring. Vol. 1, Radiation. 2nd ed. New York, NY: John Wiley.
Burnett, W. C., Schultz, M. K., Corbett, D. R., Horwitz, E. P., Chiarizia, R., Dietz, M.,
Thakkar, A., and Fern, M. 1997. Preconcentration of actinide elements from soils and
large volume water samples using extraction chromatography. J Radioanal Nucl Chem
226, 121–127.
Byrnes, M. E. 2001. Sampling and Surveying Radiological Environments. Boca Raton, FL:
Lewis Publishers, CRC Press.
Carlton, W. H., Bauer, L. R., Evans, A. G., Geary, L. A., Murphy, C. E., Jr., Pinder, J. E., and
Strom, R. N. 1992. Cesium in the Savannah River Site Environment. US Department of
Energy Report WSRC-RP-92-250. Aiken, SC: Westinghouse Savannah River Company.
Cember, H. 1996. Introduction to Health Physics. New York, NY: McGraw-Hill.
References 441
Coryell, C. D. and Sugarman, N. 1951. Radiochemical studies: the fission products. National
Nuclear Energy Series. 3 books, Vol. 9. New York, NY: McGraw-Hill.
Costa-Fernandez, J. M., Bings, N. H., Leach, A. M., and Hieftje, G. M. 2000. Rapid simul-
taneous multielemental speciation by capillary electrophoresis coupled to inductively
coupled plasma time-of-flight mass spectrometry. J Anal Atom Spectrom 15(9), 1063–
1067.
Cristy, S. S. 2000. Secondary ion mass spectrometry in Inorganic Mass Spectrometry.
Vol. 23.. Barshick, C. M., Duckworth, D. C., and Smith, D. H., eds. pp. 159–221. New
York, NY: Marcel Dekker.
Curie, P. and Sklodowska Curie, M. 1898. Sur une substance Nouvelle Radio-active, con-
tenue dans la Pechblende. Compt. Rend. 127, 175.
Currie, L. A. 1968. Limits for qualitative detection and quantitative determination—
application to radiochemistry. Anal Chem 40, 586–593.
Curtiss, L. F. and Davis, F. J. 1943. A counting method for the determination of small
amounts of radium and radon. J Res Nat Bur Stand. 31, 181–195.
Dai, M. H., Buesseler, K. O., Kelley, J. M., Andrews, J. E., Pike, S., and Wacker, J. F. 2001.
Size-fractionated plutonium isotopes in a coastal environment. J Environ Radioactiv
53(1), 9–25.
Daly, N. R. 1960. Scintillation type mass spectrometer ion detector. Rev Sci Instrum 31(3),
264–267.
Danesi, P. R., Bleise, A., Burkart, W., Cabianca, T., Campbell, M. J., Makarewicz, M.,
Moreno, J., Tuniz, C., and Hotchkis, M. 2003. Isotopic composition and origin of uranium
and plutonium in selected soil samples collected in Kosovo. J Environ Radioactiv 64(2–
3), 121–131.
Daubert v. Merrell Dow Pharmaceuticals 1993. 509 U.S. 579.
de Laeter, J. R. 1996. The role of off-line mass spectrometry in nuclear fission. Mass
Spectrom Rev 15(4), 261–281.
de Laeter, J. R. 2001. Applications of Inorganic Mass Spectrometry. New York, NY: John
Wiley.
de las Heras, L. A., Bocci, F., Betti, M., and Actis-Dato, L. O. 2000. Comparison between the
use of direct current glow discharge mass spectrometry and inductively coupled plasma
quadrupole mass spectrometry for the analysis of trace elements in nuclear samples.
Fresen J Anal Chem 368(1), 95–102.
De, A. K., Khopkar, S. M., and Chalmers, R. A. 1970. Solvent Extraction of Metals. London:
Van Nostrand Reinhold Co.
Dean, J. A. 1992. Lange’s Handbook of Chemistry. 14th ed. Table 11.22, pp. 11–61. New
York: McGraw-Hill.
Dean, J. A. 1999. Lange’s Handbook of Chemistry. 15th ed. New York, NY: McGraw-Hill.
Delanghe, D., Bard, E., and Hamelin, B. 2002. New TIMS constraints on the uranium-238
and uranium-234 in seawaters from the main ocean basins and the Mediterranean Sea.
Mar Chem 80(1), 79–93.
Delmore, J. E., Appelhans, A. D., and Peterson, E. S. 1995. A rare-earth-oxide matrix for
emitting perrhenate anions. Int J Mass Spectrom Ion Processes 146, 15–20.
DeVoe, J. R. 1962. Application of Distillation Techniques To Radiochemical Separations.
National Research Council Report NAS-NS-3108. Washington, DC: NRC.
Deyl, Z., ed. 1979. Electrophoresis: a survey of techniques and applications, Part A. Tech-
niques. J Chromatogr Libr Vol. 18. New York, NY: Elsevier.
DHHS 1982. Department of Health and Human Services. DHHS evaluation of results
of environmental chemical testing by EPA in the vicinity of Love Canal—Implications
References 443
Erdmann, N., Nunnemann, M., Eberhardt, K., Herrmann, G., Huber, G., Kohler, S., Kratz,
J. V., Passler, G., Peterson, J. R., Trautmann, N., and Waldek, A. 1998. Determination
of the first ionization potential of nine actinide elements by resonance ionization mass
spectroscopy (RIMS). J Alloys Compd 271, 837–840.
ETP 2005. Online at http://www.etpsci.com/htm/etp/tech info/tech articles/active film.asp.
Viewed on November 15, 2005.
Evans, R. D. 1955. The Atomic Nucleus. Pp. 470–510, 625 New York, NY: McGraw-Hill.
Faires, R. A. and Boswell, G. G. J. 1981. Radioisotope Laboratory Techniques. 4th ed.
Oxford, London: Butterworth-Heinemann.
Fassel, V. A. 1971. Spectroscopic properties and analytical applications of induction-
coupled plasmas. Appl Spectrosc 25(1), 148–&.
Fassel, V. A. and Kniseley, R. N. 1974 Inductively coupled plasmas. Anal Chem 46(13),
1155–&.
Faure, G. 1986. Principles of Isotope Geology. P. 399. New York, NY: John Wiley.
Fearey, B. L., Tissue, B. M., Olivares, J. A., Loge, G. W., Murrell, M. T., and Miller,
C. M. 1992. High-precision thorium RIMS for geochemistry. Inst Phys Conf Ser (128),
209–212.
Fifield, L. K. 1999. Accelerator mass spectrometry and its applications. Rep Prog Phys
62(8), 1223–1274.
Fifield, L. K. 2000. Advances in accelerator mass spectrometry. Nucl Instrum Meth B 172,
134–143.
Fifield, L. K., Carling, R. S., Cresswell, R. G., Hausladen, P. A., di Tada M. L., and Day,
J. P. 2000. Accelerator mass spectrometry of Tc-99. Nucl Instrum Meth B 168(3), 427–
436.
Finney, G. D. and Evans, R. 1935. The radioactivity of solids determined by alpha-ray
counting. Phys Rev 48, 503.
Finston, H. L. and Kinsley, M. T. 1961. The Radiochemistry of Cesium. National Research
Council Report NAS-NS 3035. Washington, DC: National Research Council.
Firestone, R. B. 1990. Table of Isotopes. 8th ed. New York, NY: John Wiley.
Firestone, R. B. 1996. Table of Isotopes. 9th ed. New York, NY: John Wiley.
Friedlander, G. and Herrman, G. 2003. History of nuclear and radiochemistry in Handbook
of Nuclear Chemistry. Vol. I. Chapter 1. Vertes, A., Nagy, S., and Klencsar, Z., eds.
Dordrecht: Kluwer Academic.
Friedlander, G., Kennedy, J. W., Macias, E. S., and Miller, J. M. 1981. Nuclear and Radio-
chemistry. 3rd ed. New York, NY: John Wiley.
Fritz, J. S. and Umbreit, G. R. 1958. Ion exchange separation of metal ions. Anal Chim Acta
19, 509–516.
FRMAC 2002. Federal Radiological Monitoring and Assessment Center Report
DOE/NV/11718—181. Monitoring and Sampling Manual. Vol. 1, Rev. 1. Washington
DC: FRMAC.
Gäggeler, H. W. 1994. On-line gas chemistry experiments with transactinide elements. J
Radioanal Nucl Chem Articles 183, 261.
Gäggeler, H. W., Jost, D. T., Kovacs, J., Scherer, U. W., Weber, A., Vermeulen, D., Türler,
A., Gregorich, K. E., Henderson, R. A., Czerwinski, K. R., Kadkhodayan, B., Lee,
D. M., Nurmia, M., Hoffman, D. C., Kratz, J. V., Gober, M. K., Zimmermann, H. P.,
Schädel, M., Brüchle, W., Schimpf, E., and Zvara, I. 1992. Gas phase chromatography
experiments with bromides of tantalum and element-105. Radiochim Acta 57,
93–100.
446 References
Garcia, R. and Kahn, B. 2001. Total dissolution of environmental and biological samples for
radiometric analysis by closed-vessel microwave digestion. J Radiological Nucl Chem
250, 85–91.
Gärtner, M., Boettger, M., Eichler, B., Gäggeler, H. W., Grantz, M., Hübener, S., Jost,
D. T., Piguet, D., Dressler, R., Türler, A., and Yakushev, A. B. 1997. On-line gas chro-
matography of Mo, W and U (oxy) chlorides. Radiochim Acta 78, 59–68.
Gastel, M., Becker, J. S., Kuppers, G., and Dietze, H. J. 1997. Determination of long-
lived radionuclides in concrete matrix by laser ablation inductively coupled plasma mass
spectrometry. Spectrochim Acta B 52(14), 2051–2059.
Gerlich, D. 1992. Inhomogeneous Rf-fields—a versatile tool for the study of processes with
slow ions. Adv Chem Phys 82, 1–176.
Ghiorso, A., Lee, D., Somerville, L. P., Loveland, W., Nitschke, J. M., Ghiorso, W., Seaborg,
G. T., Wilmarth, P., Leres, R., Wydler, A., Nurmia, M., Gregorich, K., Czerwinski, K.,
Gaylord, R., Hamilton, T., Hannink, N. J., Hoffman, D. C., Jarzynski, C., Kacher, C.,
Kadkhodayan, B., Kreek, S., Lane, M., Lyon, A., McMahan, M. A., Neu, M., Sikkeland,
T., Swiatecki, W. J., Türler, A., Walton, J. T., and Yashita, S. 1995a. Evidence for the
synthesis of element 267 110 produced by the 59 Co + 209 Bi reaction. Nucl Phys 583,
861–866.
Ghiorso, A., Lee, D., Somerville, L. P., Loveland, W., Nitschke, J. M., Ghiorso, W., Seaborg,
G. T., Wilmarth, P., Leres, R., Wydler, A., Nurmia, M., Gregorich, K., Czerwinski, K.,
Gaylord, R., Hamilton, T., Hannink, N. J., Hoffman, D. C., Jarzynski, C., Kacher, C.,
Kadkhodayan, B., Kreek, S., Lane, M., Lyon, A., McMahan, M. A., Neu, M., Sikkeland,
T., Swiatecki, W. J., Türler, A., Walton, J. T., and Yashita, S. 1995b. Evidence for the
possible synthesis of element 267 110 produced by the 59 Co + 209 Bi reaction. Phys Rev C
51, R2293–R2297.
Ghiorso, A., Nitschke, J. M., Alonso, J. R., Alonso, C. T., Nurmia, M., Seaborg, G. T.,
Hulet, E. K., and Lougheed, R. W. 1974. Element 106. Phys Rev Lett 33, 1490–1493.
Gibson, J. K. 1998. Laser ablation mass spectrometry of actinide dioxides: ThO2 , UO2 ,
NpO2 , PuO2 and AmO2 . Radiochim Acta 81(2), 83–91.
Gill, C. G., Daigle, B., and Blades, M. W. 1991. Atomic mass-spectrometry of solid samples
using laser ablation-ion trap mass-spectrometry (LAITMS). Spectrochim Acta B 46(8),
1227–1235.
Gilmore, G. and Hemingway, J. D. 1995. True coincidence summing in Practical Gamma-
Ray Spectrometry. Chapter 7. West Sussex, England: John Wiley.
Gindler, J. E. 1962. The Radiochemistry of Uranium. National Research Council Report
NAS-NS 3050. Washington, DC: National Research Council.
Gonzalez, J., Mao, X. L., Roy, J., Mao, S. S., and Russo, R. E. 2002. Comparison of 193,
213 and 266 nm laser ablation ICP-MS. J Anal Atom Spectrom 17(9), 1108–1113.
Gordon, L., Salutsky, M. L., and Willard, H. H. 1959. Precipitation from Homogeneous
Solution. New York, NY: John Wiley.
Granstrom, M. L. and Kahn, B. 1955. The adsorption of cesium in low concentration by
paper. J Phys Chem 59, 408–410.
Grate, J. W. and Egorov, O. 2003. Automated radiochemical separation, analysis, and sens-
ing in Handbook of Radioactivity Analysis. pp. 1129–1164. San Diego, CA: Elsevier.
Gray, A. L. and Date, A. R. 1983. Inductively coupled plasma source-mass spectrometry
using continuum flow ion extraction. Analyst 108(1290), 1033–1050.
Gregorich, K. E., Ginter, T., Loveland, W., Peterson, D., Patin, J. B., Folden, C. M., III,
Hoffman, D. C., Lee, D. M., Nitsche, H., Omtvedt, J. P., Omtvedt, L. A., Stavsetra, L.,
References 447
Sudowe, R., Wilk, P. A., Zielinski, P., and Aleklett, K. 2003. Cross section limits for the
208
Pb(86 Kr,n)293 118 reaction. Eur Phys J A 18, 633–638.
Gregorich, K. E., Henderson, R. A., Lee, D. M., Nurmia, M. J., Chasteler, R. M., Hall,
H. L., Bennett, D. A., Gannett, C. M., Chadwick, R. B., Leyba, J. D., Hoffman, D. C.,
and Herrmann, G. 1988. Aqueous chemistry of element l05. Radiochim Acta 43, 223–231.
Groenewold, G. S., Delmore, J. E., Olson, J. E., Appelhans, A. D., Ingram, J. C., and Dahl,
D. A. 1997. Secondary ion mass spectrometry of sodium nitrate: Comparison of ReO4-
and Cs+ primary ions. Int J Mass Spectrom Ion Processes 163(3), 185–195.
Gruning, C., Huber, G., Klopp, P., Kratz, J. V., Kunz, P., Passler, G., Trautmann, N., Waldek,
A., and Wendt, K. 2004. Resonance ionization mass spectrometry for ultratrace analysis
of plutonium with a new solid state laser system. Int J Mass Spectrom 235(2), 171–178.
Guillaumont, R., Adloff, J.-P., and Peneloux, A. 1989. Kinetic and thermodynamic aspect
of tracer-scale and single atom chemistry. Radiochim Acta 46, 169–176.
Guillaumont, R., Adloff, J.-P., Peneloux, A., and Delamoye, P. 1991. Sub-tracer scale be-
haviour of radionuclides. Application to actinide chemistry. Radiochim Acta 54, 1–15.
Gunnink, R. and Niday, J. B. 1972. Description of the GAMANAL program, UCRL-51061
in Computerized Quantitative Analysis by Gamma-ray Spectrometry. Vol. I. Livermore,
CA: Lawrence Livermore Laboratory.
Gunther, D., Horn, I., and Hattendorf, B. 2000. Recent trends and developments in laser
ablation-ICP-mass spectrometry. Fresen J Anal Chem 368(1), 4–14.
GV Instruments 2005. Application Report AN-20. Online at http://www.gvinstruments.
co.uk/index.asp.
Gy, Pierre M. 1992. Sampling of Heterogeneous and Dynamic Material Systems: Theories
of Heterogeneity, Sampling and Homogenizing. Amsterdam, The Netherlands: Elsevier.
Hahn, O. 1936. Applied Radiochemistry. Ithaca, NY: Cornell University Press.
Haissinsky, M. 1964. Nuclear Chemistry and its Applications. Reading, MA: Addison-
Wesley.
Haldimann, M., Baduraux, M., Eastgate, A., Froidevaux, P., O’Donovan, S., Von Gunten,
D., and Zoller, O. 2001. Determining picogram quantities of uranium in urine by iso-
tope dilution inductively coupled plasma mass spectrometry. Comparison with alpha-
spectrometry. J Anal Atom Spectrom 16(12), 1364–1369.
Harris, M. K., Herrington, P. B., Miley, H. S., Ellis, J. E., McKinnon, A. D., and St. Pierre,
D. E. 1999. Data authentication demonstration radionuclide stations in 21st Seismic
Research Symposium: Technologies for Monitoring the Comprehensive Nuclear-Test-
Ban Treaty. pp. 331–337. Los Alamos, NM: Los Alamos National Laboratory.
Heimbuch, A. M., Gee, H. Y., DeHaan, A. Jr., and Leventhal, L. 1965. The assay of alpha-
and beta-emitters by means of scintillating ion-exchange resins in Radioactive Sample
Measurement Techniques in Medicine and Biology. pp. 505–519. Vienna, Austria: Inter-
national Atomic Energy Agency.
Heinrich, C. A., Pettke, T., Halter, W. E., Aigner-Torres, M., Audetat, A., Gunther, D.,
Hattendorf, B., Bleiner, D., Guillong, M., and Horn, I. 2003. Quantitative multi-element
analysis of minerals, fluid and melt inclusions by laser-ablation inductively-coupled-
plasma mass-spectrometry. Geochim Cosmochim Ac 67(18), 3473–3497.
Hellborg, R., Faarinen, M., Kiisk,M., Magnusson, C. E., Persson, P., Skog, G., and
Stenstrom, K. 2003. Accelerator mass spectrometry—an overview. Vacuum 70(2–3),
365–372.
Henry, R., Koller, D., Liezers, M., Farmer, O. T., Barinaga, C., Koppenaal, D. W., and
Wacker, J. 2001. New advances in inductively coupled plasma—mass spectrometry
448 References
(ICP-MS) for routine measurements in the nuclear industry. J Radioanal Nucl Ch 249(1),
103–108.
Hensley, J. W., Long, A. O., and Willard, J. E. 1949. Reactions of ions in aqueous solution
with glass and metal surfaces. Ind Eng Chem 41, 1415–1421.
Herrmann, G., Riegel, J., Rimke, H., Sattelberger, P., Trautmann, N., Urban, F. J., Ames,
F., Otten, E. W., Ruster, W., and Scheerer, F. 1991. Resonance ionization mass-
spectroscopy of uranium. Inst Phys Conf Ser (114), 251–254.
Herzog, G. F. 1994. Applications of accelerator mass-spectrometry in extraterrestrial ma-
terials. Nucl Instrum Meth B 92(1–4), 492–499.
Heumann, K. G. 1992. Isotope-dilution mass-spectrometry. Int J Mass Spectrom Ion Pro-
cesses 118, 575–592.
Heusser, G. 2003. Low-background gamma spectroscopy of natural decay chain activity
at the μBq/kg level. Paper presented at the national meeting of the American Chemical
Society, Division of Nuclear Chemistry and Technology; Paper No. 15. New York. Con-
tact address for Heusser: Max Planck Institut fuer Kernphysik, POB 103 980 D-69029,
Heidelberg, Germany.
Hidaka, H., Holliger, P., and Gauthier-Lafaye, F. 1999. Tc/Ru fractionation in the Oklo and
Bangombe natural fission reactors, Gabon. Chem Geol 155(3–4), 323–333.
Hindman, F. D. 1986. Actinide separations for α spectrometry using neodymium fluoride
coprecipitation. Anal Chem 58, 1238–1241.
Hoffman, D. C. 1994. The heaviest elements. Chem & Eng News 72, 24–34.
Hoffman, D. C. and Lee, D. M. 1999. Chemistry of the heaviest elements—One atom at a
time. J Chem Ed 76, 331–347.
Hoffman, D. C. and Lee, D. M. 2003. Superheavy Elements in Handbook of Nuclear
Chemistry, vol. 2. Chapter 10. pp. 397–416. Attila Vértes, S. N., Zoltán Klencsár, eds.
Norwell, Massachusetts: Kluwer Academic Publishers.
Hoffman, D. C., Hamilton, T. M., and Lane, M. R. 1996. Spontaneous Fission, in Nuclear
Decay Modes, Chapter 10. Poenaru, D. N. ed. Bristol: Institute of Physics Publishing.
Hoffman, D. C., Ghiorso, A., and Seaborg, G. T. 2000. The Transuranium People: The
Inside Story. pp. 258–298, 379–396. London: Imperial College Press.
Hoffman, D. C., Lee, D. M., and Pershina, V. 2004. Transactinide Elements and Future
Elements in: The Chemistry of the Actinide and Transactinide Elements. 3rd ed. vol. III.
Chapter 14. Katz, J. J., Morss, L. R., Edelstein, N., Fuger, J. eds. London: Kluwer
Academic Publishers.
Hofmann, S., Hessberger, F. P., Ackermann, D., Münzenberg, G., Antalic, S., Cagarda, P.,
Kindler, B., Kojouharova, J., Leino, M., Lommel, B., Mann, R., Popeko, A. G., Reshitko,
S., Saro, S., Uusitalo, J., and Yeremin, A. V. 2002. New results on elements 111 and 112.
Eur Phys J A 14, 147–157.
Hofmann, S., Ninov, V., Hessberger, F. P., Armbruster, P., Folger, H., Münzenberg, G.,
Schott, H. J., Popeko, A. G., Yeremin, A. V., Andreyev, A. N., Saro, S., Janik, R., and
Leino, M. 1995a. Production and decay of 269 110. Z Phys A 350, 277–280.
Hofmann, S., Ninov, V., Hessberger, F. P., Armbruster, P., Folger, H., Münzenberg, G.,
Schott, H. J., Popeko, A. G., Yeremin, A. V., Andreyev, A. N., Saro, S., Janik, R., and
Leino, M. 1995b. The new element 111. Z Phys A 350, 281.
Hofmann, S., Ninov, V., Hessberger, F. P., Armbruster, P., Folger, H., Münzenberg, G.,
Schott, H. J., Popeko, A. G., Yeremin, A. V., Saro, S., Janik, R., and Leino, M. 1996. Z
Phys A 354, 229–230.
Hofstetter, K. J., Cable, P. R., and Beals, D. M. 1999. Field analyses of tritium at environ-
mental levels. Nucl Instr Methods in Phys Res A422, 761–766.
References 449
Kim, C. K., Seki, R., Morita, S., Yamasaki, S., Tsumura, A., Takaku, Y., Igarashi, Y., and
Yamamoto, M. 1991. Application of a high-resolution inductively coupled plasma mass-
spectrometer to the measurement of long-lived radionuclides. J Anal Atom Spectrom 6(3),
205–209.
Kim, C. S., Kim, C. K., Lee, J. I., and Lee, K. J. 2000. Rapid determination of Pu isotopes
and atom ratios in small amounts of environmental samples by an on-line sample pre-
treatment system and isotope dilution high resolution inductively coupled plasma mass
spectrometry. J Anal Atom Spectrom 15(3), 247–255.
King S. E., Phillips G. W., August R. A., Beach L. A., Cutchin J. H., and Castaneda C.
1987. Detection of tritium using accelerator mass-spectrometry. Nucl Instrum Meth B
29(1–2), 14–17.
Kingston, H. M. and Haswell, S., eds. 1997. Microwave Enhanced Chemistry: Fundamen-
tals, Sample Preparation, and Applications. Washington, DC: ACS.
Kirbach, U. W., Folden, C. M., III, Ginter, T. N., Gregorich, K. E., Lee, D. M., Ninov,
V., Omtvedt, J. P., Patin, J. B., Seward, N. K., Strellis, D. A., Sudowe, R., Türler,
A., Wilk, P. A., Zielinski, P. M., Hoffman, D. C., and Nitsche, H. 2002. The Cryo-
Thermochromatographic Separator (CTS): A new rapid separation and a detection system
for on-line chemical studies of highly volatile osmium and hassium (Z = 108) tetroxides.
Nucl Instrum Meth A 484, 587–594.
Kirby, H. W. and Salutsky, M. L. 1964. The Radiochemistry of Radium. National Research
Council Report NAS-NS 3057. Washington, DC: National Research Council.
Kleinberg, J. and Cowan, G. A. 1960. The Radiochemistry of Fluorine, Chlorine, Bromine
and Iodine. National Research Council Report NAS-NS 3005. Washington, DC: National
Research Council.
Knoll, G. F. 1989. Radiation Detection and Measurement. 2nd ed. New York, NY:
Wiley.
Knoll, G. F. 2000. Radiation Detection and Measurement. 3rd ed. New York, NY: John
Wiley.
Kocher, D. C. 1977. Nuclear Decay Data for Radionuclides Occurring in Routine Re-
leases from Nuclear Fuel Cycle Facilities. Oak Ridge National Laboratory Report
ORNL/NUREG/TM-102. Oak Ridge, TN.
Kolthoff, I. M. 1932. Theory of coprecipitation: the formation and properties of crystalline
precipitates. J Phys Chem 36, 860–881.
Kolthoff, I. M. and Elving P. J. 1971. Treatise on Analytical Chemistry. Pt. 1, Vol. 9. New
York, NY: John Wiley.
Kolthoff, I. M. and Lingane, J. J. 1939. The fundamental principles and applications of
electrolysis with the dropping mercury electrode and Heyrovsk’s polarographic method
of chemical analysis. Chem Rev 24, 1–94.
Koppenaal, D. W., Eiden, G. C., and Barinaga, C. J. 2004. Collision and reaction cells in
atomic mass spectrometry: development, status, and applications. J Anal Atom Spectrom
19(5), 561–570.
Koppenol, W. H. 2002. Naming of new elements (IUPAC Recommendations 2002). Pure
Appl Chem 74, 787–791.
Korenman, I. M. 1968. Analytical Chemistry of Low Concentrations. Jerusalem: Israel
program for scientific translations ltd.
Korkisch, J. 1989. Handbook of Ion Exchange Resins: Their Application to Inorganic
Analytical Chemistry. Boca Raton, FL: CRC Press.
Kovach, J. L. 1968. Adsorbents; review and projection in Proceedings of the tenth AEC
air cleaning conference. First, M. W. and Morgan, J. M., Jr., eds. pp. 149–166.US DOE
452 References
Report CONF-680821. Springfield, VA: Clearinghouse for federal scientific and technical
information.
Kratz, J. V. 1999a. Fast chemical separation procedures for transactinides in Heavy Elements
and Related New Phenomena. Greiner, W. and Gupta, R. K. eds. pp. 43–63. Singapore:
World Scientific.
Kratz, J. V. 1999b. Chemical properties of the transactinide elements in Heavy Elements
and Related New Phenomena. Greiner, W. and Gupta, R. K. eds. pp. 129–193. Singapore:
World Scientific.
Kraus, K. A. and Nelson, F. 1956. Anion exchange studies of the fission products in
Proceedings of the 1st International Conference on Peaceful Uses of Atomic Energy.
Vol. 7. A/CONF.8/7. pp. 113–125. New York, NY: United Nations Publishers.
Kusaka, Y. and Meinke, W. W. 1961. Rapid Radiochemical Separations. National Research
Council Report NAS-NS 3104. Washington, DC: National Research Council.
Kutschera, W. 1994. Atom counting of long-lived radionuclides. Nuclear Instruments &
Methods in Physics Research Section a-Accelerators Spectrometers Detectors and Asso-
ciated Equipment 353(1–3), 562–562.
Ladd, M. F. C. and Lee, W. H. 1964. Elementary Practical Radiochemistry. London: Cleaver-
Hume Press.
Laitinen, H. A. and Watkins, N. H. 1975. Cathodic stripping coulometry of lead. Anal Chem
47, 1352–1358.
Langmuir I. and Kingdon K. H. 1925. Thermionic effects caused by vapours of alkali metals.
P Roy Soc A 107, 61–79.
Larsen, B. S. and McEwen, C. N. 1998. Mass Spectrometry of Biological Materials. pp. 469.
New York, NY: Marcel Dekker.
Laue, C. A. and Nash, K. L., eds. 2003. Radioanalytical Methods in Interdisciplinary
Research, ACS Symposium Series 868. Washington, DC: American Chemical Society.
Lazarev, Y. A., Lobanov, Y. V., Oganessian, Y. T., Utyonkov, V. K., Abdullin, F. S.,
Buklanov, G. V., Gikal, B. N., Iliev, S., Mezentsev, A. N., Polyakov, A. N., Sedykh,
I. M., Shirokovsky, I. V., Subbotin, V. G., Sukhov, A. M., Tsyganov, Y. S., Zhuchko, V.
E., Lougheed, R. W., Moody, K. J., Wild, J. F., Hulet, E. K., and McQuaid, J. H. 1994.
Discovery of enhanced nuclear stability near the deformed shells N=162 and Z=108.
Phys Rev Lett 73, 624–627.
Lazarev, Y. A., Lobanov, Y. V., Oganessian, Y. T., Utyonkov, V. K., Abdullin, F. S., Polyakov,
A. N., Rigol, J., Shirokovsky, I. V., Tsyganov, Y. S., Iliev, S., Subbotin, V. G., Sukhov,
A. M., Buklanov, G. V., Gikal, B. N., Kutner, V. B., Mezentsev, A. N., Subotic, K., Wild,
J. F., Longheed, R. W., and Moody, K. J. 1996. Alpha decay of (273)110 - shell closure
at N=162. Phys Rev C 54, 620–625.
Lederer, C. M., Hollander, J. M., and Perlman, I. 1967. Table of Isotopes, 6th ed. New York,
NY: John Wiley.
Lee, S. H., Gastaud, J., La Rosa, J. J., Kwong, L. L. W., Povinec, P. P., Wyse, E., Fifield,
L. K., Hausladen, P. A., Di Tada, L. M., and Santos, G. M. 2001. Analysis of plutonium
isotopes in marine samples by radiometric, ICP-MS and AMS techniques. J Radioanal
Nucl Ch 248(3), 757–764.
Lee, T., Ku, T. L., Lu, H. L., and Chen, J. C. 1993. 1st detection of fallout Cs-135 and
potential applications of Cs-137/Cs-135 ratios. Geochim Cosmochim Ac 57(14), 3493–
3497.
Lewis, D., Miller, G., Duffy, C. J., Efurd, D. W., Inkret, W. C., and Wagner, S. E.
2001. Los Alamos National Laboratory thermal ionization mass spectrometry results
from intercomparison study of inductively coupled plasma mass spectrometry, thermal
References 453
Marcus, Y. and Kertes, A. S. 1969. Ion Exchange and Solvent Extraction of Metal Complexes.
New York, NY: Wiley Interscience.
Marinsky, J. A., Glendenin, L. E., and Coryell, C. D. 1947. J Am Chem Soc 69, 2781.
Martin, B., Ockenden, D. W., and Foreman, J. K. 1961. The solvent extraction of plutonium
and americium by tri-n-octylphosphine oxide. J Inorg Nucl Chem 21, 96–107.
Martin, M. J. and Blichert-Tolf, P. H. 1970. Radioactive atoms. Nucl Data Tables A8, 1–198,
1970.
McAninch, J. E., Hamilton, T. F., Brown, T. A., Jokela, T. A., Knezovich, J. P., Ognibene,
T. J., Proctor, I. D., Roberts, M. L., Sideras-Haddad, E., Southon, J. R., and Vogel J. S.
2000. Plutonium measurements by accelerator mass spectrometry at LLNL. Nucl Instrum
Meth A 172, 711–716.
McAninch, J. E., Marchetti, A. A., Bergquist, B. A., Stoyer, N. J., Nimz, G. J., Caffee,
M. W., Finkel, R. C., Moody, K. J., Sideras-Haddad, E., Buchholz, B. A., Esser, B. K.,
and Proctor, I. D. 1998. Detection of Tc-99 by accelerator mass spectrometry: preliminary
investigations. J Radioan Nucl Ch 234(1–2), 125–129.
McDowell, W. J. 1992. Photon/electron-rejecting alpha liquid scintillation (PERALS) spec-
trometry. A review. Radioactivity and Radiochemistry 3, 26–53.
McIntyre, J. I., Abel, K. H., Bowyer, T. W., Hayes, J. C., Heimbigner, T. R., Panisko, M. E.,
Reeder, P. L., and Thompson, R. C. 2001. Measurements of ambient radioxenon levels
using the automated radioxenon sampler/analyzer (ARSA). J Radioan Nucl Ch 248 (3),
629–635.
McMillan, E. M. and Abelson, P. H. 1940. Radioactive element 93. Phys Rev 57, 1185.
Meeks, A. M., Giaquinto, J. M., and Keller, J. M. 1998. Application of ICP-MS radionuclide
analysis to “Real World” samples of Department of Energy radioactive waste. J Radioan
Nucl Ch 234(1–2), 131–135.
Miley, H. S., Bowyer, S. M., Hubbard, C. W., McKinnon, A. D., Perkins, R. W., Thompson,
R. C., and Warner, R. A. 1998. A description of the DOE radionuclide aerosol sam-
pler/analyzer for the Comprehensive Test Ban Treaty. J Radioan Nucl Ch 235 (1–2),
83–87.
Miller, D. R. 1988. Free jet sources in Atomic and Molecular Beam Methods. Vol. 1.
Chapter 2. Scoles, G., ed. Oxford: Oxford University Press.
Moghissi, A. A., Bretthauer, E. W., and Compton, E. H. 1973. Separation of water from
biological and environmental samples for tritium analysis. Anal Chem 45, 1565–1566.
Momyer, F., Jr. 1960. The Radiochemistry of the Rare Gases. National Academy of
Sciences—National Research Council Report NAS-NS 3025. Washington, DC: National
Research Council.
Montaser, A. 1992. Inductively Coupled Plasmas in Analytical Atomic Spectrometry. New
York, NY: John Wiley.
Montaser, A. 1998. Inductively Coupled Plasma Mass Spectrometry. pp. 964. New York,
NY: Wiley-VCH.
Moreno, J. M. B., Betti, M., and Alonso, J. I. G. 1997. Determination of neptunium
and plutonium in the presence of high concentrations of uranium by ion chromatog-
raphy inductively coupled plasma mass spectrometry. J Anal Atom Spectrom 12(3), 355–
361.
Münzenberg, G., Armbruster, P., Folger, H., Hessberger, F. P., Hoffman, S., Keller, K.,
Poppensieker, K., Reisdorf, W., Schmidt, K.-H., and Schott, H.-J. 1984. The identification
of element 108. Z Phys A, 317, 235–236.
Münzenberg, G., Armbruster, P., Hessberger, F. P., Hofmann, S., Poppensieker, K., Reisdorf,
W., Schneider, J. H. R., Schneider, W. R. W., Schmidt, K.-H., Sahm, C.-C., and Vermeulen,
References 455
D. 1982. Evidence for element 109 from one correlated decay sequence following the
fussion of 58 Fe with 209 Bi. Z Phys A 315, 145.
Münzenberg, G., Hofmann, S., Hessberger, F. P., Reisdorf, W., Schmidt, K. H., Schneider,
J. H. R., Armbruster, P., Sahm, C.-C., and Thuma, B. 1981. Identification of element 107
by α correlation chains. Z Phys A 300, 107.
Muramatsu, Y., Hamilton, T., Uchida, S., Tagami, K., Yoshida, S., and Robison, W. 2001.
Measurement of Pu-240/Pu-239 isotopic ratios in soils from the Marshall Islands using
ICP-MS. Sci Total Environ 278(1–3), 151–159.
Muramatsu, Y., Uchida, S., Tagami, K., Yoshida, S., and Fujikawa, T. 1999. Determination
of plutonium concentration and its isotopic ratio in environmental materials by ICP-
MS after separation using ion-exchange and extraction chromatography. J Anal Atom
Spectrom 14(5), 859–865.
Murthy, G. K. and Campbell, J. E. 1960. A simplified method for the determination of
Iodine131 in milk. J Dairy Science 8, 1042–1049.
Murthy, G. K., Coakley, J. E., and Campbell, J. E. 1960. A method for the elimination of
ashing in strontium-90 determination of milk. Dairy Sci 43, 151–154.
NAS-NRC 1960a. National Academy of Sciences—National Research Council Reports
NAS-NS 30xx. The Radiochemistry of ELEMENT. Washington, DC: Volumes in the
series published by the National Research Council, 1960–1988.
NAS-NRC 1960b. National Academy of Sciences—National Research Council Reports
NAS-NS-31xx: Nuclear Science Series. Radiochemistry Technique. Available from
NTIS, Springfield, VA. Volumes in the series published 1960–1988.
Navratil, O., Hala, J., Kopunec, R., Macasek, F., Mikulaj, V., and Leseticki, L. 1992. Nuclear
Chemistry. New York, NY: Ellis Horwood, PTR Prentice Hall.
NCRP 1976a. National Council on Radiation Protection and Measurements, Report No. 47.
Tritium Measurement Techniques. Bethesda, MD: NCRP.
NCRP 1976b. National Council on Radiation Protection and Measurements, Report No. 50.
Environmental Radiation Measurements. Bethesda, MD: NCRP.
NCRP 1979. National Council on Radiation Protection and Measurements,Report No. 62.
Tritium in the Environment. Bethesda, MD: NCRP.
NCRP 1983. National Council on Radiation Protection and Measurements Report No. 75.
I-129: Evaluation of Releases from Nuclear Power Generation. Bethesda, MD: NCRP.
NCRP 1985a. National Council on Radiation Protection and Measurements, Report No. 81.
Carbon-14 in the Environment. Bethesda, MD: NCRP.
NCRP 1985b. National Council of Radiation Protection and Measurements Report
No. 58. A Handbook of Radioactivity Measurement Procedures. 2nd ed. Bethesda, MD:
NCRP.
NCRP 1987a. National Council on Radiation Protection and Measurements, Report No. 94.
Exposure of the Population in the United States and Canada from Natural Background
Radiation. Bethesda, MD: NCRP.
NCRP 1987b. National Council on Radiation Protection and Measurements, Report
No. 87. Use of Bioassay Procedures for Assessment of Internal Radionuclide Deposition,
Bethesda, MD: NCRP, 1987.
Nelms S. 2005. Inductively Coupled Plasma Mass Spectrometry Handbook. Victoria,
Australia: Blackwell Publishing.
Nichols, M. 2006. Monte Carlo Simulation of Beta Particle Detection. Ph.D. thesis. Atlanta,
GA 30332: Georgia Institute of Technology.
Ninov, V., Gregorich, K. E., Loveland, W., Ghiorso, A., Hoffman, D. C., Lee, D. M., Nitsche,
H., Swiatecki, W. J., Kirbach, U. W., Laue, C. A., Adams, J. L., Patin, J. B., Shaughnessy,
456 References
D. A., Strellis, D. A., and Wilk, P. A. 1999. Observation of superheavy nuclei produced
in the reaction of 86 Kr with 208 Pb. Phys Rev Lett 83, 1104–1107.
Ninov, V., Gregorich, K. E., Loveland, W., Ghiorso, A., Hoffman, D. C., Lee, D. M., Nitsche,
H., Swiatecki, W. J., Kirbach, U. W., Laue, C. A., Adams, J. L., Patin, J. B., Shaughnessy,
D. A., Strellis, D. A., and Wilk, P. A. 2002. Retraction of 1999 published paper. Phys
Rev Lett 89, 039901–1.
NIST 1994. National Institute of Standards and Technology, Technical Note 1297. Guide-
lines for Evaluating and Expressing the Uncertainty of NIST Measurement Results.
Gaithersburg, MD: NIST.
Odom, R. W., Strathman, M. D., Buttrill, S. E., and Baumann, S. M. 1990. Nondestructive
imaging detectors for energetic particle beams. Nucl Instrum Meth B 44(4), 465–472.
Oganessian, Y. T. 2001. The synthesis and decay properties of the heaviest elements. Nucl
Phys A 685, 17c–36c.
Oganessian, Y. T. 2002. Synthesis and properties of even-even isotopes with Z = 110—116
in 48 Ca induced reactions. J Nucl Radiochem Sci 3, 5–8.
Oganessian, Y. T., Yeremin, A. V., Popeko, A. G., Bogomolov, S. L., Buklanov, G. V.,
Chelnokov, M. L., Chepigin, V. I., Gikal, B. N., Gorshkov, V. A., Gulbekian, G. G., Itkis,
M. G., Kabachenko, A. P., Lavrentev, A. Y., Malyshev, O. N., Rohac, J., Sagaidak, R. N.,
Hofmann, S., Saro, S., Giardinas, G., and Morita, K. 1999a. Synthesis of nuclei of the
superheavy element 114 in reactions induced by 48 Ca. Nature 400, 242–245.
Oganessian, Y. T., Utyonkov, V. K., Lobanov, Y. V., Abdullin, F. S., Polyakov, A. N.,
Shirokovsky, I. V., Tsyganov, Y. S., Gulbekian, G. G., Gobomolov, S. L., Gikal, B. N.,
Mezentsev, A. N., Iliev, S., Subbotin, V. G., Sukhov, A. M., Buklanov, G. V., Subotic,
K., Itkis, M. G., Moody, K. J., Wild, J. F., Stoyer, N. J., Stoyer, M. A., and Lougheed,
R. W. 1999b. Synthesis of superheavy nuclei in the 48 Ca + 244 Pu reaction. Phys Rev Lett
83, 3154–7.
Oganessian, Y. T., Yeremin, A. Y., Gulbekian, G. G., Bogomolov, S. L., Chepigin, V. I.,
Gikal, B. N., Gorshkov, V. A., Itkis, M. G., Kabachenko, A. P., Kutner, V. B., Lavrentev,
A. Y., Malyshev, O. N., Popeko, A. G., Rohac, J., Sagaidak, R. N., Hofmann, S.,
Münzenberg, G., Veselsky, M., Saro, S., Iwasa, N., and Morita, K. 1999c. Search for
new isotopes of element 112 by irradiation of 238 U with 48 Ca. Eur Phys J A 5, 63–
68.
Oganessian, Y. T., Utyonkov, V. K., Lobanov, Y. V., Abdullin, F. S., Polyakov, A. N.,
Shirokovsky, I. V., Tsyganov, Y. S., Gulbekian, G. G., Bogomolov, S. L., Gikal, B. N.,
Mezentsev, A. N., Iliev, S., Subbotin, V. G., Sukhov, A. M., Ivanov, O. V., Buklanov,
G. V., Subotic, K., Itkis, M. G., Moody, K. J., Wild, J. F., Stoyer, N. J., Stoyer, M. A., and
Lougheed, R. W. 2000a. Synthesis of superheavy nuclei in the Ca-48+Pu-244 reaction:
(288)114 - art. no. 041604. Phys Rev C 62, 041604 (R) 1–4.
Oganessian, Y. T., Utyonkov, V. K., Lobanov, Y. V., Abdulin, F. S., Polyakov, A. N.,
Shirokovsky, I. V., Tsyganov, Y. S., Gulbekian, G. G., Bogomolov, S. L., Gikal, B. N.,
Mezentsev, A. N., Iliev, S., Subbotin, V. G., Sukhov, A. M., Ivanov, O. V., Buklanov,
G. V., Subotic, K., Itkis, M. G., Moody, K. J., Wild, J. F., Stoyer, N. J., Stoyer, M. A.,
Lougheed, R. W., Laue, C. A., Karelin, Y. A., and Tatarinov, A. N. 2000b. Observation
of the decay of 292 116. Phys Rev C 63, 011301(R) 1–2.
Oganessian, Y. T., Utyonkov, V. K., Lobanov, Y. V., Abdullin, F. S., Polyakov, A. N.,
Shirokovsky, I. V., Tsyganov, Y. S., Gulbekian, G. G., Bogomolov, S. L., Gikal, B. N.,
Mezentsev, A. N., Iliev, S., Subbotin, V. G., Sukhov, A. M., Ivanov, O. V., Buklanov,
G. V., Subotic, K., Voinov, A. A., Itkis, M. G., Moody, K. J., Wild, J. F., Stoyer, N. J.,
Stoyer, M. A., Lougheed, R. W., and Laue, C. A. 2002. Eur Phys J A 15, 201–204.
References 457
Oganessian, Y. T., Utyonkoy, V. K., Lobanov, Y. V., Abdullin, F. S., Polyakov, A. N.,
Shirokovsky, I. V., Tsyganov, Y. S., Gulbekian, G. G., Bogomolov, S. L., Mezentsev,
A. N., Iliev, S., Subbotin, V. G., Sukhov, A. M., Voinov, A. A., Buklanov, G. V., Subotic,
K., Zagrebaev, V. I., Itkis, M. G., Patin, J. B., Moody, K. J., Wild, J. F., Stoyer, M. A.,
Stoyer, N. J., Shaughnessy, D. A., Kenneally, J. M., and Lougheed, R. W. 2004a. Phys
Rev C 69, 021601(R).
Oganessian, Y. T., Yeremin, A. V., Popeko, A. G., Malyshev, O. N., Belozerov, A. V.,
Buklanov, G. V., Chelnokov, M. L., Chepigin, V. I., Gorshkov, V. A., Hofmann, S., Itkis,
M. G., Kabachenko, A. P., Kindler, B., Munzenberg, G., Sagaidak, R. N., Saro, S., Schott,
H.-J., Streicher, B., Shutov, A. V., Svirikhin, A. I., and Vostokin, G. K. 2004b. Eur Phys
J A 19, 3–6.
Oganessian, Y. T., Utyonkoy, V. K., Lobanov, Y. V., Abdullin, F. S., Polyakov, A. N.,
Shirokovsky, I. V., Tsyganov, Y. S., Gulbekian, G. G., Bogomolov, S. L., Gikal, B. N.,
Mezentsev, A. N., Iliev, S., Subbotin, V. G., Sukhov, A. M., Voinov, A. A., Buklanov,
G. V., Subotic, K., Zagrebaev, V. I., Itkis, M. G., Patin, J. B., Moody, K. J., Wild, J. F.,
Stoyer, M. A., Stoyer, N. J., Shaughnessy, D. A., Kenneally, J. M., and Lougheed, R. W.
2004c. Phys Rev C 69, 054607.
Omtvedt, J. P., Alstad, J., Breivik, H., Dyve, J. E., Eberhardt, K., Folden III, C. M., Ginter,
T., Gregorich, K. E., Hult, E. A., Johansson, M., Kirbach, U. W., Lee, D. M., Mendel, M.,
Nahler, A., Ninov, V., Omtvedt, L. A., Patin, J. B., Skarnemark, G., Stavsetra, L., Sudowe,
R., Wiehl, N., Wierczinski, B., Wilk, P. A., Zielinski, P. M., Kratz, J. V., Trautmann, N.,
Nitsche, H., and Hoffman, D. C. 2002. SISAK liquid-liquid extraction experiments with
pre-separated 257 Rf. J Nucl Radiochem Sci 3, 121–124.
Ostlund, H. G. and Mason, A. S. 1985. Atmospheric tritium in: Tritium Laboratory Data
Report No. 14. Rosenstiel School of Marine and Atmospheric Science. Miami, FL:
University of Miami.
OTA 1983. Office of Technology Assessment, NTIS order #PB84-114917. Habitability of
the Love Canal area: An analysis of the technical basis for the decision on the habit-
ability of the emergency declaration area. Washington, DC: U.S. Government Printing
Office.
Oughton, D. H., Fifield, L. K., Day, J. P., Cresswell, R. C., Skipperud, L., Di Tada,
M. L., Salbu, B., Strand, P., Drozcho, E., and Mokrov, Y. 2000. Plutonium from Mayak:
Measurement of isotope ratios and activities using accelerator mass spectrometry. Environ
Sci Technol 34(10), 1938–1945.
Oughton, D. H., Skipperud, L., Fifield, L. K., Cresswell, R. G., Salbu, B., and Day, P. 2004.
Accelerator mass spectrometry measurement of Pu-240/Pu-239 isotope ratios in Novaya
Zemlya and Kara Sea sediments. Appl Radiat Isotopes 61(2–3), 249–253.
Pajo, L., Schubert, A., Aldave, L., Koch, L., Bibilashvili, Y. K., Dolgov, Y. N., and
Chorokhov, N. A. 2001a. Identification of unknown nuclear fuel by impurities and phys-
ical parameters. J Radioanal Nucl Ch 250(1), 79–84.
Pajo L., Tamborini G., Rasmussen G., Mayer K., and Koch L. 2001b. A novel isotope anal-
ysis of oxygen in uranium oxides: comparison of secondary ion mass spectrometry, glow
discharge mass spectrometry and thermal ionization mass spectrometry. Spectrochim
Acta B 56(5), 541–549.
Pappas, R. S., Ting, B. G., and Paschal, D. C. 2004. Rapid analysis for plutonium-239 in
1 ml of urine by magnetic sector inductively coupled plasma mass spectrometry with a
desolvating introduction system. J Anal Atom Spectrom 19(6), 762–766.
Parrington, J. R., Knox, H. D., Breneman, S. L., Baum, E. M., and Feiner, F. 1996. Chart
of the Nuclides. 15th ed. San Jose, CA: GE Nuclear Energy.
458 References
Pasternack, B. S. and Harley, N. H. 1971. Detection limits for radionuclides in the analysis
of multi-component gamma ray spectrometer data. Nucl Inst and Methods 91, 533–540.
Penneman, R. A. and Keenan, T. K. 1960. The Radiochemistry of Americium and Curium.
National Research Council Report NAS-NS 3006. Washington, DC: National Research
Council.
Percival, D. R. and Martin, D. B. 1974. Sequential determination of radium-226, radium-
228, actinium and thorium isotopes in environmental and process waste samples. Anal
Chem 46, 1742–1749.
Perkins, R. W., Robinson, D. E., Thomas, C. W., and Young, J. A. 1989. IAEA-SM-306/125.
Comparison of nuclear accident and nuclear test debris in Proceedings of the Interna-
tional Symposium on Environmental Contamination Following a Major Nuclear Accident.
pp. 111–139. Vienna: IAEA, October 16–20, 1989.
Perrier, C. and Segrè, E. 1937. Some chemical properties of element 43. J Chem Phys 5,
712.
Perrin, R. E., Knobeloch, G. W., Armijo, V. M., and Efurd, D. W. 1985. Isotopic analysis of
nanogram quantities of plutonium by using a sid ionization source. Int J Mass Spectrom
Ion Processes 64(1), 17–24.
Perrin, D. D. 1964.Organic Complexing Reagents: Structure, Behavior, and Application to
Inorganic Analysis. New York: Interscience Publishers.
Pershina, V. 1998a. Solution chemistry of element 105-Part I: Hydrolysis of group 5 cations:
Nb, Ta, and Pa. Radiochim Acta 80, 65–73.
Pershina, V. 1998b. Solution chemistry of element 105-part II: Hydrolysis and complex
formation of Nb, Ta, Ha and Pa in HCl solutions. Radiochim Acta 80, 75–84.
Pershina, V. and Bastug, T. 1999. Solution chemistry of element 105-Part III: Hydrolysis
and complex fomation of Nb, Ta, Db and Pa in HF and HBr solutions. Radiochim Acta
84, 79–84.
Pershina, V. and Bastug, T. 2000. The electronic structure and properties of group 7 oxychlo-
rides, MO3 Cl, where M=Tc, Re, and element 107, Bh. J Chem Phys 113, 1441–1446.
Pershina, V. and Fricke, B. 1996. Group 6 dioxydichlorides M2 Cl2 (M=Cr, Mo, W, and
element 106, Sg)—the electronic structure and thermochemical stability. J Phys Chem
100, 8748–8751.
Pershina, V. and Fricke, B. 1999. Electronic structure and chemistry of the heaviest elements
in Heavy Elements and Related New Phenomena. vol. 1. pp. 184–262. Greiner, W. and
Gupta, R. K., eds. Singapore: World Scientific.
Pershina, V. and Hoffman, D. C. 2003. The chemistry of the heaviest elements in Theoretical
chemistry and physics of heavy and superheavy elements in the series, “Progress in
Theoretical Chemistry and Physics”. Chapter 3. pp. 55–114. Kaldor, U. and Wilson, S.,
eds. Dordrecht, The Netherlands: Kluwer Academic Publishers.
Pershina, V. and Kratz, J. V. 2001. Solution chemistry of element 106: Theoretical predic-
tions of hydrolysis of group 6 cations Mo, W and Sg. Inorg Chem 40, 776–780.
Pershina, V., Sepp, W.-D., Bastug, T., Fricke, B., and Ionova, G. V. 1992. Relativistic effects
in physics and chemistry of element 105. III. Electronic structure of hahnium oxyhalides
as analogs of group 5 elements oxyhalides, J Chem Phys 97, 1123–1131.
Pershina, V., Bastug, T., Fricke, B., and Varga, S. 2001. The electronic structure and prop-
erties of group 8 oxides MO4 , where M=Ru, Os, and Element 108, Hs. J Chem Phys
115, 792–799.
Pershina, V., Bastug, T., Jacob, T., Fricke, B., and Varga, S. 2002. Intermetallic compounds
of the heaviest elements: The electronic structure and bonding of dimers of element 112
and its homolog Hg. Chem Phys Lett 365, 176–183.
References 459
Pibida, L., Nortershauser, W., Hutchinson, J. M. R., and Bushaw, B. A. 2001. Evaluation of
resonance ionization mass spectrometry for the determination of Cs-135/Cs-137 isotope
ratios in low-level samples. Radiochimica Acta 89(3), 161–168.
Pitzer, K. S. 1975. Are elements 112, 114, and 118 relatively inert gases? J Chem Phys 63,
1032–1033.
Platzner, I. T. 1997. Modern Isotope Ratio Mass Spectrometry. New York, NY: John
Wiley.
Platzner, I. T., Becker, J. S., and Dietze, H. J. 1999. Stability study of isotope ratio mea-
surements for uranium and thorium by ICP-QMS. Atomic Spectroscopy 20(1), 6–12.
Porter, C. R. and Kahn, B. 1964. Improved determination of strontium-90 in milk by an
ion-exchange method. Anal Chem 36, 676–678.
Porter, C. R., Kahn, B., Carter, M. W., Rehnberg, G. L., and Pepper, E. W. 1967. De-
termination of radiostrontium in food and other environmental media. Env Sci Tech 1,
745–750.
Poupard, D. and Jouniaux, B. 1990. Determination of picogram quantities of americium
and curium by thermal ionization mass-spectrometry (Tims). Radiochimica Acta 49(1),
25–28.
Praphairaksit, N. and Houk, R. S. 2000. Reduction of mass bias and matrix effects in
inductively coupled plasma mass spectrometry with a supplemental electron source in a
negative extraction lens. Anal Chem 72, 4435–4440.
Priest, N. D., Pich, G. M., Fifield, L. K., and Cresswell, R. G. 1999. Accelerator mass
spectrometry for the detection of ultra-low levels of plutonium in urine, including that
excreted after the ingestion of Irish sea sediments. Radiat Res 152(6), S16–S18.
Probst, T. U. 1996. Studies on the long-term stabilities of the background of radionuclides
in inductively coupled plasma mass spectrometry (ICP-MS)—A review of radionuclide
determination by ICP-MS. Fresen J Anal Chem 354(7–8), 782–787.
Raisbeck, G. M. and Yiou, F. 1999. I-129 in the oceans: origins and applications. Sci Total
Environ 238, 31–41.
Ratliff, T. A. 2003. Laboratory Quality Assurance System. 3rd ed. pp. 4, 31–32. Hoboken,
NJ: John Wiley.
Rehkamper, M., Schonbachler, M., and Stirling, C. H. 2001. Multiple collector ICP-MS: In-
troduction to instrumentation, measurement techniques and analytical capabilities. Geo-
standard Newslett 25(1), 23–40.
Reyes, L. H., Gayon, J. M. M., Alonso, J. I. G., and Sanz-Medel, A. 2003. Determination
of selenium in biological materials by isotope dilution analysis with an octapole reaction
system ICP-MS. J Anal Atom Spectrom 18(1), 11–16.
Richardson, S. D. 2001. Mass spectrometry in environmental sciences. Chem Rev 101(2),
211–254.
Riegel, J., Deissenberger, R., Herrmann, G., Kohler, S., Sattelberger, P., Trautmann, N.,
Wendeler, H., Ames, F., Kluge, H. J., Scheerer, F., and Urban, F. J. 1993. Resonance
ionization mass spectroscopy for trace analysis of neptunium. Appl Phys B Laser O
56(5), 275–280.
Rieman, W. and Walton, H. F. 1970. Ion Exchange in Analytical Chemistry. New York, NY:
Pergamon Press.
Roberts, M. L. and Caffee, M. W. 2000. I-129 interlaboratory comparison: Phase II results.
Nucl Instrum Meth Phys B 172, 388–394.
Rodushkin, I., Lindahl, P., Holm, E., and Roos, P. 1999. Determination of plutonium con-
centrations and isotope ratios in environmental samples with a double-focusing sector
field ICP- MS. Nucl Instrum Meth Phys A 423(2–3), 472–479.
460 References
Rokop, D. J., Perrin, R. E., Knobeloch, G. W., Armijo, V. M., and Shields, W. R. 1982.
Thermal ionization mass spectrometry of uranium with electrodeposition as a loading
technique. Anal Chem 54(6), 957–960.
Rondinella, V. V., Betti, M., Bocci, F., Hiernaut, T., and Cobos, J. 2000. IC-ICP-MS applied
to the separation and determination of traces of plutonium and uranium in aqueous
leachates and acid rinse solutions of UO2 doped with Pu-238. Microchem J 67(1–3),
301–304.
Rosen, A. 1998. Twenty to thirty years of DV-X α calculation: A survey of accuracy and
application. Adv Quant Chem 29, 1–30.
Rosson, R., Jakiel, R., Klima, S., Kahn, B., and Fledderman, P. 2000. Correcting tritium
concentrations in water vapor monitored with silica gel. Health Phys 78, 68–73.
Rucklidge, J. 1995. Accelerator mass-spectrometry in environmental geoscience—A re-
view. Analyst 120(5), 1283–1290.
Russo, R. E., Mao, X. L., Liu, H. C., Gonzalez, J., and Mao, S. S. 2002. Laser ablation in
analytical chemistry—A review. Talanta 57(3), 425–451.
Sahoo, S. K., Yonehara, H., Kurotaki, K., Fujimoto, K., and Nakamura, Y. 2002. Precise
determination of U-235/U-238 isotope ratio in soil samples by using thermal ionisation
mass spectrometry. J Radioanal Nucl Ch 252(2), 241–245.
Schädel, M. 1995. Chemistry of the transactinide elements. Radiochim Acta 70/71, 207–
223.
Schädel, M., Brüchle, W., Dressler, R., Eichler, B., Gäggeler, H. W., Günther, R., Gregorich,
K. E., Hoffman, D. C., Hübener, S., Jost, D. T., Kratz, J. V., Paulus, W., Schumann, D.,
Timokhin, S., Trautmann, N., Türler, A., Wirth, G., and Yakuschev, A. 1997a. Chemical
properties of element 106 (seaborgium). Nature Lett 388, 55–57.
Schädel, M., Brüchle, W., Schausten, B., Schimpf, E., Jäger, E., Wirth, G., Günther, R.,
Kratz, J. V., Paulus, W., Seibert, A., Thorle, P., Trautmann, N., Zauner, S., Schumann, D.,
Andrassy, M., Misiak, R., Gregorich, K. E., Hoffman, D. C., Lee, D. M., Sylwester, E. R.,
Nagame, Y., and Oura, Y. 1997b. First aqueous chemistry with seaborgium (element 106).
Radiochim Acta 77, 149–159.
Schädel, M., Brüchle, W., Jäger, E., Schausten, B., Wirth, G., Paulus, W., Günther, R.,
Eberhardt, K., Kratz, J. V., Seibert, A., Strub, E., Thörle, P., Trautmann, N., Waldek,
W., Zauner, S., Schumann, D., Kirbach, U., Kubica, B., Misiak, R., Nagame, Y., and
Gregorich, K. E. 1998. Aqueous chemistry of seaborgium (Z=106). Radiochim Acta 83,
163–165.
Schaumloffel, D., Giusti, P., Zoriy, M. V., Pickhardt, C., Szpunar, J., Lobinski, R., and
Becker, J. S. 2005. Ultratrace determination of uranium and plutonium by nano-volume
flow injection double-focusing sector field inductively coupled plasma mass spectrometry
(nFI-ICP-SFMS). J Anal Atom Spectrom 20(1), 17–21.
Schmitt, G., Markell, C., and Hagen, D. 1990. Membrane approach to solid-phase extrac-
tions. Anal Chim Acta 236, 157–164.
Schneier, B. 2000. Secrets & Lies: Digital Security in a Networked World. New York, NY:
John Wiley.
Schulze, J., Auer, M., and Werzi, R. 2000. Low level radioactivity measurement in support
of the CTBTO. Appl Radiat Isotopes 53, 23–30.
Scoles, G. 1988. Atomic and Molecular Beam Methods. Vol. 1. Oxford: Oxford University
Press.
Seaborg, G. T. 1945. The chemical and radioactive properties of the heavy elements. Chem
Eng News 23, 2190.
Seaborg, G. T. 1967. Actinides reviews. Actin Rev 1, 3–38.
References 461
Seaborg, G. T., and Keller, O. L., Jr. 1986. Future elements in The Chemistry of the Actinide
Elements. 2nd ed. Katz, J. J. Seaborg, G. T. and Morss, L. R. eds. pp. 1635–1643. London:
Chapman and Hall.
Seaborg, G. T., Wahl, A. C., and Kennedy, J. W. 1946. Radioactive element 94 from
deuterons on uranium. Phys Rev 69, 367.
Sekine, T. and Hasegaw, Y. 1977. Solvent Extraction Chemistry: Fundamentals and Appli-
cations. New York, NY: Marcel Dekker.
Seth, M., Schwerdtfeger, P., Dolg, M., Faegri, K., Hess, B. A., and Kaldor, U. 1996. Large
relativistic effects in molecular properties of the hydride of superheavy element 111.
Chem Phys Lett 250, 461–465.
Seth, M., Cooke, F., Schwerdtfeger, P., Heully, J.-L.,and Pelissier, M. 1998a. The chemistry
of the superheavy elements. II. The stability of high oxidation states in group 11 elements:
Relativistic coupled cluster calculations for the di-, tetra- and hexafluoro metallates of
Cu, Ag, Au, and element 111. J Chem Phys 109, 3935–3943.
Seth, M., Faegri, K., and Schwerdtfeger, P. 1998b. The stability of the oxidation state +4
in group 14 compounds from carbon to element 114. Angew Chem Int Ed Engl 37,
2493–2496.
Seubert, A. 2001. On-line coupling of ion chromatography with ICP-AES and ICP-MS.
Trac-Trend in Anal Chem 20(6–7), 274–287.
Shen, C. C., Cheng, H., Edwards, R. L., Moran, S. B., Edmonds, H. N., Hoff, J. A., and
Thomas, R. B. 2003. Measurement of attogram quantities of Pa-231 in dissolved and
particulate fractions of seawater by isotope dilution thermal ionization mass spectroscopy.
Anal Chem 75(5), 1075–1079.
Sherwin, C. W. 1951.Vacuum evaporation of radioactive materials. Rev Sci Instrum 22 (5),
339–341.
Shleien, B. ed. 1992. The Health Physics and Radiological Health Handbook. Silver
Springs, MD: Scinta.
Shleien, B., Slaback, L., Jr., and Birky, B. 1998. Handbook of Health Physics and Radio-
logical Health. 3rd ed. Baltimore, MD: Lippincott.
Shuping, R. E., Philips, C. R., and Moghissi, A. A. 1970. Krypton-85 levels in the envi-
ronment determined from dated krypton gas samples. Radiol Health Data R 11, 667–
672.
Shuter, E. and Teasdale, W. E. 1989. Application of drilling coring, and sampling techniques
to test holes and wells in Techniques of Water-Resources Investigations. Chapter F1.
Book 2. Reston, VA: US Geological Survey.
Sill, C. W. and Olson, D. G. 1970. Sources and prevention of recoil contamination of
solid-state alpha detectors. Anal Chem 42, 1596–1607.
Sill, C. W. and Williams, R. L. 1969. Radiochemical determination of uranium and the
transuranium elements in process solutions and environmental samples. Anal Chem
41(12), 1624–1632.
Sill, C. W., Puphal, K. W., and Hindman, F. D. 1974. Simultaneous determination of alpha-
emitting nuclides of radium through californium in soil. Anal Chem 46, 1725–1737.
Silva, R., Harris, J., Nurmia, M., Eskola, K., and Ghiorso, A. 1970. Chemical separation of
rutherfordium. Inorg Nucl Chem Lett 6, 871–877.
Skipperud, L. and Oughton, D. H. 2004. Use of AMS in the marine environment. Environ
Int 30(6), 815–825.
Smart, J. E. 1998. Real-time airborne radiation analysis and collection (RTARAC) searching
for airborne species characteristic to nuclear proliferation. J Radioanal Nucl Ch. 235,
105–108.
462 References
Underhill, D. W. 1996. The adsorption of argon, krypton and xenon on activated charcoal.
Health Phys 71 (2), 160–166.
UNSCEAR 2000a. United Nations Scientific Committee on the Effects of Atomic Radiation,
Annex B. Exposures from Natural Radiation Sources. Vienna: UN.
UNSCEAR 2000b. United Nations Scientific Committee on the Effects of Atomic Radiation,
Annex C. Exposures from Man-made Sources of Radiation. Vienna: UN.
Vanhaecke, F. and Moens, L. 1999. Recent trends in trace element determination and speci-
ation using inductively coupled plasma mass spectrometry. Fresen J Anal Chem 364(5),
440–451.
Vertes, A., Nagy, S., and Klencsar Z., eds. 2003. Handbook of Nuclear Chemistry. 5 books.
Dordrecht: Kluwer Academic.
Vonderheide, A. P., Montes-Bayon, M., and Caruso, J. A. 2002. Development and appli-
cation of a method for the analysis of brominated flame retardants by fast gas chro-
matography with inductively coupled plasma mass spectrometric detection. J Anal Atom
Spectrom 17(11), 1480–1485.
Waber, J. T. and Averill, F. W. 1974. Molecular orbitals of PtF6 and E110 F6 calculated by
the self-consistent multiple scattering X α method. J Chem Phys 60, 4460–4470.
Wahl, A. C. and Bonner, N. A. 1951. Radioactivity Applied to Chemistry. New York: John
Wiley.
Wakat, M. A. 1971. Catalogue of γ -rays emitted by radionuclides. Nucl Data Tables 8,
445–666.
Walker, R. L., Carter, J. A., and Smith, D. H. 1981. A bulk resin bead procedure to obtain
uranium and plutonium from radioactive solutions for mass-spectrometric analysis. Anal
Lett Pt A 14(19), 1603–1612.
Walton, H. F. and Rocklin, R. D. 1990. Ion Exchange in Analytical Chemistry. Boca Raton,
FL: CRC Press.
Wang, Z., Kahn, B., and Valentine, J. D. 2002. Efficiency calculation and coincidence
summing correction for germanium detectors by Monte Carlo simulation. IEEE T Nucl
Sci 49, 1925–1931.
Watanabe, K., Iguchi, T., Ogita, T., Uritani, A., and Harano, H. 2001. Development of failed
fuel detection and location technique using resonance ionization mass spectrometry. J
Nucl Sci Technol 38(10), 844–849.
Weast, R. C. 1985. CRC Handbook of Chemistry and Physics. Boca Raton, FL: CRC
Press.
Wendt, K., Bhowmick, G. K., Bushaw, B. A., Herrmann, G., Kratz, J. V., Lantzsch,
J., Muller, P., Nortershauser, W., Otten, E. W., Schwalbach, R., Seibert, U. A., Traut-
mann, N., and Waldek, A. 1997. Rapid trace analysis of Sr-89, Sr-90 in environmental
samples by collinear laser resonance ionization mass spectrometry. Radiochim Acta 79(3),
183–190.
Wendt, K., Blaum, K., Bushaw, B. A., Gruning, C., Horn, R., Huber, G., Kratz, J. V., Kunz,
P., Muller, P., Nortershauser, W., Nunnemann, M., Passler, G., Schmitt, A., Trautmann, N.,
and Waldek, A. 1999. Recent developments in and applications of resonance ionization
mass spectrometry. Fresen J Anal Chem 364(5), 471–477.
Wendt, K., Trautmann, N., and Bushaw, B. A. 2000. Resonant laser ionization mass spec-
trometry: An alternative to AMS? Nucl Instrum Meth B 172, 162–169.
Wilk, P. A., Gregorich, K. E., Türler, A., Laue, C. A., Eichler, R., Ninov, V., Adams, J.
L., Kirbach, U. W., Lane, M. R., Lee, D. M., Patin, J. B., Shaughnessy, D. A., Strellis,
D. A., Nitsche, H., and Hoffman, D. C. 2000. Evidence for new isotopes of element 107:
266
Bh and 267 Bh. Phys Rev Lett 85, 2697–2700.
References 465
Zvara, I., Belov, V. Z., Domanov, V. P., Korotkin, Y. S., Chelnokov, L. P., Shalayevsky,
M. R., Shchegolev, V. A., and Hussonois, M. 1972. Chemical isolation of Kurchatovium.
Sov Radiochem 14, 115–118.
Zvara, I., Aikhler, V., Belov, V. Z., Zvarova, T. S., Korotkin, Y. S., Shalayevsky, M. R.,
Shchegolev, V. A., and Hussonois, M. 1974. Gas chromatography and thermochromato-
graphy in the study of transuranium elements. Sov Radiochem 16, 709–715.
Zvara, I., Belov, V. Z., Domanov, V. P., and Shalayevsky, M. R. 1976. Chemical isolation of
nielsbohrium as ekatantalum in the form of the anhydrous bromide; II experiments with
a spontaneously fissioning isotope of nielsbohrium. Sov Radiochem 18, 328–334.
Index
467
468 Index
chain of custody, 78, 186, 222, 298 in Ge detectors, 140, 168, 258, 322
chelation, 52 in LS, 34, 37–38, 97, 100, 105–106, 108–109,
Chemical Hygiene Officer (CHO), 295–296, 117, 127–128, 130, 134, 137, 145,
302–303, 308, 314 151–153, 158, 165–166, 183–184, 255,
Chemical Hygiene Plan (CHP), 295, 297, 308, 327
310, 312–313, 316 in organic scintillators, 34
chemical hygiene, 294–295, 316 in surface barrier detectors, 32, 157, 348, 360
staffing, 295–296 practical counting efficiency, 128, 135, 137,
Cherenkov radiation, 18, 23, 130, 152, 423 190
chromatography, 46, 54, 348, 353, 364, 387, 418 counting
gas, 353, 364, 387 alpha-particle, 124, 126, 150
liquid, 348 beta-particle, 37, 75, 95, 109, 125, 127, 151,
paper, 46 180, 182, 184, 253, 255
clear melt, 423 gamma-ray, 161, 180, 201
Code of Federal Regulations (CFR), 271, 284, gross alpha/beta, 95, 98, 100, 126–127,
286, 293, 295–296, 303–304, 308–310, 182–183, 253
314–316 liquid scintillation (see liquid scintillation)
explained, 314–316 statistics, 37, 198–199
coincidence, 34, 141, 145 counting efficiency, 13, 34, 36, 37, 103, 106,
combustible liquids, 303–304, 423 110
command and control, 320–321 critical value, 189, 204
calculation of, 205
complex-forming, 48
and MDA, 204
Comprehensive Nuclear Test Ban Treaty
cross-section, 45, 47, 373
(CNTBT), 329
Curie, Marie, 338
compton scattering, 23–25, 28, 140, 146, 160,
161, 424
Data Quality Objectives (DQOs), 238
contaminants, 39–40, 43–44, 48, 50–51, 57, 105,
108, 121, 128, 132, 145, 164, 171, 203, data
210, 224, 230, 258, 270, 273, 302, 312, 332 assessment, 222, 236
control chart, 208–211, 231–232 authentication, 321–322, 430
conversion electrons, 142 presentation of, 6, 190, 216–219
emission of, 163, 166 review, 190, 213–216
co-precipitation, 3, 42, 66-68, 74, 109, 110, 112, security, 279, 281–283
113 surety, 321–322
correction validation, 241, 273
energy, 15, 140 verification, 220
mass, 133 Daubert v. Merrill Dow, 243
cosmic rays, 1, 14, 103, 144 dead time (see also resolving time)
cost decay constant, 12, 176, 178, 191, 209,
laboratory, 288 425
of analysis, 99, 222, 290 decay fraction, 116, 137, 142, 162, 166–168,
counters (see also detectors) 175, 179, 190, 425
Geiger-Mueller, 29, 139, 148, 427 decay schemes
ionization chambers, 29–30, 148, 162 positron emission and electron capture, 181
maintenance alpha particle and gamma ray, 181
proportional, 97, 100, 105–106, 109, 116, beta particle, 179
117, 128, 134, 137, 139–140, 145, 148, 151 beta particle and electron capture, 179–180
and source thickness, 123–124 beta particle and gamma ray, 180
calculation of, 123–126 conventions, 172
for alpha particles, 124, 126, 150 decommissioning, 94, 220–221, 286
for beta particles, 37, 75, 95, 109, 125, 127, decontamination factor (DF), 39, 95, 102, 171
151, 180, 182, 184, 253, 255 defensibility of results, 242–243
for gamma rays, 131–133 depletion layer, 33
Index 469
ingrowth, 3, 14, 95, 98, 106, 110 cocktail, 34, 75, 97, 127, 129, 130, 152–156,
integrity 183, 336
in business practices, 286–287 counter, 18, 23, 104, 151–156, 276
interference, 28 system, 256, 350
isobaric, 390–391 Love Canal, 239–240, 267
radionuclide, 93, 95, 142 low-concentration behavior, 66–68
interlaboratory testing, 241, 428 Lucas cells, 98, 110, 84, 199, 429
internal conversion, 10, 422, 424, 429, 436 luminescence, 33, 127, 153, 156, 255
International Union of Pure and Applied
Chemistry (IUPAC), 341–344 maintenance
intralaboratory testing (see quality control) equipment, 280–281
iodine, 33, 37, 41–42, 83, 97, 101, 108, 128, mass defect, 429
213, 400, 415–416 mass excess, 15
ion cyclotron resonance (ICR), 377 mass spectrometer
ion trap mass spectrometer (ITMS), 376–377 accelerator, 364, 398–401
ion-exchange applications, 406–409
resins, 44–48, 99, 185, 327 ion detection, 373, 379–382, 386
separation, 44–49 ion source, 365–369
ionization, 18–19, 28, 30, 32, 36–37, 84, 105, mass analyzer, 370–379
137, 147–148, 162, 365–369, 388–389, 393 vacuum chamber, 382
chamber, 29–30, 147–148, 162, 427 mass spectrometry
isobar, 9, 362, 391, 429 of actinides, 338–361
isotope, 1, 3, 7–11, 73 of activation products, 409
isotope dilution, 3, 72–73, 364, 385 of fission products, 408–409
isotope enrichment, 65 of uranium, 407–408
isotope separation, 39, 65, 113, 363 method
isotopic exchange, 74, 253 instability of, 245
IVO-COLD, 347, 357 modification of, 119–120
development of, 118–120
laboratory Minimum Detectable Activity (MDA), 287
license, 285 molecular sieves, 46, 60
notebooks, 228–229, 243 monitoring, 2, 18, 83, 91, 93–94, 147, 182, 187,
access, 224, 269, 281 220–221, 240, 261–263, 306–308
airflow in, 269–271 Monte Carlo
design, 261–291 evaluation, 355–356
discharge, 129, 268–272, 314, 336 programs, 248, 356
flooring, 269 simulation, 125–126, 133, 135–136, 255, 257
hazard levels, 262 multi-channel analyzer, 276
management, 261–291
operating procedures, 222, 224–225 NaI(Tl) detector, 34, 37, 132, 160, 183
operation, 6, 119, 222, 224, 236, 245–246, 293 negatrons
safety, 292–317 emission of, 9, 422
staffing, 78 neutrino, 9, 145
laminar flow boxes, 302, 429 neutron activation, 16, 82, 84, 103–104, 108,
laser ablation, 365–369, 378, 387, 403–404 111, 114, 430
legislation neutrons
Atomic Energy Act, 284, 421 epithermal, 17
RCRA, 315–316 fast, 17, 430
regulatory, 217, 219, 295, 421 high-energy, 17
license resonance energy, 17
radioactive materials, 284–285 sources, 108
liquid-liquid extraction, 50–54, 349–350 thermal, 16, 430
liquid scintillation (see also scintillation and normal distribution, 19, 196, 206
scintillation detectors) nuclear reaction, 431
Index 471