• No results found

Uranium Aerosol Activity Size Distributions at a Nuclear Fuel Fabrication Plant

N/A
N/A
Protected

Academic year: 2021

Share "Uranium Aerosol Activity Size Distributions at a Nuclear Fuel Fabrication Plant"

Copied!
15
0
0

Loading.... (view fulltext now)

Full text

(1)

Downloaded from http://journals.lww.com/health-physics by BhDMf5ePHKav1zEoum1tQfN4a+kJLhEZgbsIHo4XMi0hCywCX1AWnYQp/IlQrHD3bhnalqTQiPuXw799DX1z331J56NmOmbCKbDOAkAjv6Ro1LiYjfk+HQ== on 10/01/2020 Downloadedfrom http://journals.lww.com/health-physicsby BhDMf5ePHKav1zEoum1tQfN4a+kJLhEZgbsIHo4XMi0hCywCX1AWnYQp/IlQrHD3bhnalqTQiPuXw799DX1z331J56NmOmbCKbDOAkAjv6Ro1LiYjfk+HQ==on 10/01/2020

Paper

Uranium Aerosol Activity Size Distributions at a

Nuclear Fuel Fabrication Plant

Edvin Hansson,

1,2

Håkan B.L. Pettersson,

3

and Mats Eriksson

1,4

Abstract—Inhalation of uranium aerosols is a concern in nuclear fuel fabrication. Determination of committed effective doses and lung equivalent doses following inhalation intake requires knowl-edge about aerosol characteristics; e.g., the activity median aerody-namic diameter (AMAD). Cascade impactor sampling of uranium aerosols in the breathing zone of nuclear operators was carried out at a nuclear fuel fabrication plant producing uranium dioxide via ammonium uranyl carbonate. Complementary static sampling was carried out at key process steps. Uranium on impaction sub-strates was measured using gross alpha counting and alpha spectrometry. Activity size distributions were evaluated for both unimodal and bimodal distributions. When a unimodal distribu-tion was assumed, the average AMAD in the operator breathing zone at the workshops was 12.9–19.3 mm, which is larger than found in previous studies. Certain sampling occasions showed variable isotope ratios (234U/238U) at different impactor stages, indicating more than one population of particles; i.e., a multi-modal activity size distribution. When a bimulti-modal distribution (coarse and fine fraction) was assumed, 75–88% of the activity was associated with an AMAD of 15.2–18.9 mm (coarse frac-tion). Quantification of the AMAD of the fine fraction was asso-ciated with large uncertainties. Values of 1.7–7.1 mm were obtained. Static sampling at key process steps in the workshops showed AMADs of 4.9–17.2 mm, generally lower than obtained by breathing zone sampling, when a unimodal distribution was assumed. When a bimodal distribution was assumed, a smaller fraction of the activity was associated with the coarse fraction compared to breathing zone sampling. This might be due to

impactor positioning during sampling and sedimentation of large particles. The average committed effective dose coefficient for breathing zone sampling and a bimodal distribution was 1.6–2.6 mSv Bq−1for234

U when Type M/S absorption parameters were as-sumed (5.0mSv Bq−1for an AMAD of 5mm). The corresponding lung equivalent dose coefficient was 3.6–10.7 mSv Bq−1(29.9mSv Bq−1 for an AMAD of 5 mm). The predicted urinary excretion level 100 d after inhalation intake was found to be 13-34% of that corre-sponding to an AMAD of 5mm. Uranium aerosols generated at a nuclear fuel fabrication plant using ammonium uranyl carbonate route of conversion were associated with larger AMADs compared to previous work, especially when sampling of aerosols was carried out in the operator breathing zone. A bimodal activity size distribu-tion can be used in calculadistribu-tions of committed effective doses and lung equivalent doses, but parameters associated with the fine frac-tion must be interpreted with care due to large uncertainties. Health Phys. 119(3):327–341; 2020

Key words: aerosols; contamination, internal; lungs, human; nuclear workers

INTRODUCTION

IN NUCLEARfuel fabrication, detailed knowledge of the

inha-lation exposure to uranium aerosols and the corresponding committed effective doses (CED) needs to be gained in or-der to ensure worker safety and regulatory compliance. Re-cent work has found evidence for increased lung cancer risk among uranium and plutonium workers (median, mean, and max lung absorbed doses of 2.4, 7.3, and 316 mGy, respec-tively) following inhalation intake (Grellier et al. 2017). It is thus important to ensure low intake levels and to accurately determine CED following inhalation exposure, which may be challenging due to confounding factors and uncertainties (ICRP 2007; Gilbert 2009; Harrison and Day 2008).

To accurately determine the CED following inhalation intake of uranium aerosols, the deposition of inhaled parti-cles in the airways must be well understood. Particle depo-sition is commonly modeled using the Human Respiratory Tract Model (HRTM) designed by the International Commis-sion on Radiological Protection (ICRP) (ICRP 1994, 2015). The model assumes a population of particles following a

1Department of Medical and Health Sciences, Linköping University,

58183 Linköping, Sweden;2Westinghouse Electric Sweden AB, Bränslegatan 1, 72136 Västerås, Sweden;3Department of Radiation Physics, and

Department of Medical and Health Sciences, Linköping University, 58183 Linköping, Sweden;4Swedish Radiation Safety Authority, 17116 Stockholm, Sweden.

The authors declare no conflicts of interest.

For correspondence contact: Edvin Hansson, Department for Nuclear Safety, Westinghouse Electric Sweden AB, 72136 Västerås, Sweden, or email at hanssoea@westinghouse.com.

(Manuscript accepted 9 December 2019) 0017-9078/20/0

Copyright © 2020 The Author(s). Published by Wolters Kluwer Health, Inc. on behalf of the Health Physics Society. This is an open-access article distributed under the terms of the Creative Commons Attribution-Non Commercial-No Derivatives License 4.0 (CCBY-NC-ND), where it is permissible to download and share the work provided it is properly cited. The work cannot be changed in any way or used commercially without permission from the journal.

(2)

log-normal distribution, which can be described by the Activity Median Aerodynamic Diameter (AMAD) and its geometric standard deviation (GSD). The AMAD is the aerodynamic particle size where 50% of the total activity is associated with particles with an aerodynamic diameter

greater than the AMAD (d50%). The GSD describes the

spread of the activity size distribution. The frequency function is defined (eqn 1) as

dA¼ ffiffiffiffiffiffi 1 2p p lnGSD ð Þexp − lndae− lnAMAD ð Þ2 2 lnGSDð Þ2 " # d ln dae; ð1Þ

where dA is the fraction of radioactivity in the size range

be-tween ln daeand ln dae+ d ln dae, where daeis the

aerody-namic particle diameter (Hinds 1999). The aerodyaerody-namic particle diameter of a non-spherical particle is defined as

dae = de[r(r0χ)1]0.5, where de is the diameter of a

spherical particle with the same volume as the particle

con-sidered,r (g cm−3) is the density of the non-spherical

parti-cle, r0 is the reference density (1 g cm−3), and χ the

dynamic shape factor (dimensionless, typically assumed to be 1.5) (Hinds 1999; ICRP 1994, 1997).

A GSD value of 1 describes a monodispersed aerosol, and a large GSD indicates a large spread in the activity size distribution. The equation is applicable for log-normal dis-tributions only, which is typically assumed in the field of ra-diation protection. In the absence of detailed information,

the ICRP recommends a default value of AMAD = 5mm

and GSD = 2.5 (ICRP 1994). These parameters indicate that 95% of the activity is associated with particles in the

aerody-namic size range 0.8mm (5 mm2.52= 0.8mm) and 31

mm (5 mm2.52≈ 31 mm) (Hinds 1999). The activity size

distribution of aerosols is important for modeling deposition in the airways using the HRTM model; e.g., the alveolar de-position is expected to be 3% of the total inhaled activity if

the AMAD is 15mm, whereas the corresponding figure for

an AMAD of 5mm is 10%, thus increasing the lung

equiv-alent dose (Hlung) (ICRP 2015).

The AMAD and GSD can be determined experimentally by sampling aerosols using cascade impactors. A common evaluation method is to plot the accumulated fraction of activ-ity as a function of impactor cut-point in a log-probabilactiv-ity graph. This method may give too much weight to early and late impaction stages (Hinds 1999). Alternatively, a nonlinear regression fit to the frequency function (eqn 1) can be applied to determine the AMAD and GSD (Thiel 2002).

The aerosol size distribution will affect the deposition pattern in the respiratory tract and, as a consequence, the bio-logical fate of inhaled radioactive aerosols. Thus, the rate of excretion (e.g., via urine) and CED per inhaled unit of activity will vary with particle size (AMAD and GSD) in addition to the chemical composition/solubility class affecting particle dissolution rate in the respiratory tract.

Early studies have reported cascade impactor measure-ments of uranium aerosol AMADs in nuclear fuel

fabrica-tion. Schieferdecker et al. reported an AMAD of 8.2mm

(Schieferdecker et al. 1985). Thind reported an AMAD of

6.1mm and GSD of 2.1 (Thind 1986). A more recent review

paper stated an average AMAD of 6.2mm with a GSD of

2.3 (Davesne and Blanchardon 2014). An earlier review paper

stated that results below 1mm were inconsistent, suggesting a

bimodal size distribution (Dorrian and Bailey 1995). Avail-able data are based on static sampling and might not be repre-sentative of the breathing zone of operators (Davesne and Blanchardon 2014). To the best of our knowledge, AMADs determined from cascade impactor sampling in the breathing zone of workers in nuclear fuel fabrication have not been pub-lished by open sources.

Uranium dioxide (UO2) powder for nuclear fuel and

light-water reactors is typically produced from uranium

hexafluoride (UF6) using integrated dry route (IDR)

con-version or wet-route concon-version via ammonium diuranate (ADU), alternatively via ammonium uranyl carbonate (AUC)

(IAEA 2006). These production methods yield UO2powder

with different properties; e.g., the AUC route of conversion

produces a UO2powder consisting of coarser particles

com-pared to the other methods (IAEA 2006). From a radiation protection perspective, process history is important since it af-fects the physicochemical properties of aerosols associated with a certain compound (Ansoborlo et al. 1999). From pub-lished data on uranium aerosols in nuclear fuel fabrication, it can be difficult to distinguish which production method was used. To the best of our knowledge, only one study has cov-ered activity size distributions at a site using wet-route conversion via AUC and reported an average AMAD of

8.2mm (Schieferdecker et al. 1985).

The aim of the present work is to evaluate uranium aerosol activity size distributions (with corresponding AMADs and GSDs) in the breathing zone of operators at a nuclear fuel fabrication plant using wet-route conversion via AUC. Portable cascade impactors were used to sample uranium aerosols in the operator breathing zone. In addition, static sampling was carried out at certain process steps. Activity size distributions were eval-uated by assuming unimodal as well as bimodal distributions, and the impact on dosimetry calculations was evaluated based on the obtained results. The information in the present work is important in order to improve internal dosimetry, worker safety, and ensure regulatory compliance. Improved internal dosimetry might add to the overall knowledge about health effects as-sociated with low-level exposure to ionizing radiation.

MATERIALS AND METHODS Impactor sampling

Cascade impactors (Marple 298, Prod. No. SE298; Thermo Fisher Scientific, Waltham, MA), which operate

(3)

at 2.0 L min−1were used for aerosol sampling. The impactor

used in the present work has eight impaction stages (A–H)

with corresponding cut-points of 21.3, 14.8, 9.8, 6.0, 3.5,

1.6, 0.9, and 0.5mm, respectively. The cut-point is defined

as the aerodynamic diameter of particles having 50% prob-ability of impaction at the given stage. The last stage is followed by a final filter to collect remaining particles. Gilian 5000 sampling pumps (Sensidyne, St. Petersburg, FL) were used with the impactors. The pumps were calibrated and flowrates checked using the methodology described in previous work, with flow rates showing variations typically <1% (Hansson et al. 2017). The small size of the impactor al-lows for sampling in the breathing zone, which was carried out by personnel carrying the impactor attached to the collar of the overalls throughout the work day. Static sampling was carried out by positioning the impactor as close as practically possi-ble to work stations of particular interest.

Mixed cellulose ester membrane (MCE) (Thermo Sci-entific, SEC-290-MCE) was used as impaction substrate,

and polyvinyl chloride (PVC) (SEF-290-P5) with a 5mm

pore size was used as final collection filter unless otherwise stated. Other authors have used an identical setup (Cheng et al. 2009). In the present work, impactor substrates were not coated to prevent particle bounce due to difficulties in reproducing application of thin, even layers of coating material. Furthermore, coating of substrates would distort comparison with planned in vitro dissolution rate experiments using the same sampling methodology. However, an attempt to investigate the presence of particle bounce was carried out by parallel sampling using cascade impactors with coated and uncoated substrates. The coating material (PRF Industrial Line Silicon Oil, Taerosol Oy, Kangasala, Finland) was sprayed onto each MCE substrate in a sweeping motion for 1-2 s at a distance of approximately 15 cm. Parallel sampling was carried out twice (three impactors per occasion) at a work station where scrap pellets are oxidized to triuranium

octoxide (U3O8). Impaction stages G and H, as well as the

final collection filters, were examined using scanning electron microscopy with energy-dispersive x-ray spectrometry (Phenom ProX, Phenom-World BV, Eindhoven, The Netherlands).

Sampling was carried out in the operator breathing zone at the four major workshops at the site: conversion, powder preparation, pelletizing, and burnable absorber (BA)

pelletizing. At the conversion workshop, UO2is formed from

UF6via AUC. At the powder preparation workshop, powder

milling, powder blending, and oxidizing of waste materials is

carried out. The pelletizing workshop produces UO2pellets

by pressing, sintering, and grinding. The BA pelletizing

work-shop is similar but uses milled UO2powder, which is blended

with Gd2O3. An overview of the main uranium flows is

pre-sented in Fig. 1. A more detailed description of the processes and typical operations has been described in previous work (Hansson et al. 2017).

Uranium material at the site is handled in batches,

where the enrichment levels (mass-percent235U) vary

be-tween 0.71% and 4.95% (average enrichment level 3.8%).

The activity ratio234U/238U increases approximately

line-arly with enrichment level. As a consequence, enrichment levels at the workshops may vary from day to day.

Sampling of the breathing zone was carried out on 18 occasions (five occasions at each workshop except for the BA pelletizing workshop, where only three occasions could be completed due to production-related circumstances). In addition to the breathing zone sampling, static sampling was carried out at 13 key locations at the different workshops: • Conversion workshop: Static sampling at the transporta-tion of AUC to the fluidizing bed furnaces and of the gen-eral workshop air;

• Powder preparation workshop: Static sampling at the sta-tions for powder milling station, oxidizing of discarded pellets, oxidizing of waste from pellet grinding, humidity

check (for criticality safety reasons) of UO2powder and

of the general workshop air;

• Pelletizing workshop: Static sampling at one of the pellet pressing stations, emptying of sintered pellets and of the general workshop air; and

• BA pelletizing workshop: Static sampling at the stations

for UO2/Gd2O3 blending and oxidizing of waste from

pellet grinding.

Sampling times were based on previous knowledge of airborne activity concentration levels and determined so as to collect sufficient amounts of radioactivity for reliable measurements but avoiding particle overload on the impaction substrates.

Radioactivity measurements

The total amount of alpha activity at all impactor

sub-strates and final collection filters were measured for 20–24 h

using an LB 790 10-Channel a-b Low-Level Counter

Fig. 1. A schematic overview of the main flows of different uranium materials between the four main workshops.

(4)

Ta b le 1 . Uranium acti v ity concentrations and acti v ity ratio s (br eat hing zone samp ling) b ase d o n radio metr ic da ta fo llo wing al ph a spe ctro metr y. Ura n ium ac ti vity cor re spon ds to th e sum of 23 4 U, 23 5 U, an d 23 8 U acti v ity . U ncer tainties cor re sp ond to ± 1 st andard de viation d ue to counting statistics and flo w rate measurements (2.00 ± 0.05 L m in − 1 ). Sa mpli ng-ID Ura n iu m acti v ity co n cen tra tio n (B q m – 3) and Acti vity ratio 234 U/ 238 U Sampling time (min) To ta l concentration (Bq m – 3) Stage A 21. 3 m m Sta g e B 14 .8 m m Stage C 9.8 m m Stage D 6. 0 m m St age E 3. 5 m m Sta g e F 1.6 m m Stage G 0. 9 m m Stage H 0.5 m m Fi n al filt er Con ver sion wo rk shop Sam p li ng 1 1 .2 9 (0.04 ) 1 .03 (0. 03) 1. 62 (0 .0 4) 0 .77 (0 .03 ) 0.1 9 8 (0.0 06) 0 .09 9 (0 .00 3) 0. 090 (0 .00 3 ) 0 .03 3 (0 .00 3 ) 0 .24 9 (0 .00 7 ) 3 68 5.4 (0 .7) 7. 0 (0.2) 6.9 (0. 3) 6 .9 (0.2 ) 6.9 (0 .3) 6 .2 (0.2 ) 6. 5 (0 .2) 6.5 (0 .3) 6. 3 (0 .6) 6. 8 (0 .1) Sam p li ng 2 0 .13 2 (0 .00 7 ) 0 .28 (0. 01) 0.1 5 9 (0.0 06) 0. 076 (0 .00 5 ) 0 .0 19 (0 .0 02) 0 .02 6 (0 .00 2) 0. 013 (0 .00 1 ) 0 .00 5 (0 .00 1 ) 0 .15 1 (0 .00 6 ) 6 69 0.9 (0 .2) 5. 9 (0.3) 6.6 (0. 2) 6 .5 (0.2 ) 5.7 (0 .3) 5 .9 (0.4 ) 6. 1 (0 .5) 6.4 (0 .3) 5. 2 (0 .4) 5. 7 (0 .2) Sam p li ng 3 0 .19 2 (0 .00 8 ) 0 .159 (0. 006 ) 0 .1 60 (0 .0 07) 0. 099 (0 .00 5 ) 0 .0 37 (0 .0 02) 0 .02 4 (0 .00 1) 0. 015 (0 .00 1 ) 0 .00 9 (0 .00 1 ) 0 .05 2 (0 .00 3 ) 6 69 0.8 (0 .1 ) 5. 9 (0.3) 6.2 (0. 2) 6 .3 (0.3 ) 6.3 (0 .3) 6 .4 (0.4 ) 6. 9 (0 .4) 6.2 (0 .4) 5. 8 (0 .4) 5. 7 (0 .3) Sam p li ng 4 0 .4 6 (0.02 ) 0 .64 (0. 02) 0. 75 (0 .0 3) 0 .46 (0 .01 ) 0.1 5 5 (0.0 08) 0 .07 8 (0 .00 4) 0. 042 (0 .00 3 ) 0 .02 9 (0 .00 2 ) 0 .14 7 (0 .00 7 ) 4 45 2.8 (0 .4) 6. 7 (0.2) 6.5 (0. 2) 6 .3 (0.3 ) 6.3 (0 .2) 5 .8 (0.3 ) 6. 4 (0 .3) 5.9 (0 .4) 6. 0 (0 .3) 6. 2 (0 .3) Sam p li ng 5 1 .1 7 (0.04 ) 0 .99 (0. 03) 0. 59 (0 .0 2) 0 .32 (0 .01 ) 0.1 0 7 (0.0 06) 0 .04 7 (0 .00 3) 0. 027 (0 .00 3 ) 0 .01 8 (0 .00 2 ) 0 .11 8 (0 .00 5 ) 3 00 3.4 (0 .6) 6. 7 (0.2) 6.3 (0. 1) 5 .6 (0.1 ) 4.8 (0 .1) 3 .4 (0.2 ) 3. 5 (0 .2) 4.6 (0 .4) 3. 7 (0 .2) 5. 3 (0 .2) P o wde r pr ep a rati o n w or ks ho p Sam p li ng 1 0 .5 7 (0.02 ) 1 .11 (0. 03) 1. 42 (0 .0 6) 1 .05 (0 .03 ) 0. 47 (0 .0 2) 0.3 4 (0 .01 ) 0. 128 (0 .00 9 ) 0 .11 7 (0 .00 6 ) 0 .14 5 (0 .00 4 ) 8 14 5.4 (0 .7) 6. 6 (0.2) 6.5 (0. 1) 6 .7 (0.5 ) 6.2 (0 .1) 5 .3 (0.2 ) 3. 6 (0 .1) 3.3 (0 .1) 4. 1 (0 .2) 5. 6 (0 .2) Sam p li ng 2 0 .7 6 (0.02 ) 0 .65 (0. 02) 1. 81 (0 .0 5) 1 .41 (0 .04 ) 0. 91 (0 .0 3) 0.9 9 (0 .03 ) 0. 233 (0 .00 7 ) 0 .11 1 (0 .00 4 ) 0 .5 6 (0 .02 ) 8 93 7.4 (0 .7) 5. 6 (0.2) 5.6 (0. 2) 5 .6 (0.1 ) 5.7 (0 .2) 5 .6 (0.2 ) 5. 5 (0 .2) 5.4 (0 .1) 5. 5 (0 .2) 5. 7 (0 .2) Sam p li ng 3 0 .3 5 (0.02 ) 0 .77 (0. 03) 1. 05 (0 .0 3) 0 .78 (0 .03 ) 0. 38 (0 .0 2) 0 .22 1 (0 .00 7) 0. 107 (0 .00 4 ) 0 .08 9 (0 .00 5 ) 0 .14 7 (0 .00 6 ) 8 02 3.9 (0 .5) 5. 3 (0.3) 5.4 (0. 2) 5 .5 (0.2 ) 5.5 (0 .2) 5 .9 (0.3 ) 5. 6 (0 .1) 5.7 (0 .2) 5. 6 (0 .3) 5. 7 (0 .2) Sam p li ng 4 2 .2 5 (0.07 ) 2 .52 (0. 08) 4 .5 (0.2 ) 4.3 (0 .1) 9 .4 (0.3 ) 9. 8 (0 .3) 1 .11 (0 .04 ) 0.3 6 (0 .01 ) 0.3 5 (0 .02 ) 5 6 2 3 5 (4 ) 6. 5 (0.3) 6.3 (0. 3) 6 .6 (0.5 ) 6.8 (0 .2) 6 .9 (0.8 ) 7. 1 (0 .3) 7.0 (0 .4) 7. 1 (0 .3) 6. 3 (0 .4) Sam p li ng 5 1 .3 4 (0.04 ) 0 .92 (0. 03) 1. 87 (0 .0 5) 1 .18 (0 .04 ) 0. 53 (0 .0 2) 0.4 2 (0 .02 ) 0. 202 (0 .00 8 ) 0 .12 6 (0 .00 7 ) 0 .2 9 (0 .01 ) 5 93 6.9 (0 .8) 6. 6 (0.2) 6.2 (0. 2) 6 .7 (0.2 ) 6.6 (0 .3) 7 .0 (0.2 ) 6. 7 (0 .4) 6.4 (0 .3) 6. 9 (0 .5) 6. 5 (0 .3) P elletiz ing w or kshop Sam p li ng 1 0 .02 0 (0 .00 1 ) 0 .022 (0. 001 ) 0 .0 29 (0 .0 02) 0. 021 (0 .00 1 ) 0 .0 09 (0 .0 01) 0 .00 5 (0 .00 1) 0.0 022 (0 .00 03) 0. 002 0 (0 .00 03) 0 .00 8 (0 .00 1) 1 149 0. 12 (0 .04 ) 6. 1 (0.3) 5.9 (0. 3) 6 .2 (0.4 ) 5.9 (0 .3) 5 .8 (0.6 ) 5. 5 (0 .6) 5.6 (0 .7) 6. 1 (0 .7) 6. 2 (0 .5) Sam p li ng 2 0 .13 8 (0 .00 4 ) 0 .247 (0. 007 ) 0 .2 81 (0 .0 1) 0. 175 (0 .00 5 ) 0 .0 81 (0 .0 02) 0 .06 2 (0 .00 2) 0. 041 (0 .00 2 ) 0 .01 8 (0 .00 1 ) 0 .11 3 (0 .00 4 ) 1 082 1.2 (0 .1 ) 5. 8 (0.1) 6.1 (0. 1) 6 .4 (0.3 ) 5.8 (0 .1) 6 .0 (0.2 ) 6. 0 (0 .2) 6.5 (0 .2) 6. 0 (0 .3) 6. 2 (0 .2) Sam p li ng 3 0 .13 0 (0 .00 4 ) 0 .113 (0. 004 ) 0 .1 57 (0 .0 05) 0. 116 (0 .00 4 ) 0 .0 67 (0 .0 03) 0 .04 8 (0 .00 2) 0. 028 (0 .00 1 ) 0 .01 3 (0 .00 2 ) 0 .02 1 (0 .00 1 ) 1 827 0.7 (0 .1) 3. 9 (0.1) 4.1 (0. 1) 3 .9 (0.1 ) 4.1 (0 .1) 4 .2 (0.1 ) 4. 2 (0 .1) 4.5 (0 .2) 4. 8 (0 .4) 4. 6 (0 .2) Sam p li ng 4 0 .24 3 (0 .00 7 ) 0 .210 (0. 006 ) 0 .2 82 (0 .0 09) 0 .19 (0 .01 ) 0.0 7 8 (0.0 03) 0 .03 8 (0 .00 2) 0. 021 (0 .00 2 ) 0 .01 3 (0 .00 1 ) 0 .04 9 (0 .00 2 ) 1 292 1.1 (0 .2) 6. 4 (0.2) 6.2 (0. 2) 6 .1 (0.2 ) 6.9 (0 .5) 6 .0 (0.2 ) 6. 4 (0 .4) 5.8 (0 .4) 5. 7 (0 .6) 6. 8 (0 .4) Sam p li ng 5 0 .09 0 (0 .00 3 ) 0 .184 (0. 007 ) 0 .3 1 (0.0 1 ) 0 .214 (0 .00 6 ) 0 .0 84 (0 .0 03) 0 .06 2 (0 .00 3) 0. 042 (0 .00 2 ) 0 .01 7 (0 .00 1 ) 0 .11 6 (0 .00 4 ) 2 469 1.1 (0 .1) 7. 1 (0.3) 7.1 (0. 3) 7 .0 (0.3 ) 7.1 (0 .2) 7 .2 (0.3 ) 7. 2 (0 .5) 7.4 (0 .3) 6. 9 (0 .3) 6. 9 (0 .3) Continued n ext p a g e

(5)

(Berthold Technologies, USA, LLC, Oak Ridge, TN). Back-ground levels were corrected for and frequently evaluated.

The breathing zone samples were also analyzed using alpha spectrometry in order to obtain uranium isotopic data.

Samples were acid dissolved in concentrated HNO3

followed by aqua regia after addition of traceable amounts

of232U yield determinant (Isotrak, AEA Technology, PLC,

Didcot, UK). Uranium was then extracted from 8 M

HNO3 by tributyl phosphate (TBP) and back-extracted

into demineralized water (Holm 1984) and finally electro-deposited onto stainless steel planchets (Hallstadius 1984). Measurements were carried out using passivated implanted planar silicon detectors (Canberra PIPS, Mirion Technologies Inc, San Ramon, CA), ORTEC (Oak Ridge,

TN) OctêteRMCA system, and ORTEC MaestroRsoftware.

Counting times varied between 1–4 d.

The activity ratio from measurements by alpha spec-trometry compared to total alpha activity followed a normal distribution with a mean of 1.01 and a standard deviation of 0.21. Two data points with very low activities deviated from the mean with more than 3 standard deviations and were considered outliers, resulting in a mean ratio of 0.99 with a standard deviation of 0.13. The agreement was considered satisfactory, and thus the remaining samples (static sam-pling) were not analyzed using alpha spectrometry. Data reduction

Activity size distributions were evaluated using a unimodal approach (eqn 1). However, alpha spectrometry

measure-ments revealed variable isotope ratios (234U/238U) during

some sampling occasions, suggesting a multi-modal activity size distribution. Thus a bimodal activity size distribution was evaluated for each sampling occasion (eqn 2). A data

re-duction protocol developed by O’Shaughnessy and Raabe

for sampling carried out with the Marple 298 impactor was

used to evaluate activity size distributions (O’Shaughnessy

and Raabe 2003). The protocol, which assumes a unimodal log-normal distribution (eqn 1), was modified as to model a bimodal activity size distribution:

dA¼ ffiffiffiffiffiffi f 2p p lnGSD1 ð Þexp − lndae− lnAMAD1 ð Þ2 2 lnGSD1ð Þ2 " # ( þ ffiffiffiffiffiffi1−f 2p p lnGSD2 ð Þexp − lndae− lnAMAD2 ð Þ2 2 lnGSD2ð Þ2 " #

g

d ln dae; ð2Þ

where f and 1−f is the fraction of activity in the coarse and

fine fraction, respectively. The same approach has been used by other authors investigating bimodal activity size distributions (Cheng et al. 2009).

The measured amount of radioactivity at each impac-tion substrate was corrected for sampling efficiency by ap-plying a stage impaction efficiency of 0.52, 0.61, 0.78, 0.89, 0.95, 0.96, 0.97, and 0.99 for impactor stages A-H,

B A pe lle tiz in g w or ks ho p Sam p li ng 1 0 .05 2 (0 .00 2 ) 0 .049 (0. 002 ) 0 .0 80 (0 .0 04) 0. 066 (0 .00 3 ) 0 .0 32 (0 .0 02) 0 .02 1 (0 .00 2) 0. 009 (0 .00 1 ) 0 .00 8 (0 .00 1 ) 0 .03 2 (0 .00 2 ) 5 84 0.35 (0 .08 ) 5. 7 (0.2) 5.1 (0. 2) 5 .9 (0.2 ) 5.6 (0 .2) 5 .1 (0.3 ) 5. 4 (0 .4) 5.9 (0 .6) 4. 8 (0 .4) 5. 9 (0 .3) Sam p li ng 2 0 .01 5 (0 .00 1 ) 0 .023 (0. 001 ) 0 .0 33 (0 .0 02) 0. 023 (0 .00 2 ) 0 .0 09 (0 .0 01) 0 .00 6 (0 .00 1) 0. 003 (0 .00 1 ) 0 .00 3 (0 .00 1 ) 0 .01 2 (0 .00 1 ) 5 99 0.13 (0 .08 ) 4. 4 (0.3) 4.6 (0. 3) 4 .5 (0.3 ) 4.8 (0 .3) 4 .5 (0.5 ) 5. 4 (0 .8) 4.1 (0 .8) 3. 8 (0 .7) 4. 5 (0 .2) Sam p li ng 3 0 .06 5 (0 .00 3 ) 0 .067 (0. 002 ) 0 .1 19 (0 .0 05) 0. 092 (0 .00 3 ) 0 .0 63 (0 .0 02) 0 .04 4 (0 .00 3) 0. 018 (0 .00 1 ) 0 .01 2 (0 .00 1 ) 0 .02 3 (0 .00 1 ) 9 36 0.5 (0. 0 7) 5. 8 (0.2) 5.7 (0. 2) 5 .2 (0.2 ) 5.7 (0 .1) 5 .8 (0.2 ) 5. 8 (0 .3) 5.9 (0 .3) 5. 9 (0 .3) 5. 8 (0 .2)

(6)

respectively, as recommended by the manufacturer (Thermo Fisher 2009). Lower and upper bounds considered in the

data reduction were 0.1mm and 50 mm, respectively, as in

previous studies (Cheng et al. 2009).

The AMADs, GSDs, and f for each sampling occasion were evaluated by applying a nonlinear least squares regres-sion fit (Microsoft Excel 2010 Problem Solver) to the fre-quency function (eqn 1 and 2, respectively). Simulations were repeated in R (version 3.5.1., 2018-07-02) and in-cluded standard errors of the derived parameters.

Internal dosimetry

Dose coefficients (CED per unit of inhalation intake) and predicted urinary excretion following inhalation of

1 Bq of234U were evaluated. The software Integrated

Mod-ules for Bioassay Analysis (IMBA) Professional Plus (v. 4.1) was used (Birchall et al. 2007).

The HRTM has recently been revised by the ICRP; thus, new recommendations are available. For example, the particle transport model has been revised (ICRP 2015). Default

ab-sorption parameters for UO2and U3O8are now considered

in-termediate Type M/S (rapid fraction, fr= 0.03; rapid rate, sr= 1

d−1; slow rate ss= 510−4d−1and fractional absorption in the

alimentary tract, fA= 610−4). Uranyl nitrate [UO2(NO3)2],

uranium peroxide hydrate (UO4), ADU [(NH4)2U2O7] and

uranium trioxide (UO3) are now considered intermediate Type

F/M (fr= 0.8; sr= 1 d−1; ss= 0.01 d−1, and fA= 0.016) (ICRP

2017). AUC [(NH4)4UO2(CO3)3] is not included in ICRP

Publication 137.

An add-on to IMBA has previously been used to model urinary excretion according to the ICRP 130 recommenda-tions (Birchall et al. 2017). This add-on was acquired and used in the present work to model the urinary excretion

fol-lowing inhalation of 1 Bq of234U according to the ICRP

130 particle transport model. Activity size distributions rived in the present work were used in combination with

de-fault absorption parameters, aerosol density (3 g cm−3), and

shape factor (1.5) (ICRP 1994, 2015). For the conversion workshop Type F/M and Type M/S materials were considered. For remaining workshops, only Type M/S material was considered.

To the best of our knowledge, as of today, no software exists that allows for calculation of dose coefficients accord-ing to the ICRP 130 particle transport recommendations. Thus, dose coefficients were evaluated using the ICRP 66 particle transport model but with the revised ICRP 137 ab-sorption parameters.

RESULTS AND DISCUSSION Radioactivity measurements and activity size distributions

Radiometric data following alpha spectrometry of breath-ing zone samples are presented in Table 1, and radiometric data following alpha counting of static sampling are presented

Fig. 2. Box plot of uranium measured in breathing zone samplings for each workshop, indicating median fractions and q1 (25th)/q3 (75th) percen-tiles. X-axes correspond to impactor stage cut-point (final collection filter and stage H-A, respectively). Data points are considered outliers if they are greater than q3 + 1.5 (q3−q1) or less than q1−1.5  (q3−q1) (Matlab R2018a default setting).

(7)

in Appendix A (Table A1). It is evident from Table 1 that

iso-tope ratios (234U/238U) might vary across impactor stages

dur-ing a sdur-ingle round of sampldur-ing (e.g., sampldur-ing 5 at the conversion workshop, sampling 1 at the powder preparation workshop, sampling 3 at the pelletizing workshop). This im-plies that at least two populations of particles with different size distributions and enrichment levels (i.e., originating from different batches of uranium material) might be pres-ent during a sampling occasion. Thus, a bimodal approach might be preferable to describe the data. Many sampling oc-casions showed uniform isotope ratios (e.g., sampling 1 at the pelletizing workshop). This does not contradict a bimodal distribution, as different process steps might be associated with different size distributions but identical enrichment levels (depending on material batch). It is also noteworthy from Table 1 that the concentration of sampled activity varies with up to an order of magnitude. This implies that day-to-day variations in exposure might be large, presumably due to variations in production and work tasks carried out.

Data from Table 1 are presented as fractions in Fig. 2 for easy comparison between the different workshops. Ac-tivity sampled at the conversion workshop was collected at the early impaction stages to a greater extent compared to the other workshops.

Activity size distributions

Each sampling occasion was evaluated by assuming both a unimodal and a bimodal activity size distribution. Results are presented in Table 2 (breathing zone sampling) and Table 3 (static sampling). The average parameters (Table 2) for each workshop were used to visually compare activity size distributions (Fig. 3).

From Tables 2 and 3 and Fig. 3, it is evident that a typ-ical AMAD at the site is larger than the ICRP default

distri-bution (AMAD = 5 mm, GSD = 2.5). Sampling in the

breathing zone of the operators tended to generate more pre-dominant coarse fractions (larger f) compared to static sam-pling. Static sampling can be difficult to carry out in a way

Table 2. Breathing zone sampling—estimated AMAD, GSD and f for coarse and fine fractions assuming unimodal and bimodal activity size distributions. Numbers in brackets indicate the standard error from the curve-fitting procedure unless stated otherwise.

Sampling-ID Unimodal fit

Bimodal fit

Coarse fraction Fine fraction

Conversion workshop AMAD (mm) GSD f AMAD 1 (mm) GSD 1 1-f AMAD 2 (mm) GSD 2

Sampling 1a 17.7 (0.9) 1.8 (0.07) − − − − − Sampling 2b 19.9 (0.5) 1.4 (0.03) Sampling 3 19.8 (0.4) 1.8 (0.04) 0.90 (0.23) 20.1 (0.5) 1.7 (0.1) 0.10 (0.23) 5.3 (17.4) 3.1 (5.2) Sampling 4 16.9 (0.3) 1.7 (0.02) 0.85 (1.12) 17.7 (1.1) 1.6 (0.2) 0.15 (1.12) 8.9 (43.8) 2.1 (3.2) Sampling 5a 22.0 (0.4) 1.5 (0.03) Average (1 SD) 19.3 (1.8) 1.6 (0.2) 0.88 (0.03) 18.9 (1.2) 1.7 (0.1) 0.12 (0.03) 7.1 (1.8) 2.6 (0.5) Powder preparation workshop

Sampling 1 15.2 (0.6) 1.8 (0.05) 0.77 (0.52) 16.1 (0.5) 1.6 (0.2) 0.23 (0.52) 6.5 (14.1) 2.4 (2.1) Sampling 2b 12.0 (1.6) 2.3 (0.2) − − − − − Sampling 3 14.7 (0.6) 1.8 (0.05) 0.73 (0.51) 15.7 (0.6) 1.6 (0.2) 0.27 (0.51) 6.8 (12.3) 2.4 (1.6) Sampling 4c 6.5 (1.0) 2.9 (0.4) 0.56 (0.06) 13.2 (1.1) 2.0 (0.2) 0.44 (0.06) 3.5 (0.1) 1.3 (0.1) Sampling 5 16.1 (1.4) 2.1 (0.1) 0.94 (0.09) 15.9 (1.6) 2.0 (0.2) 0.06 (0.09) 1.6 (1.5) 1.6 (1.8) Average (1 SD) 12.9 (3.5) 2.2 (0.4) 0.75 (0.13) 15.2 (1.2) 1.8 (0.2) 0.25 (0.13) 4.6 (2.2) 1.9 (0.5) Pelletizing workshop Sampling 1 16.4 (0.4) 1.9 (0.04) 0.90 (0.12) 16.5 (0.3) 1.8 (0.1) 0.10 (0.12) 2.7 (5.5) 3.9 (5.6) Sampling 2 16.5 (0.8) 1.7 (0.06) 0.76 (0.19) 17.0 (0.5) 1.6 (0.1) 0.24 (0.19) 5.1 (7.4) 3.4 (2.4) Sampling 3 16.8 (0.8) 2.1 (0.07) 0.90 (0.08) 17.0 (0.9) 1.9 (0.1) 0.10 (0.08) 2.2 (1.4) 2.1 (1.3) Sampling 4 17.5 (0.4) 1.9 (0.04) 0.97 (0.04) 17.5 (0.5) 1.9 (0.1) 0.03 (0.04) 1.8 (1.8) 1.9 (2.3) Sampling 5 14.2 (0.7) 1.8 (0.06) 0.74 (0.10) 14.1 (0.4) 1.6 (0.1) 0.26 (0.10) 4.3 (5.6) 6.0 (3.0) Average (1 SD) 16.3 (1.1) 1.9 (0.1) 0.85 (0.09) 16.4 (1.2) 1.8 (0.1) 0.15 (0.09) 3.2 (1.3) 3.5 (1.5) BA pelletizing workshop Sampling 1 15.1 (0.8) 2.1 (0.08) 0.89 (0.09) 15.0 (0.8) 2.0 (0.1) 0.11 (0.09) 0.9 (1.0) 3.2 (4.6) Sampling 2 15.2 (0.6) 1.8 (0.05) 0.84 (0.12) 15.3 (0.4) 1.7 (0.1) 0.16 (0.12) 2.5 (5.5) 5.4 (6.5) Sampling 3b 14.0 (0.9) 2.2 (0.1) Average (1 SD) 14.8 (0.5) 2.0 (0.2) 0.87 (0.02) 15.2 (0.1) 1.8 (0.1) 0.13 (0.02) 1.7 (0.8) 4.3 (1.1)

aNo solution for bimodal activity size distribution, suggesting weak fine fraction or unimodal distribution. b

Fine and coarse fractions indistinguishable. Derived parameters not included in average parameter values.

(8)

that is representative of worker exposure due to positioning and the fact that operators carry out work at multiple po-sitions during a work period. The distance to the source of particle dispersion is likely important since large particles settle more rapidly than small particles (Hinds 1999). Our results suggest that static sampling might underestimate the amount of activity present in the breathing zone and that the derived AMAD might be underestimated.

A bimodal activity size distribution appears to give a better curve fit compared to a unimodal fit (Fig. 3). This is in accordance with the observed variable isotope ratios for breathing zone sampling (Table 1). However, the AMAD and GSD for the fine fraction were difficult to quantify and were frequently not statistically significant (Tables 2 and 3). This was particularly obvious for the con-version workshop, where the fine fraction appeared to be very indistinct in comparison to the coarse fraction.

One could consider weighting the measurement points in the curve-fitting process (Kemmer and Keller 2010). An attempt was made using relative weighting for each of the samplings carried out at the pelletizing work-shop. For the fine fraction, the mean AMAD increased

from 3.2mm to 5.0 mm and the mean GSD from 3.5 to

9.1, compared to the unweighted curve-fitting procedure. That large a GSD indicates a very broad peak. We believe that this is a case of overfitting data to the model, which

might be misleading, in particular if the late impactor stages are affected by particle bounce. The effect on the coarse fraction was less prominent.

The derived AMADs in the present work are generally considerably higher than previously published data (Davesne and Blanchardon 2014). This might be due to aerosol characteristics associated with the particular production method. The wet-chemical AUC conversion

process is known to generate a coarser UO2powder

com-pared to wet-chemical ADU conversion (Bergqvist and Sahle 2010; Palheiros et al. 2009). The sampling method (i.e., breathing zone vs. static sampling) can also play an im-portant role.

Dose coefficients

The calculated dose coefficients (CED per unit of234U

inhalation intake) for breathing zone sampling at the four workshops are presented in Table 4. The dose coefficient associated with the ICRP default assumption (AMAD = 5 mm, GSD = 2.5) is included for comparison. Potential

expo-sure at the workshops is associated with CED and Hlung

coef-ficients being 21–51% and 12–29%, respectively, of those

obtained for the default 5mm and Type M/S assumption.

Lower dose coefficients are expected since a high AMAD results in a greater fraction of activity depositing in the up-per respiratory tract, resulting in a lower dose to the lung.

Table 3. Static sampling—estimated AMAD, GSD, and f for coarse and fine fractions assuming unimodal and bimodal ac-tivity size distributions. Numbers in brackets indicate the standard error from the curve-fitting procedure.

Sampling position

Unimodal fit

Bimodal fit

Coarse fraction Fine fraction AMAD (mm) GSD f AMAD 1 (mm) GSD 1 1-f AMAD 2 (mm) GSD 2 Conversion workshop

Fluidizing bed furnace 11.7 (1.5) 2.3 (0.2) 0.34 (0.12) 13.8 (0.9) 1.5 (0.1) 0.66 (0.12) 6.8 (2.2) 3.5 (0.6) General ventilation exhaust 10.8 (1.0) 2.2 (0.2) 0.33 (0.07) 11.4 (0.4) 1.4 (0.1) 0.67 (0.07) 8.9 (1.7) 3.5 (0.6) Powder preparation workshop

Powder milling 4.9 (1.0) 3.5 (0.6) 0.55 (0.05) 10.0 (0.7) 1.9 (0.1) 0.45 (0.05) 1.8 (0.1) 1.6 (0.1) Humidity check 13.5 (1.1) 2.7 (0.2) 0.87 (0.10) 14.5 (1.5) 2.2 (0.3) 0.13 (0.10) 2.3 (0.9) 1.6 (0.5) Oxidizing of discarded pelletsa 12.4 (0.6) 1.8 (0.1) 0.50 (0.09) 15.3 (2.5) 2.8 (0.5) 0.50 (0.09) 11.8 (0.2) 1.40 (0.04) Oxidizing of grinding waste 9.5 (0.7) 2.6 (0.2) 0.54 (0.12) 13.7 (0.8) 1.8 (0.1) 0.46 (0.12) 4.0 (1.0) 2.4 (0.3) General workshop air 16.2 (0.7) 1.8 (0.05) 0.57 (0.84) 18.9 (1.2) 1.5 (0.2) 0.43 (0.84) 10.4 (13.6) 2.0 (0.8) Pelletizing workshop

Pellet pressb 10.9 (0.6) 2.2 (0.1) 0.87 (0.03) 10.9 (0.3) 2.0 (0.1) 0.13 (0.03) 0.3 (0.1) 2.6 (1.0)

Emptying of sintered pelletsc 8.6 (0.3) 2.20 (0.07) 0.91 (0.02) 9.3 (0.4) 2.3 (0.1) 0.09 (0.02) 6.6 (0.6) 1.2 (0.2) Pellet inspection 15.1 (0.8) 2.10 (0.08) 0.69 (0.79) 17.1 (1.4) 1.8 (0.3) 0.31 (0.79) 6.3 (16.1) 2.6 (2.5) General ventilation exhaustc 12.4 (0.8) 2.1 (0.1) 0.11 (0.17) 18.6 (6.3) 1.3 (0.3) 0.89 (0.17) 10.8 (1.4) 2.1 (0.2)

BA pelletizing workshop

Mixing with Gd2O3c 7.6 (0.5) 2.2 (0.1) 0.17 (0.04) 10.0 (0.1) 1.10 (0.02) 0.83 (0.04) 6.9 (0.3) 2.3 (0.1)

Oxidizing discarded pellets,

grinding waste 17.2 (0.4) 1.8 (0.03) 0.88 (0.40) 17.7 (1.1) 1.7 (0.1) 0.12 (0.40) 6.3 (19.1) 2.3 (3.1)

a

Based on the average of three rounds of sampling.

bFinal filter affected by particle bounce. c

(9)

Urinary excretion

The resulting urinary excretion rate following

inhala-tion intake of 1 Bq234U is presented in Fig. 4. The urinary

excretion rate associated with the ICRP default assumption

(AMAD = 5mm, GSD = 2.5) is included for comparison.

The expected urinary excretion per unit of inhalation

intake 100 d after exposure was found to be 13–34% of that

corresponding to the default assumption of AMAD 5mm

and Type M/S material. Lower excretion rates are expected since a high AMAD results in a smaller fraction of activity reaching the alveoli and thus the blood.

Uncertainties

Although variable isotope ratios (Table 1) are a strong indicator of bimodal/multimodal activity size distributions, the parameters associated with the fine fractions proved

Fig. 3. Activity size distributions in the breathing zone at the different workshops. Coarse and fine fractions from bimodal curve-fitting for average fractions of activity at each impactor stage are compared with the ICRP default assumption. The coarse and fine fractions were summed and nor-malized in the lower subplot.

Table 4. Dose coefficients (CED and Hlung) at the different workshops, following inhalation intake of234U. ICRP default AMAD (5mm) for both solubility classes (Type F/M and Type M/S) are included for comparison. Doses are based on an alpha radiation weighting factor of 20 (ICRP 2007). Absorption parameters from ICRP 137 were used.

Coarse fraction Fine fraction Weighted dose coefficient

f CED (mSv Bq1) Hlung (mSv Bq1) 1-f CED (mSv Bq1) Hlung (mSv Bq1) CED (mSv Bq1) Hlung (mSv Bq1) Conversion (Type F/M) 0.88 0.1 0.2 0.12 0.4 2.7 0.1 0.5 Conversion (Type M/S) 0.88 1.3 1.1 0.12 4.0 21.9 1.6 3.6

Powder preparation (Type M/S) 0.75 1.6 3.0 0.25 5.5 33.6 2.6 10.7

Pelletizing (Type M/S) 0.85 1.5 2.4 0.15 5.5 36.0 2.1 7.5

BA pelletizing (Type M/S) 0.87 1.6 3.0 0.13 5.6 46.0 2.1 8.6

ICRP 5mm, Type F/M − − − − − − 0.6 3.6

ICRP 5mm, Type M/S − − − − − − 5.0 29.9

1

Calculated by using ICRP Publication 137 absorption data in the ICRP 66 particle transport model. Dose coefficient presented in the ICRP 137 publication is 5.5mSv Bq1.

(10)

difficult to be quantify as indicated by large uncertainties (Tables 2 and 3). Dose coefficients and excretion patterns are based on these derived parameters and must thus be interpreted with care.

Total analytical uncertainties due to counting statistics, detector efficiency (gross alpha counting only), tracer un-certainty, etc., were estimated to be <10%. The analytical uncertainties were considered lower than sampling-related uncertainties, including sample representativeness, particle bounce/roll-off, and impactor stage collection efficiency. Particle bounce/roll-off can be reduced by coating impac-tion substrates; e.g., by applying a thin layer of silicone oil. It is important to consider coating material, thickness, and homogeneity. To improve reproducibility, simplify sampling, and allow for easy comparison with planned in vitro dissolution rate studies, coating of the impaction sub-strates was not carried out in the present work except for a comparative measurement (Appendix B). Previous studies using the same type of impactor have concluded that some roll-off can be expected at impaction stages G and H (Rubow et al. 1987). Cheng et al. used the same type of impactor with coating applied on Stage A only. Stages G and H were uncoated, but the authors reported that bounce did not appear to be significant based on deposition patterns (Cheng et al. 2009). Another study concluded that measurements with uncoated impaction substrates tended to generate lower

AMADs, although the difference was not statistically significant (Kirychuk et al. 2009). In the present work, the effect of coated and uncoated substrates was evaluated by parallel rounds of stationary sampling (Appendix B, Table B1 and B2, Fig. B1). Coating of substrates generated a somewhat lower fraction of activity on late impactor stages, in particular stages H and G, as well as the final collection filter. This might indicate that some sampling artefact, e.g. particle bounce, occurs. The presence of particles larger than expected compared to impactor stage cut-point (stages G-H and final collection filter) was verified using scanning electron microscopy for both coated and uncoated rounds of sampling. This might introduce a bias toward higher 1-f and GSD, and lower AMAD, for the fine fraction. However, modeled AMADs and GSDs remained similar for coated and uncoated rounds of sampling, regardless of whether a unimodal or bimodal approach was used (Fig. B1). We consider the effect of coating difficult to quantify due to variations between different rounds of samplings and uncertainties introduced in the coating procedure. It should be mentioned that particle bounce/roll-off is most pronounced at high particle loads (Fujitani et al. 2006). Particle loads in the present work were lower (< 0.1 mg

assuming a specific activity of approximately 90 Bq mg−1)

compared to other studies (approximately 0.2–0.9 mg)

(Cheng et al. 2009).

Fig. 4. Predicted urinary excretion rate following inhalation of 1 Bq234U based on derived AMADs and GSDs at the different workshops. The ICRP 130 particle transport and ICRP 137 absorption parameters for Type F/M and Type M/S were used. Type F/M was considered for the con-version workshop only.

(11)

Impactor stage collection efficiencies were determined by Rubow et al. (1987) using monodisperse aerosols. Early impaction stages are associated with lower collection efficiencies; thus, the impact on the overall uncertainty is greater when a large fraction of the activity is collected on early impaction stages. In the present work, at least 50% of the measured activity was typically collected at impaction stages A-C (collection efficiencies of 0.52, 0.61, and 0.78, respectively). The impact on derived AMADs, GSDs, and f for the pelletizing workshop was tested by modifying the collection efficiencies in the data reduction protocol with ± 20% for stages A-C. The impact on the derived f, AMAD, and GSD for the coarse fraction was less than 5%. The impact on the AMAD and GSD for the fine fraction was within uncertainties presented in Table 2.

CONCLUSION

In the present work, cascade impactors were used to evaluate activity size distributions at a nuclear fuel fabrica-tion plant using wet-route AUC conversion. Sampling was carried out in the operator breathing zone at the four main

workshops and at certain process steps. Variable234U/238U

isotope ratios indicated a multimodal rather than unimodal activity size distribution for breathing zone sampling. A bi-modal distribution (coarse and fine fraction) was assumed.

Most activity (75–88%) was associated with the coarse

frac-tion (AMAD 15.2–18.9 mm). The AMAD of the fine fraction

was 1.7–7.1 mm, but uncertainties were substantial. When

static sampling was carried out, the coarse fraction consisted of a smaller fraction of the activity, and the AMAD was lower. We conclude that although the parameters associated with the fine fractions were difficult to quantify, the presence of a fine fraction is nonetheless very important to consider in CED assessments. The predicted CED per unit of inhalation

intake (234U) at the four main workshops was estimated to

1.6-2.6mSv Bq−1for the Type M/S solubility class, compared

to 5.0mSv Bq−1when assuming an AMAD of 5mm. The

predicted urinary excretion of234U per unit of inhaled activity

at the four workshops at 100 d after intake was 13–34% of the

excretion rate when an AMAD of 5mm was assumed.

The findings in the present work allow for more realis-tic assumptions regarding activity size distributions, improv-ing CED evaluations at the site. It also shows that internal dosimetry preferably should be based on breathing zone data rather than static sampling. Improved dosimetry can aid in radiation protection optimization and in future epidemiolog-ical studies. Future work includes determination of absorp-tion parameters and fracabsorp-tional uptake to the alimentary tract.

Acknowledgments—The Swedish Radiation Safety Authority is acknowledged for funding the present work (grant number SSM2016-589-2). Westinghouse Electric Sweden AB is acknowledged for participation in the study and funding of the PhD program.

Special thanks go to Patrick O’Shaughnessy at the University of Iowa for help with the data reduction protocol, and to Jörgen Gustafsson and numerous colleagues at Westinghouse Electric Sweden AB for invaluable discussions. The operators at the site are thanked for assisting with sampling of aerosols.

REFERENCES

Ansoborlo E, Henge-Napoli MH, Chazel V, Gibert R, Guilmette RA. Review and critical analysis of available in vitro dissolution tests. Health Phys 77:638–645; 1999. DOI 10.1097/00004032-199912000-00007.

Bergqvist H, Sahle W. Electron microscopy characterization and x-ray measurements of UO2 powder and UO2 based fuel pellet materials. Royal Institute of Technology, Stockholm, Sweden; FNM/ICT, KTH Report No. 100331; 2010.

Birchall A, Puncher M, Marsh JW, Davis K, Bailey MR, Jarvis NS, Peach AD, Dorrian MD, James AC. IMBA Professional Plus: a flexible approach to internal dosimetry. Radiat Protect Dosim 125:194–197; 2007. DOI 10.1093/rpd/ncl171. Birchall A, Vostrotin V, Efimov A, Sokolova A, Suslova K,

Zhdanov A, Schadilov A, Puncher M, Dorrian M-D, Napier B, Strom DJ, Scherpelz R, Miller S. The Mayak worker dosimetry system (MWDS-2013) for internally deposited plutonium: an overview. Radiat Protect Dosim 176:10–31; 2017. DOI 10. 1093/rpd/ncx014.

Chang M, Kim S, Sioutas C. Experimental studies on particle im-paction and bounce: effects of substrate design and material. Atmos Environ 33:2313–2322; 1999. DOI 10.1016/S1352-2310(99)00082-5.

Cheng YS, Kenoyer JL, Guilmette RA, Parkhurst MA. Physico-chemical characterization of Capstone depleted uranium aerosols II: particle size distributions as a function of time. Health Phys 96: 266–75; 2009. DOI 10.1097/01.HP.0000290613.41486.cb. Davesne E, Blanchardon E. Physico-chemical characteristics of

uranium compounds: a review. Int J Radiat Biol 90:975–988; 2014. DOI 10.3109/09553002.2014.886796.

Dorrian MD, Bailey MR. Particle size distributions of radioactive aerosols measured in workplaces. Radiat Protect Dosim 60: 119–133; 1995. DOI 10.1097/HP.0b013e318287321d. Fujitani Y, Hasegawa S, Fushimi A, Kondo Y, Tanabe K, Kobayashi

S, Kobayashi T. Collection characteristics of low-pressure impac-tors with various impaction substrate materials. Atmos Environ 40:3221–3229; 2006. DOI 10.1016/j.atmosenv.2006.02.001. Gilbert ES. Ionising radiation and cancer risks: what have we

learned from epidemiology? Int J Radiat Biol 85:467–482; 2009. DOI 10.1080/09553000902883836.

Grellier J, Atkinson W, Berard P, Bingham D, Birchall A, Blanchardon E, Bull R, Guseva Canu I, Challeton-De Vathaire C, Cockerill R, Do MT, Engels H, Figuerola J, Foster A, Holmstock L, Hurtgen C, Laurier D, Puncher M, Riddell AE, Samson E, Thierry-Chef I, Tirmarche M, Vrijheid M, Cardis E. Risk of lung cancer mortality in nuclear workers from internal exposure to alpha particle-emitting radionuclides. Epidemiol 28: 675–684; 2017. DOI 10.1097/EDE.0000000000000684. Hallstadius L. A method for the electrodeposition of actinides.

Nucl Instrum Meth Phys Res 223:266–267; 1984. DOI 10. 1016/0167-5087(84)90659-8.

Hansson E, Pettersson HBL, Fortin C, Eriksson M. Uranium aero-sols at a nuclear fuel fabrication plant: characterization using scanning electron microscopy and energy dispersive x-ray spectroscopy. Spectrochimica Acta Part B: Atomic Spectros-copy 131:130–137; 2017. DOI 10.1016/j.sab.2017.03.002. Harrison J, Day P. Radiation doses and risks from internal

emit-ters. J. Radiol. Prot. 28:137–159; 2008. DOI 10.1088/0952-4746/28/2/R01.

(12)

Hinds WC. Aerosol technology: properties, behavior, and measure-ment of airborne particles. New York: Wiley & Sons; 1999. Holm E. Review of alpha-particle spectrometric measurements of

actinides. Int J of Applied Radiation and Isotopes 35(4):285-290; 1984. DOI 10.1016/0020-708X(84)90070-X.

International Atomic Energy Agency. The front end of the ura-nium fuel cycle. 50th IAEA General Conference, Documents [online]. 2006. Available at https://www.iaea.org/About/ Policy/GC/GC50/GC50InfDocuments/English/gc50inf-3-att6[lowen]en.pdf. Accessed 1 August 2018.

International Commission on Radiological Protection. Human respi-ratory tract model for radiological protection. Oxford: Pergamon Press; ICRP Publication 66; Ann ICRP 24(1–3); 1994.

International Commission on Radiological Protection. Individual monitoring for internal exposure of workers. Oxford: Pergamon Press; ICRP Publication 78; Ann ICRP 27(3–4); 1997. International Commission on Radiological Protection.

Recom-mendations of the International Commission on Radiological Protection. Oxford: Pergamon Press; ICRP Publication 103; Ann ICRP 37(2–4); 2007.

International Commission on Radiological Protection. Occupational intakes of radionuclides: part 1. Oxford: Pergamon Press; ICRP Publication 130; Ann ICRP 44(2); 2015.

International Commission on Radiological Protection. Occupa-tional intakes of radionuclides: part 3. Oxford: Pergamon Press; ICRP Publication 137; Ann ICRP 46(3–4); 2017. Kemmer G, Keller S. Nonlinear least-squares data fitting in Excel

spreadsheets. Nature Protocols 5:267–281; 2010. DOI 10. 1038/nprot.2009.182.

Kirychuk SP, Reynolds SJ, Koehncke N, Nakatsu J, Mehaffy J. Comparison of endotoxin and particle bounce in Marple cascade

samplers with and without impaction grease. J Agromed 14: 242–249; 2009. DOI 10.1080/10599240902845062.

O'Shaughnessy PT, Raabe OG. A comparison of cascade impactor data reduction methods. Aerosol Sci Technol 37:187–200; 2003. DOI 10.1080/02786820300956.

Palheiros F, Gonzaga R, Soares A. Comparative study of the differ-ent industrial manufacturing routes for UO2pellet specifications

through the wet process. NAC 2009: International nuclear atlan-tic conference Innovations in nuclear technology for a sustain-able future, Brazil [online]. 2009. Availsustain-able at https://inis.iaea. org/collection/NCLCollectionStore/_Public/41/057/41057359. pdf?r=1&r=1. Accessed 17 February 2020.

Rubow K, Marple V, Olin J, Mccawley M. A personal cascade impac-tor: design, evaluation and calibration. Am Industrial Hygienists Assoc J 48:532–538; 1987. DOI 10.1080/15298668791385174. Schieferdecker H, Dilger H, Doerfel H, Rudolph W, Anton R.

Inhalation of U aerosols from UO2 fuel element fabrication.

Health Phys 48:29–48; 1985.

Thermo Fisher Scientific Inc. Marple personal cascade impactors Series 290, instruction manual, part number 100065-00. 2009. Thiel CG. Cascade impactor data and the lognormal distribution: nonlinear regression for a better fit. J Aerosol Med 15:369–378; 2002. DOI 10.1089/08942680260473443.

Thind KS. Determination of particle size for airborne UO2dust at

a fuel fabrication work station and its implication on the deriva-tion and use of ICRP Publicaderiva-tion 30 derived air concentraderiva-tion values. Health Phys 51:97–105; 1986.

(13)

APPENDIX A—RADIOMETRIC DATA FOR STATIC SAMPLING Ta b le A 1 . Uranium act iv ity concentrations (static samplin g) ba se d o n g ro ss alp h a coun ting .Unc er ta intie s cor re spond to ± 1 standard d ev ia tion due to countin g stat istics and flo w rate me asu rem ents (2 .00 ± 0.05 L m in − 1 ). Urani u m act iv it y concentration (Bq m  3) Sampling time (m in) To ta l concentration (Bq m  3 ) Stage A 21 .3 m m St ag e B 14 .8 m m St age C 9. 8 m m Sta g e D 6. 0 m m Sta g e E 3. 5 m m Sta g e F 1. 6 m m Sta g e G 0. 9 m m Sta g e H 0. 5 m m Fi n al filter Co n versio n wor ksho p F lu id iz in g b ed fu rn ac e 0. 0013 (0 .0002) 0.0035 (0 .0003) 0.0062 (0 .0004) 0.0052 (0.0004) 0.0039 (0.0003) 0.0041 (0.0003) 0.0011 (0.0001) 0.0004 (0. 0001) 0.0047 (0.0003) 2 6 8 2 0.03 (0.01) Gene ral ve n tila tion exhaust 0 .0008 (0 .0002) 0.0013 (0 .0003) 0.0030 (0 .0003) 0.0030 (0.0004) 0.0018 (0.0002) 0.0015 (0.0003) 0.0003 (0.0002) 0.00 03 (0.0003) 0.0015 (0.0003) 1 9 8 3 0.01 (0.02) P o w d er pr ep a ration w or ks h o p P o w d er m il li n g 0 .034 (0 .003) 0 .051 (0 .004) 0.133 (0. 008) 0.16 (0.01) 0.136 (0.008) 0.30 (0.02) 0.17 (0.01) 0.0404 (0.003) 0 .034 (0.003) 3 9 6 1.0 (0.2) H u m id it y ch ec k 0. 52 (0 .03) 0.39 (0 .02) 0.71 (0 .04) 0.66 (0.04) 0.50 (0.03) 0.55 (0.03) 0.118 (0.007) 0.075 (0.004) 0 .122 (0.007) 1 3 8 2 3.6 (0.6) Oxi dizing of dis carded pe llets 0. 23 (0 .03) 0.35 (0 .01) 0.84 (0 .04) 0.68 (0.03) 0.21 (0.04) 0.09 (0.03) 0.07 (0.02) 0.05 (0.02) 0.11 (0.02) 1 3 3 5 2.64 (0.09) O x id iz in g of g ri nd in g w as te 0 .141 (0 .008) 0 .22 (0. 01) 0.40 (0 .02) 0.46 (0.03) 0.40 (0.02) 0.42 (0.02) 0.118 (0.007) 0.053 (0.003) 0 .037 (0.002) 9 4 2 2. 3( 0 .4 ) G en er al w o rk sh o p ai r 0. 042 (0 .003) 0 .066 (0 .004) 0.075 (0. 004) 0.059 (0.003) 0.028 (0.002) 0.018 (0.001) 0.0072 (0.0005) 0.0056 (0.0004) 0.0052 (0. 0004) 1 2 5 5 0.30 (0.06) P elletiz ing w or kshop P el le t p re ss 0 .081 (0 .005) 0 .128 (0 .007) 0.25 (0 .01) 0.30 (0.02) 0.19 (0.01) 0.099 (0.006) 0.026 (0.002) 0.0158 (0.001) 0 .17 (0.01) 1 4 2 5 1.2 (0.2) E m pt yi ng of si nt er ed pe lle ts 0 .3 0 (0.02) 0.47 (0 .03) 0.83 (0 .05) 1.54 (0.09) 1.20 (0.07) 0.66 (0.04) 0.070 (0.004) 0.026 (0.002) 0 .126 (0.007) 5 5 9 6 5. 2( 0 .9 ) P el le t in sp ec ti o n 0 .022 (0 .001) 0 .027 (0 .002) 0.033 (0. 002) 0.032 (0.002) 0.018 (0.001) 0.0135 (0.0008) 0.0052 (0.0004) 0.0026 (0.0002) 0.0043 (0. 0003) 2 8 8 7 0.16 (0.03) Gene ral ve n tila tion exhaust 0 .0077 (0 .0006) 0.0169 (0 .001) 0.022 (0. 001) 0.026 (0.002) 0.0156 (0.0009) 0.0105 (0.0007) 0.0038 (0.0003) 0.0023 (0. 0002) 0.0025 (0.0002) 3 0 4 6 0.11 (0.02) B A pel let izin g w or ks hop Mixing wi th Gd 2 O3 0. 0119 (0 .0009) 0.0188 (0 .0013) 0.0627 (0 .0037) 0.0893 (0.0051) 0.0784 (0.0045) 0.0588 (0.0034) 0.0129 (0.0009) 0.0005 (0.0002) 0.0077 (0.0005) 1 9 3 8 0.3 (0.1) Oxidi zing o f d is carded p ell ets and g rin ding wa st e 0. 22 (0 .01) 0.26 (0 .01) 0.30 (0 .02) 0.19 (0.01) 0.079 (0.005) 0.052 (0.003) 0.021 (0.001) 0.013 (0.001) 0 .022 (0.001) 1 1 1 7 1.2 (0.2)

(14)

APPENDIX B—PARALLEL SAMPLING

■■

Ta b le B 1 . F rac tion o f acti v ity at each imp acto r st age , ba sed o n radiometric data follo wing total alpha countin g o f sa m ple s co llec ted du ring pa ralle l sa m plin g to in v estigate th e ef fect from p ar -tic les bou nce . Sampl ing w as car ri ed o u t at th e p o wde r pre p ar atio n w o rksh op, w h er e o x idizi ng of w aste p ell ets is ca rr ie d o u t, at tw o d if fe re nt o cca sio ns with th ree im p actors ru nning in p arallel. Unc er ta intie s cor re spon d to ± 1 sta nda rd de v iatio n due to cou n ting st atisti cs and flo w rate m easurements (2.00 ± 0 .05 L m in − 1 ). F rac tio n of ac ti v ity (-) Stage A 21 .3 m m Sta g e B 14.8 m m Sta g e C 9.8 m m St ag e D 6. 0 m m St age E 3. 5 m m Sta g e F 1. 6 m m Stage G 0.9 m m Stage H 0.5 m m F inal filt er Occasi on 1 Imp ac tor 1 – n o co ating ap p lie d 0 .105 (0 .00 6 ) 0 .14 0 (0 .00 8 ) 0 .3 2 (0 .02 ) 0 .2 6 (0. 01) 0.0 6 7 (0.0 04) 0. 022 (0 .001 ) 0 .01 8 (0 .00 1 ) 0 .016 0 (0 .00 09) 0. 057 (0 .00 3 ) Imp ac tor 2 – co atin g appli ed 0 .103 (0 .00 6 ) 0 .16 5 (0 .00 9 ) 0 .3 4 (0 .02 ) 0 .2 5 (0. 01) 0.0 7 2 (0.0 04) 0. 027 (0 .002 ) 0 .010 0 (0.00 06 ) 0 .006 0 (0 .00 04) 0. 035 (0 .00 2 ) Imp ac tor 3 – co atin g appli ed 0 .107 (0 .00 6 ) 0 .16 3 (0 .00 9 ) 0 .3 4 (0 .02 ) 0 .2 4 (0. 01) 0.0 8 4 (0.0 05) 0. 027 (0 .002 ) 0 .008 0 (0.00 05 ) 0 .005 0 (0 .00 03) 0. 019 (0 .00 1 ) Occasi on 2 Imp ac tor 1 – n o co ating ap p lie d 0 .081 (0 .00 5 ) 0 .12 1 (0 .00 7 ) 0 .3 2 (0 .02 ) 0 .2 6 (0. 01) 0.0 9 7 (0.0 06) 0. 043 (0 .003 ) 0 .03 0 (0 .00 2 ) 0 .016 0 (0 .00 09) 0. 035 (0 .00 2 ) Imp ac tor 2 – n o co ating ap p lie d 0 .071 (0 .00 4 ) 0 .13 9 (0 .00 8 ) 0 .3 3 (0 .02 ) 0 .2 5 (0. 01) 0.0 7 6 (0.0 04) 0. 041 (0 .002 ) 0 .02 5 (0 .00 2 ) 0 .03 0 (0 .00 2 ) 0 .038 (0 .00 2 ) Imp ac tor 3 – co atin g appli ed 0 .098 (0 .00 6 ) 0 .14 7 (0 .00 8 ) 0 .3 2 (0 .02 ) 0 .2 6 (0. 01) 0.1 0 1 (0.0 06) 0. 033 (0 .002 ) 0 .012 0 (0.00 08 ) 0 .006 0 (0 .00 04) 0. 031 (0 .00 2 ) Av er a g e Co ati n g appl ied 0 .10 (0 .01 ) 0 .1 6 (0 .02 ) 0 .3 3 (0 .03 ) 0 .2 5 (0. 02) 0.0 8 6 (0.0 08) 0. 029 (0 .003 ) 0 .01 0 (0 .00 1 ) 0 .006 0 (0 .00 07) 0. 028 (0 .00 3 ) No co at ing app lie d 0 .086 (0 .00 8 ) 0 .1 3 (0 .01 ) 0 .3 2 (0 .03 ) 0 .2 6 (0. 03) 0.0 8 0 (0.0 08) 0. 035 (0 .004 ) 0 .02 4 (0 .00 3 ) 0 .02 0 (0 .00 2 ) 0 .043 (0 .00 4 ) Ratio coating/no coating 1.2 (0 .2) 1. 2 (0 .2) 1. 0 (0 .1) 1 .0 (0. 1) 1 .1 (0.1 ) 0.8 (0. 1) 0.4 1 (0 .06 ) 0.2 8 (0 .04 ) 0 .66 (0 .09 )

Fig. B1. Comparison of activity size distributions from sampling with and without coated impactor substrates.

(15)

Ta b le B 2 . Deri v ed av era ge (1 S D ) p ar ame ter s for co ar se an d fin e fra ctio ns ba sed o n d ata in T ab le B1. Unimodal fi t Bimoda l fit Coarse fracti on F ine fraction AMAD (m m) GSD A MAD 1 (m m) GSD 1 f A MAD 2 (m m) GSD 2 1 -f Co ati n g appl ied 1 3.1 (0. 3) 1. 7 (0 .03 ) 14. 2 (0 .6) 2 .1 (0.1 ) 0 .7 (0.1 ) 12. 1 (0 .1) 1 .4 (0.0 3 ) 0 .3 (0 .1) No co at ing app lie d 1 2.4 (0. 3) 1. 8 (0 .03 ) 15. 0 (1 .9) 2 .9 (0.4 ) 0 .5 (0.1 ) 11. 8 (0 .2) 1 .4 (0.0 2 ) 0 .5 (0 .1)

References

Related documents

Figure 21: Nuclear data uncertainty for Normal Case using fast TMC with selected libraries from benchmark LEU-COMP-THERM-016 Case

In view of the inherent danger entailed in the passage of nuclear-powered vessels or vessels carrying nuclear material or other material of a similar nature and in view of the

The main function of it is to simulate the transport of multi-particles including neutrons, photons and electrons (both individually and together) with continuous

These five explicit learning tools were presented to us at Ringhals: Introduction to New Employees, Simulator Training, Competence Analysis, Experience Feedback and the

Maskinen står enligt Estermann inte bara mellan det skapande subjektet och hans verk utan också mellan det receptiva subjektet och dess konstnjut­ ning..

From the extraction of uranium from rock formations, through the milling, refining, and enriching of uranium, to the operation of reactors, and the unsolved dilemma of what to do

Han menar även att hela teamet kände sig väldigt positiva till införandet av de agila metoderna och att arbetet känns mycket roligare nu eftersom de får ökad kontroll på vad

On the other hand, if there is little investment in providing access to skilled healthcare during child birth, many skilled health- care centres are required.. Results show further