<<

FI9800105

STUK-A 145 March 1 998

Dating of and determination of sedimentation rate

Proceedings of a seminar held in Helsinki 2 - 3 April 1997

E. llus (ed.)

STUK • SATEILYTURVAKESKUS • S TR A LS A K E R H ETSC E N TR A LEN AND NUCLEAR SAFETY AUTHORITY &STUK A 1 4 5 March 1998

Dating of sediments and determination of sedimentation rate

Proceedings of a seminar held in Helsinki 2 - 3 April 1997

E. llus (ed.)

29-30

STUK • SATEILYTURVAKESKUS • STRALSAKERHETSCENTRALEN RADIATION AND NUCLEAR SAFETY AUTHORITY STUK-A145

ILUS, Erkki (ed.). Dating of sediments and determination of sedimentation rate. Proceedings of a seminar held in Helsinki 2 - 3 April 1997. Helsinki 1998, 149 pp.

ISBN 951-712-226-8 ISSN 0781-1705

Keywords Dating of sediments, sedimentation rate, 210Pb dating method, in sediments

FOREWORD

NKS (Nordic Nuclear Safety Research) is a cooperative Nordic body for nuclear safety, radiation protection and emergency preparedness. Its purpose is to combine resources available in the individual Nordic countries into joint pro- jects to yield research results, knowledge and information covering these top- ics. The work of the NKS is based on 4- programmes; the current pro- gramme having been planned for the 1994-1997. This programme com- prises 3 major fields, one of them being environmental effects (EKO). Under this umbrella there are 4 main projects. The EKO-1 Project deals with marine radioecology, in particular bottom sediments and processes.

All the Nordic countries have participated in the EKO-1 Project with their own subprojects. In addition to the research carried out in each country, the EKO-1 Group has arranged seminars, to which renowned lecturers from outside the Group and the participants in the project are invited. The NKS has financially supported the arrangement of the seminars. This kind of seminar in which outside experts and the participants can meet and exchange their knowledge and experience of the project topic has proved very useful and has formed an essential part of the project.

This booklet is a compilation of scientific papers presented in the Second NKS/EKO-1 Seminar 'Dating of sediments and determination of sedimentation rate' held at the Finnish Centre for Radiation and Nuclear Safety (STUK) on April 2-3, 1997. The programme of the Seminar consisted of 8 invited lecturers STUK-A145

and 6 other scientific presentations. The seminar was attended by 48 scientists from 9 countries.

Dating of sediments and determination of sedimentation rate are of crucial importance in all types of sedimentological study and model calculations of fluxes of substances in the aquatic environment. In many cases these tasks have been closely related to radioecological studies undertaken in marine and fresh water environments, because they are often based on measured depth profiles of certain natural or artificial radionuclides in the sediments. In this connection radioecologists have had an opportunity to give their valu- able contribution to limnological and oceanographic studies.

During recent decades Pb-21O has proved to be very useful in dating of sedi- ments, but some other radionuclides have also been successfully used, e.g. Pu- 239,240, Am-241 and Cs-137 (the last-mentioned especially after the Cherno- byl accident in 1986). The methods used in the estimation of sedimentation rate, however, are by no means totally routine procedure, but full of 'pitfalls'. The difficulties existing and problems involved in dating of sediments, as well as solutions for resolving these problems are discussed in the papers of these Proceedings.

Erkki Ilus STUK-A145

CONTENTS page

FOREWORD 3

DATING RECENT SEDIMENTS BY 21°Pb: PROBLEMS AND SOLUTIONS 7 P. G. Appleby

THE USE OF FLY-ASH PARTICLES FOR DATING LAKE SEDIMENTS 25 N. L. Rose

SEDIMENT MIXING MODELS IN THE INTERPRETATION OF 2i°Pb CONCENTRATION PROFILES IN CONTINENTAL SHELF SEDIMENTS OF EASTERN CANADA 40 S. Mulsow, B. Boudreau, J. N. Smith

THE PAST AND THE FUTURE OF THE 2iopb METHOD 54 F. El-Daoushy, R. Garcia-Tenorio

ARTIFICIAL RADIONUCLIDES IN AN INTERTIDAL SEDIMENT FROM NORTHWEST ENGLAND 62 K. Morris, M. J. Keith-Roach, A. S. Hursthouse, J. C. Butterworth, F. R: Livens, L. K. Fifield, J. P. Day, R D. Bardgett

THE MOBILITY OF RADIOCAESIUM IN LAKE SEDIMENT AND IMPLICATIONS FOR DATING STUDIES 76 J. T. Smith, D. G. Ireland, R. N. J. Comans, L. Nolan

KINETICS AND REVERSIBILITY OF RADIOCAESIUM SOLID/LIQUID PARTITIONING IN SEDIMENTS 94 R. N. J. Comans

TRACING THE CHERNOBYL FALL-OUT PEAK IN FINNISH LAKE SEDIMENTS IN ORDER TO OBTAIN A GOOD MARKER 116 H. Jungner STUK-A145

RADIONUCLIDES IN SEDIMENT CORES FROM THULE, GREENLAND, 1991 120 S. P. Nielsen, H. Dahlgaard

DATING OF LAMINATED SEDIMENTS IN SWEDISH ARCHIPELAGO AREAS OF THE BALTIC SEA 127 M. Meili, P. Jonsson, R. Carman

SEDIMENTATION AND WITHIN-BASIN VARIATIONS IN THE GULF OF FINLAND AS DETERMINED BY "'Cs TRACER 131 H. Vallius, H. Kankaanpaa, L. Niemisto, O. Sandman

EVALUATION OF SEDIMENTATION RATE AT TWO SAMPLING STATIONS IN THE GULF OF FINLAND BASED ON Pb-210, Cs-137 AND Pu-239,240 PROFILES IN SEDIMENTS 136 E. Ilus, J. Mattila, S. Klemola, T. K. Ikaheimonen

DEVELOPMENT OF INSTRUMENTATION FOR DATING METHOD 148 P. Reinikainen, J. Merilainen, I. Rekikoski, A. Virtanen, A. Witick, J. Aysto STUK-A145

DATING RECENT SEDIMENTS BY 210Pb: PROBLEMS AND SOLUTIONS

P. G. Appleby

Department of Mathematical , University of Liverpool Liverpool L69 3BX, UK

1 Introduction

Accurate dating by 210Pb is of crucial importance to a wide range of programs studying environmental records stored in natural archives such as lake sedi- ments or peat bog accumulations. In a large number of cases the method has proved to be very reliable, particularly in stable environments with uniform sediment accumulation rates where the dating calculations are unambiguous.

The method has also been found to give good result at many sites with non- uniform accumulation, though here the problem is more difficult in view of the need to determine an appropriate dating model. There are two simple models, commonly referred to as the CRS and CIC models (Appleby & Oldfield 1978, Robbins 1978). Of these, the CRS (constant rate of 210Pb supply) model is per- haps the most widely used. The main principles of this model are exemplified in Appleby et al. (1979) by cores from three Finnish lakes with annually lami- nated sediments, all of which contained layers recording dilution of the atmos- pheric 210Pb flux by increased sedimentation. 210Pb dates calculated using the CRS model were in good agreements with those determined by laminae counting. There are however circumstances where the CIC (constant initial concentration) model is appropriate, e.g. in a core from Devoke Water (Appleby & Oldfield 1992) where the CRS model was invalidated by an abrupt disconti- nuity in the sediment record.

In a very real sense these models should in the first instance be regarded as tools whose purpose is to determine, as far as practicable, the processes that have generated the data contained in the sediment record. In each case the 210Pb data must be individually assessed in light of any independent chrono- stratigraphic evidence such as that provided by 137Cs or 241Am. The object of this paper is to highlight the conceptual framework that forms the basis of this STUK-A145

assessment and to show how it can be used to solve problems that have arisen in a number of practical cases. These include the use of hybrid models where there has been a variable 210Pb supply, corrections for inaccuracies in the cal- culation of radiometric inventories, and the impact of large variations in 226Ra activity.

2 Atmospheric Fluxes

Since dating by 210Pb is based on sediment records of 210Pb fallout, a proper assessment of the record requires a good knowledge of the atmospheric flux. Fallout 210Pb derives from the decay of gaseous 222Rn introduced into the at- mosphere by from soils, and is normally assumed to have a constant flux at any given site (when averaged over a period of year or more). The amount of fallout may however vary from place to place by an order of magni- tude, depending on factors such as rainfall and geographical location. This is illustrated by the measurements of 210Pb fallout in Great Britain and Central Europe summarised in Fig. 1. These show that at a regional level there is a strong correlation with rainfall. Overall levels of fallout in Central Europe are however significantly higher than in Great Britain, presumably due to a build of 222Rn concentrations in the atmosphere as the prevailing winds transport air masses over the intervening land surface. The global summary given in Fig. 2 suggests that there is a consistent west to east increase in 210Pb fallout within the major continents, superimposed on a baseline 210Pb flux of c. 30 - 40 Bq m2 y1 per metre of rain at sites remote from major land masses.

3 Long-term Sediment Records of 210Pb fallout

Some of the earliest studies of fallout radionuclides in lake sediments were carried out by Pennington et al. (1973 & 1976) on cores from lakes in the Win- dermere catchment in Cumbria (UK), including Blelham Tarn, Esthwaite Wa- ter, Elterwater and Windermere itself. Further analyses carried out during the 1980s and 1990s at the Liverpool University Environmental Radioactivity Research Centre have resulted in a record of measurements from this region that now spans a period of 25 years. These lakes all have similar rainfall and so may be expected to have similar atmospheric fluxes of 210Pb. P.55»70xR/1O00/

m Arctic N America • Europe A Asia V Oceanic cr cr a Allies # / ° Australasia m B Antarctic

P = 30+56 X R/tOOO IX A E i_ • / a Q. X • / A S Q Central Europe ul / • O Gteal Britain

a. M / V / • « S V v 9 B* 7 V

0 SOO 1000 1500 2000 2500 3000 -180-160-140-120-100 -80 -60 -40 -20 0 20 40 60 80 100 120 140 160 180 Mean Annual Rainfall (mm y ) Longitude (degrees)

Figure 1. Values of the atmospheric 2I0Pb flux versus Figure 2. Values of atmospheric 210Pb flux per m of rainfall for Great Britain (o) and Central Europe (I)), annual rainfall, plotted against longitude. The plot c summarised from a variety of sources. The data are distinguishes values for each of the major land mas- > based on measurement from both soil cores and direct ses, the poles, and oceanic sites. CO precipitation. CJ1 STUK-A145

At a given core site, mean 210Pb fluxes to the sediment record can be estimated from the relation P = XA(0),

where X is the 210Pb constant and A(0) is the unsupported 210Pb inventory of the core. Fig. 3a shows the result of these calculations for the Windermere lakes, plotted against the date of coring. This shows that al- though the 210Pb flux appears to be relatively uniform through time, as as- sumed by the CRS model, values are about twice as high as the direct atmos- pheric flux (indicated by the dashed line).

Factors influencing differences between the atmospheric flux and sediment record may include inputs from the catchment, losses from the water column via the outflow, and sediment focussing. The influence of water column pro- cesses is shown by comparing 210Pb inventories with those of fallout 137Cs from the atmospheric testing of nuclear weapons. Values of the 137Cs/210Pb inventory ratio (corrected for 137Cs decay) in sediment cores from the Windermere lakes (Fig. 3b) are about half those expected from direct fallout, presumably due to relatively greater losses of 137Cs from the system (via the outflow) arising from its higher solubility (Appleby 1991).

4 Transport Models

These results highlight the fact that transport processes through catch- ment/lake systems have a strong influence on sediment records of the atmos- pheric 210Pb fluxes, and thus on 210Pb dating methodologies. Fig. 4 gives a simplistic framework for modelling these processes. Using simple box models, and assuming a constant 210Pb atmospheric flux P, the rate of supply of 210Pb to the bottom sediments is given by the expression

(c.f. Appleby & Oldfield 1992, Appleby & Smith 1993) where a is the catch- 210 ment/lake area ratio, t|pb is a Pb catchment/lake transport parameter and

FPb is a water column to sediment record transfer fraction, determined by the water residence time, water depth, particle settling velocity and solubility. Although this equation suggests that essential conditions for validity of the

10 Windermere 0 Elterwater H Blelham Tarn n Esthwaito Water

Aimospheric Input Ratio

.o a.

o e o • » D 9 B

0 1970 1985 Date

Figure 3. Values of the mean 210Pb flux (a) and the weapons 137Cs/ 2l0Pb inventory ratio (b) for sediment cores from I > lakes in the Windermere catchme7it, plotted against the date of coring. In both cases the dashed line shows direct at- mospheric fallout values. U1 STUK-A145

Atmospheric 2l0Pb Flux AcP Atmospheric 210Pb i FluxAP

Catchment Inventory Qdt) _ Lake Waters / . Transpor t x\ . _.. / Loss via c \ Inventory g(r) / „ from \ / outflow Catchment I To sediment record

Figure 4. A simple schematic model of transport processes controlling the sup- ply of fallout 2I0Pb to lake sediments.

CRS model include steady inputs from the catchment and constant residence time in the water column, these may not be of major significance in practice. Studies of catchment/lake transport (Lewis 1977; Scott et al. 1985; Dominik et al. 1987) suggest that just 1 - 2% of the annual 210Pb fallout on the catchment is removed to the lake. Further, since 210Pb is largely associated with particu- lates only a small fraction of the total amount entering the lake is lost via the outflow (Appleby & Smith 1993). In most cases the validity of either of the simple models will depend on the extent and nature of the processes control- ling sediment focussing.

An important exception to these conditions has been observed by Appleby et al. (1995) in a number of lakes from Signy Island (maritime Antarctic) where greatly elevated inventories of unsupported 210Pb in the sediments (relative to the atmospheric flux) were explained by the bulk of fallout onto the catchment being delivered to the lakes during the annual thaw. In terms of the above equation this corresponds to the case were r|pb « 1.

5 Validation of 210Pb dates

Although parameters such as the 210Pb inventory of core and the surficial con- centration, together with features in the unsupported 210Pb activity versus depth profile (changes in slope, non-monotonic "kinks") all play a role in the

12 STUK-A145

assessment of 210Pb data from sites with varying sediment accumulation (Appleby & Oldfield 1983), independent validation of the is essen- tial to a high level of confidence in the results. The most important means for validating dates for the last 30 - 40 years is via fallout records of artificial ra- dionuclides. This is well illustrated by results from two Windermere cores (Fig. 5), collected in 1985 and 1992. The 137Cs activity versus depth profile in the 1985 core has a single well resolved peak recording the maximum fallout of this in 1963 from the atmospheric testing of nuclear weapons. The origin of this feature is corroborated by the presence of a similar (but much smaller) 241Am peak (Appleby et al. 1991). The weapons fallout record (137Cs and 241Am) is again clearly present in the 1992 core. Now however there is a second more recent 137Cs peak recording fallout from the 1986 Chernobyl accident. The identification in this case is corroborated by 134Cs. Discharges from the Chernobyl reactor fire included both caesium radionuclides in a char- acteristic ratio Cambray et al. (1987). In many cases the presence of these peaks gives a very good control on the recent chronology. Fig. 6 shows that the dated levels determined from the 137Cs record in the 1992 Windermere core are in excellent agreement with the CRS model 210Pb dates.

Validation of older 210Pb dates near the base of a core remains a major prob- lem, particularly where there have been significant late 19th or early 20th century changes in accumulation rates. The most usual method is by inde- pendent chronostratigraphic features in the pollen, diatom or trace metal rec- ords.

6 Mixing or Acceleration?

Radionuclides delivered to the bed of the lake on settling particles may be re- distributed within the sediment column by two main processes, • physical or biological mixing at or near the sediment-water interface, • chemical diffusion or advection within the porewaters.

Sediment mixing typically results in a flattening of the 210Pb activity versus depth profile in the surficial sediment layers, and degradation of features such as 137Cs or 241Am peaks. The effects of these processes have been documented in many studies and a number of models have been developed to take account of them (Robbins et al. 1977; Oldfield & Appleby 1984). Since other processes can have similar effects on sediment records, an essential problem in using

13 STUK-A145

(a) 1985 Core (b) 1992 Core

8OOr

700

O> 600 S 600

35 _:—• m >- 500 !> o < 400 E 3 a0> .1 300 ou caes i 200 1 adi o DC EC

5 10 15 20 25 30 35 10 15 20 25 Depth (cm) Depth (cm)

Figure S. 137Cs ( ), 134Cs (.....) and 241Am (--—) activity versus depth profiles in a 1985 (a) and 1992 (b) sediment core from Lake Windermere, Cumbria, UK.

'37Cs^"Am Dales CRS2'°Pb0at«s CRS Sedimentation Rates I

OS IE 0.15 "O O

0 10 20 30 40 50 60 7C 80 90 100 110 120 130 140 150 160 170 180 Age (y)

Figure 6. CRS model 210Pb depth ( ) and sedimentation rate(—--) versus age relations for the 1992 Windermere core. Also shown are the 1963 and 1986 depths determined from l37Cs, 134Cs and 242Am records.

14 STUK-A145

mixing models is the difficulty in making objective determination of the mixing parameters. Estimates of such parameters should be based on at least two independent records. The importance of this is illustrated by the results from Zgornje Krisko Jezero (Slovenia) shown in Fig. 7 in which 210Pb activity in a sediment core was almost uniform throughout the top 20 cm. The initial pre- sumption that this might be due to rapid intensive mixing was shown to be incorrect by the presence of a well resolved 137Cs/134Cs peak at a depth of just 4 cm. This feature, recording fallout from the 1986 Chernobyl accident, together with elevated 241Am concentrations below 16 cm, demonstrated that mixing in these sediments was negligible and that the 210Pb profile was almost certainly due to very rapid sedimentation.

CO

.Q Q.

Depth (cm)

Figure 7. 210Pb ( ;, 137Cs (•—-) and l34Cs (.....) activity versus depth profiles in a sediment core from Zgornje Krisko Jezero, Slovenia.

15 STUK-A145

In those cases where mixing has occurred, Appleby & Oldfield (1992) have pointed out that corrections will often only be important in those cases where the CIC model is being used. Where the CRS model is applicable, calculations show that the use of uncorrected 210Pb dates in a sediment core with a mixing zone spanning 10 years accumulation gives a maximum error of less than 2 years.

Where physical mixing is demonstrably negligible, the possibility that flat- tened 210Pb profiles might be due to chemical remobilisation can be tested by comparing 210Pb dates with those determined from the 137Cs . The excellent agreement between 210Pb and 137Cs (or 241Am) in a large number of cases suggests that in most circumstances this is not a significant problem.

7 Corrections to the CRS Model for Incomplete Inventories

CRS model 210Pb dates are calculated using the formula

(Appleby & Oldfield 1978) where • X is the 210Pb decay constant, • A(0) is the unsupported 210Pb inventory of the entire core, • A is the unsupported 210Pb inventory below the sample being dated.

A and A(0) are both determined by numerical integration of the 210Pb profile. Accuracy in the dating, particularly in the first few samples above the equilib- rium depth, is crucially dependent on reliable estimates of the 210Pb invento- ries. Under-estimation of A in these samples will cause the 210Pb dates of these levels to be too old.

There are two common sources of error, uncertainties in the numerical proce- dures for estimating A, and gaps in the sediment record. In both cases the error can be reduced if there is a means for correcting the values given by simple numerical integration. This can be achieved in two ways: by estimating the accumulation rate at a reference point near the base of the core, or the date of that point.

16 STUK-A145

Reference accumulation rates

At sites where 210Pb activity in the older sections is evidently declining more or less exponentially with depth, indicating uniform accumulation, the mean sedimentation rate for that period, n, can be estimated from the gradient of the relevant section of the 210Pb profile (plotted logarithmically), using the formula

dm where m is the depth measured as cumulative dry mass. The 210Pb inventory below the first data point above the 210Pb dating horizon (c.100 - 150 years), At, can be estimated using the formula

(c.f. Appleby & Oldfield 1978) where d is the unsupported activity at that level. Dates for subsequent levels can then be calculated using the formula

where SA is the unsupported 210Pb inventory between the basal sample and the sample being dated.

Reference dates

The above method cannot be applied at sites where there have been non- uniform sediment accumulation rates in the 19th or early 20th century, or a hiatus in the sediment record. Accuracy in these cases can be improved by using an independently dated reference level. If AA denotes the entire unsup- ported 210Pb inventory above the reference level, determined by numerical integration of the activity versus depth profile, the inventory below the refer- ence level be estimated using the formula

17 STUK-A145

A; =•

(Oldfield & Appleby 1984), where (; is the age of the reference level. Revised 210Pb dates are then calculated as above.

The dated reference level must of course be at or above the 210Pb dating hori- zon. When it is below the 210Pb dating horizon (and there appear to have been significant increases in sedimentation rates in the late 19th century), it may instead be used with the older 2l0Pb to make estimates of the basal sedimen- tation rate. The value of Ai can then be estimated by the first method. The process can be iterated to give improved accuracy.

8 Varying 210Pb Supply Rates

Sediment cores frequently record episodes of rapid sediment accumulation, due for example to flood events, turbidity currents of major land-use changes. Even where the profile contains non-monotonic variations indicating signifi- cant dilution of the atmospheric 210Pb flux, the event recorded by these fea- tures may import significant quantities of additional fallout 210Pb. For a brief period the 210Pb supply rates may be significantly higher than at of nor- mal accumulation, and where this occurs the simple CRS and CIC models will both give erroneous dates.

Where there are suitably placed independently dated levels, determined by chronostratigraphic markers such as 137Cs, 241Am or pollen, these can be used to calculate the mean 210Pb flux for each zone. Assuming the flux to be uniform within the zone, corrected 210Pb dates and sedimentation rates for intermedi- ate depths can calculated by applying the principles of the CRS model with the relevant 21<>Pb flux.

The method governing the application of this piecewise CRS model can be summarised as follows:

I Given two core depths xi and X2 with known dates ti and t2, the mean 210Pb flux during the period spanned by this section is

18 STUK-A145

p = _XAA_

210 where AA is the Pb inventory between x} and x9

210 II Given the Pb flux P for a core section containing a dated level xv the date of sediments at depth x in this section is determined by solving the equation

210 where AA(xltx)is the Pb inventory between xx and x.

Fig. 8 shows results obtained by applying this method to a sediment core from Norrviken (). A highly irregular 210Pb activity versus depth profile with very low unsupported activities suggested rapid sediment accumulation, par- ticularly during the past 40 - 50 years. Well defined 137Cs and 241Am peaks from weapons test and Chernobyl fallout identified the 1986 and 1963 depths and showed that the core did in fact contain a good record of inputs to the lake. Discrepancies between the 137Cs dates and simple CRS model 210Pb dates (dashed line) show that the disturbances must have been associated with ma- jor variations in the 210Pb supply. Calculations based on the 137Cs dates indi- cated an 8-fold increase in sedimentation rates during 1944 - 62, with a 5-fold increase in the 210Pb flux. The solid line shows 210Pb dates calculated using the above equations. The dotted line shows the calculated sedimentation rates.

9 Variations in 226Ra Activity

In many of the early studies, in which 210Pb was determined by alpha spec- trometry, it was often assumed that 226Ra activity was uniform throughout the core and that supported 210Pb could be estimated from total 210Pb activity at depths where unsupported activity could safely be assumed negligible. Direct 226Ra measurements, where carried out, were usually limited in number. One of the major advantages of 210Pb assay by gamma spectrometry has been the determination of 226Ra in each sample, usually via the 295 keV and 352 keV emissions from the daughter radionuclide 214Pb following 3 weeks equilibration in a container sealed against 222Rn escape.

19 STUK-A145

0

10 3.5 20 — Dates CRS 210Pb Dates CRS Pb Dates (Composite model) 30 CRS Sedimentation Rates 3.0

40 2.5 c/~ 50 § o 60 2.0 Q. 2 CD

90

100

110

120 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 Age (y)

Figure 8. Depth and sedimentation rate versus age relations for a sediment core from Norrviken, Sweden, showing both the raw CRS model dates (-•—) and the corrected dates calculated using the piecewise CRS model. The elevated 210Pb flux during the period of rapid sedimentation was calculated using the 1954, 1963 and 198613TCs dates.

Although variations in 226Ra activity with depth are in most cases quite small, this is not always the case and there are situations where they can result in significant errors in the estimation of unsupported 210Pb activity. Fig. 9 illus- trates an extreme example from Dalvatn, northern Norway (Norton et al. 1992), where the sediment record contained a discrete layer (at 3 - 6 cm depth) in which 226Ra activity was an order of magnitude higher than the normal value of 30 - 50 Bq kg'1. In such cases, 222Rn diffusion, sustained over a number of decades, may result in significant differences between the supported (in, situ generated) 210Pb activity and the 226Ra activity. Calculations using a simple diffusion model show that supported 210Pb activity is reduced in the 226Ra rich sediments and increased in the adjacent layers. Unsupported 210Pb was esti- mated by subtracting the calculated values from the total 210Pb. CRS model calculations based on the results indicated that the 226Ra rich sediments were associated with an episode of rapid sedimentation around 1910.

20 STUK-A145

400

^Ra Supported "°Pb

cr 250

O Q. Q. 3 ISO

8 10 12 14 16 18 Depth (cm)

Figure 9. Supported 210Pb and S2SRa in a sediment core from Dalvatn, Norway (Norton et al. 1992). The i2sRa activities (-—-) were determined by direct gamma assay. The supported 210Pb activities ( ) were calculated using a simple diffusion model.

10 Conclusion

It is unlikely that dating by 210Pb will ever be a totally routine procedure. In the relatively closed system of a lake and its catchment where the CRS model has an underlying theoretical basis, potential complexities in transport pro- cesses are such that neither of the simple models (CRS and CIC) can be pre- sumed without independent validation. The situation is considerably more difficult in marine environments where there is no a priori reason to suppose

21 STUK-A145

that 210Pb supply rates are governed by any simple relationship. In these cir- cumstances the CRS model might just give the best results because of the ro- bustness of the dating parameter. Apart from the fact that integration is a relatively reliable numerical procedure, it also has the effect of smoothing out minor irregularities.

Since validation can at best be achieved at a small number of points, good quality results depend on the adoption of good numerical procedures that take account of the strengths and weaknesses of the particular dating model being used. The CIC model will be prone to large errors if there is significant mixing of the surficial sediments. The CRS model may give nonsensical results if there are gaps in the sediment record. Potential dating problems may be high- lighted by routine calculation of 210Pb dates using both models. Comparisons between the two sets of dates will help determine appropriate procedures for correcting or improving the initial calculations. These procedures, some of which are outlined above, are best carried out in the light of a good under- standing of the potential transport processes controlling the supply of fallout 210Pb to the sediment record.

References

Appleby, P.G. & F. Oldfield, 1978. The calculation of 210Pb dates assuming a constant rate of supply of unsupported 210Pb to the sediment. Catena, 5:1-8.

Appleby, P.G., F. Oldfield, R. Thompson, P. Huttunen & K. Tolonen, 1979. 210Pb dating of annually laminated lake sediments from Finland. Nature, 280:53-55.

Appleby, P.G. & F. Oldfield, 1983. The assessment of 210Pb data from sites with varying sediment accumulation rates. Hydrobiologia, 103:29-35.

Appleby, P.G, N. Richardson & P.J. Nolan, 1991. 241Am dating of lake sedi- ments. Hydrobiologia, 214:35-42.

Appleby, P.G. & F. Oldfield, 1992. Application of 210Pb to sedimentation stud- ies. In: M. Ivanovich & R.S. Harmon (eds.), -series Disequilibrium: Applications to , Marine & Environmental Sciences, Oxford University Press, 731-778.

22 STUK-A145

Appleby, P.G. & J.T. Smith, 1993. The transport of radionuclides in lake- catchment systems. Proc. UNESCO Workshop on Hydrological Impact of Nu- clear Power Plant Systems, Paris, 1992, 264-275.

Appleby,P.G., V.I. Jones & J.C. Ellis-Evans, 1995. Radiometric dating of lake sediments from Signy Island (maritime Antarctic): evidence of recent climatic change. J. Palaeolimn., 13:179-191.

Appleby.P.G., 1991 Sediment records of fallout radionuclides and their appli- cation to studies of sediment-water interactions. Water, Air & Soil Pollution, 99: 573-586.

Cambray, R.S., P.A. Cawse, JA. Garland, JAB. Gibson, P. Johnson, G.N.J. Lewis, D. Newton, L. Salmon & B.O. Wade, 1987. Observations of radioactivity from the Chernobyl accident. Nuclear Energy, 26:77-101.

Dominik, J., D. Burns & J.-P. Vernet, 1987. Transport of environmental radio- in an alpine watershed. Earth. Plan. Sci. Lett., 84:165-180.

Lewis, D.M., 1977. The use of 210Pb as a heavy metal tracer in Susquehanna River system. Geochim. Cosmochim. Acta, 41:1557-1564.

Norton, S.A., A. Henriksen, P.G. Appleby, L.L. Ludwig, D.V. Verault, & T.S. Traaen, 1992. Trace metal pollution in Eastern Finnmark, Norway, as evi- denced by studies of lake sediments. Norsk Institut for Vannforskning Report 487/927. 42 pp.

Oldfield, F. & P.G. Appleby, 1984. Empirical testing of 210Pb dating models. In: E.Y. Haworth and J.G. Lund (eds.), Lake Sediments and Environmental His- tory, Leicester Univ. Press, 93-124.

Pennington, W., R.S. Cambray & E.M. Fisher, 1972. Observations on lake sediments using fallout 137Cs as a tracer. Nature, 242:324-326.

Pennington, W., R.S. Cambray, J.D. Eakins & D.D. Harkness, 1976. Ra- dionuclide dating of the recent sediments of Blelham Tarn. Freshwater Biol- ogy, 6:317-331.

23 STUK-A145

Robbins, J.A., J.R. Krezoski & S.C. Mozley, 1977. Radioactivity in sediments of the Great Lakes: post-depositional redistribution by deposit feeding organisms. Earth Planet. Sci. Lett., 36:325-333.

Robbins, J.A., 1978. Geochemical and geophysical applications of radioactive lead. In: J.O. Nriagu (ed.), Biogeochemistry of Lead in the Environment. El- sevier Scientific, Amsterdam, 285-393.

Scott, M.R., R.J. Rotter & P.F. Salter, 1985. Tranport of fallout plutonium to the ocean by the Mississippi River. Earth. Plan. Sci. Lett., 75:321-326.

24 STUK-A145

THE USE OF FLY-ASH PARTICLES FOR DATING LAKE SEDIMENTS Neil L. Rose

Environmental Change Research Centre University College London, 26 Bedford Way London WC1H0AP

Abstract

Fly-ash particles comprise two particle types, spheroidal carbonaceous parti- cles (SCP) and inorganic ash spheres (IAS). In lake sediments, these particles form an unambiguous record of atmospheric deposition of industrial pollu- tants. The main temporal trends of these sediment records have been found to be remarkably consistent over wide geographical areas and these features can be used to ascribe dates to cores undated by other means. SCP are the main particle type used in this way as they are easier to extract and enumerate. IAS are morphologically similar to some natural particles (volcanic, meteoritic) and therefore show a natural background concentration at all pre-Industrial sedi- ment depths. Apart from this background temporal trends are similar to SCP and IAS could be used for dating in the same way. Gaining an indication of fuel-type change, either from IAS:SCP ratio or chemical characterisation of SCP enables additional dates to be added to sediment cores through compari- son with documentary evidence. In the future, as 210Pb becomes progressively unable to date the post-Industrial sediment record, dating using fly-ash tech- niques may become increasingly important.

Introduction

As lake sediments accumulate they store an historical record of atmospheric pollutants deposited on to the lake and its catchment. These pollutants include fly-ash particles produced from the high combustion of - fuels. Fly-ash consists of two types of particle, spheroidal carbonaceous parti- cles (SCPs) (Fig. 1) formed from the incomplete combustion of the fuel, and inorganic ash spheres (IAS) (Fig. 2) formed by the fusing of inclusions. Fly-ash particle analyses on lake sediment cores and surface sediments

25 STUK-A145

Figure 1. Spheroidal carbonaceous particle (SCP) from the incomplete com- bustion of fossil-fuel eoctracted from a lake sediment core.

Figure 2. Inorganic ash sphere (IAS) containing enclosed plerospheres. Such particles are formed by inconsistencies in melting the mineral grain.

26 STUK-A145

provide a temporal and spatial record of the impact of fossil-fuel combustion on a region.

Although physically fragile, SCP are chemically robust and so strong mineral acids (HF, HNO3 & HC1) can be used in their extraction from sediments (Rose, 1994). In addition, they are morphologically characteristic, have no natural sources and therefore in lake sediments form unambiguous indicators of at- mospheric deposition from industrial sources. IAS, however, are mainly com- posed of aluminosilicates and therefore have to be extracted using less rigor- ous means. This results in a less complete extraction and consequently IAS are more difficult to enumerate than SCP. They are also morphologically similar to volcanic microspherules and micro-. Most palaeolimnological studies have therefore used SCP in preference to IAS although some additional infor- mation is obtainable from the use of both particle types. Studies on the chemi- cal characterisation of fly-ash particles (for source apportionment studies) also utilise SCP rather than IAS. This is because the SCP chemistries reflect the fuel from which they were derived whereas the IAS chemistries reflect their mineral origins rather than being fuel-specific. Results from particle charac- terisation can also be used to ascribe additional dates to sediment cores.

Background

The first carbonaceous particle profile for lake sediments was produced by Griffin and Goldberg (1981). This was a profile of percentage by weight for a core taken from Lake Michigan using infra-red techniques on the >38p.m fraction. A carbon record was present at all sediment levels due to domestic wood burning and natural fire sources and the industrial fossil-fuel record began at around 1900. This was shown by an increase in carbon levels and a change in particle morphology towards a more spherical nature. Carbon con- centrations were shown to double in the periods 1900-1930 and 1930-1960 followed by a decline. It was concluded that particles found in lake sediments produced by the high temperature combustion of fossil-fuels were characteris- tically spheroidal, and that the change in concentrations of this type of particle reflected the changes in fossil-fuel consumption and the introduction and im- provement of particle-arresting techniques at the combustion sources.

The first European profiles of SCP concentrations were produced for Swedish lakes (Renberg & Wik, 1984, 1985a & 1985b). These were also seen to reflect

27 STUK-A145

the historical development of coal and oil burning with the beginning of parti- cle deposition seen to be around the middle of the 19th century due to in- creased coal burning in the Industrial Revolution. This was followed by a steady rise in concentration and, after the Second World War, a sharp increase in particle concentration caused by increased oil consumption. A peak concen- tration occurred around 1970 followed by a general decline in particle concen- tration. This pattern emerged so consistently that it became possible to use these SCP profiles as a dating method for sediment cores in Sweden (Renberg & Wik, 1985b).

The first SCP profile for a U.K. lake sediment core was produced for the Round Loch of Glenhead by Darley and was shown to be correlated strongly with the of fossil-fuel combustion in the U.K. (Darley, 1985). Rose et al. (1995) later used this approach to (a) date lake sediment cores from the U.K. and Ireland, and (b) to show that there are regional differences within countries which must be taken into account when attributing dates. SCP profiles have now been produced extensively within Europe and the main temporal patterns remain consistent from Svalbard to the Sierra Nevada in southern Spain and from western Ireland to the Tatra Mountains in Slovakia and Poland (Rose et al., in prep.) SCP profiles have also been determined for lake sites in many parts of the world including China, USA, Siberia and Tasmania and the prin- ciples of dating apply equally well in all locations.

SCP Profiles

The technique of dating sediment cores using fly-ash particle concentration profiles relies, in the first instance, on dating the first few profiles in a region using independent dating means such as 210Pb or varve counting. Once the main features of the particle profile have had reliable dates attri- buted to them then subsequent profiles in that region can themselves be used for dating.

In most SCP profiles there are three main features that are used for dating purposes (Fig. 3). These are the start of the particle record (A), the rapid in- crease in SCP concentration (B) and the SCP sub-surface maximum (C). These features occur in most sediment profiles, except where sediment accumulation rate or variability obscures them, and can usually be related to major events in industrial development or pollution control strategy in the main source areas.

28 STUK-A145

Figure 3. A schematic SCP profile showing the three profile features used for dating: A - the start of the particle record, B - the rapid increase in concentra- tion - defined by the change in slope of the two parts of the profile, C - the SCP concentration peak. (Figure taken from Rose et al., 1995)

It should be stressed that SCP are found in areas far removed from sources suggesting there is a possible 'northern hemispherical background' of SCP deposition. However, in most regions this background has regional or local deposition patterns superimposed upon it and it is these patterns that are most useful in profile interpretation and dating.

29 STUK-A145

The start of the record

Using the U.K. as an example, the start of the SCP record is generally seen to occur in the middle of the 19th century and this is generally attributed to de- velopments in the Industrial Revolution and the combustion of larger quanti- ties of coal at higher . The Industrial Revolution in Britain began in the early years of the eighteenth century with the development of a pump- ing engine by Savery, Papin and Newcomen, but it was not until the late nine- teenth century and the work of Faraday and Gramme that the industrial gen- eration of electricity finally became practicable. The first fossil-fuelled power station supplying electricity to the public was situated at Holborn Viaduct in London and began generation on 12th January 1882, and by the end of the century most large towns had their own electricity generating stations. The start of the SCP record in lake sediments therefore pre-dates the first power generating station and therefore during this period reflects the combustion of coal in industry as a whole. This may account for the variability in the start of the particle record both within and between regions, as local industries would have had a large impact on atmospheric deposition especially as industrial chimney heights were generally low at this time.

The dates for the start of the SCP record cover the largest range of the dating and so can be attributed to nothing more accurate than 'the mid-19th century' although the 1860s appear to be the most common decade. In Ireland, there is also a wide range of dates but these are generally later than for the U.K. and cover the late 19th and early 20th centuries. The same is true of many areas of Europe, but at remote sites the start of the record is often seen to be much later. This is probably because concentrations are much lower and so the start of the record at these sites becomes the date at which the SCP concentration first exceeds the detection limit of the technique rather than the true first date. In these cases the start of the record appears to show reasonable agree- ment with the rapid increase feature in more polluted lakes in the region.

Although such a wide margin may appear to limit the usefulness of this fea- ture as a dating tool, it should be noted that errors on rh dates for this pe- riod are frequently ± 15-20 years and sometimes more (e.g. 1896 ± 24 is the earliest date available for a Loch Uisge, Scotland core) making the dating margin for both of similar magnitude.

30 STUK-A145

The rapid increase feature

From the 1920s and 1930s onwards, the consumption of coal in the U.K. con- tinued to increase, and this accelerated after the Second World War due to a considerable increase in total energy demand. An abundance of cheap fuel-oil also became available at this time and led to the commissioning of the first oil- fired power station in the U.K. at Bankside, London in 1952. During this pe- riod of expansion and increased consumption in the power generation industry there was a substantial reduction in coal consumption in other industrial mar- kets and the railways. Particle emission legislation was also limited at this time as the Clean Air Act was not passed until 1956.

The SCP record for this period shows a rapid increase in most U.K. sediment cores and it seems probable that this mainly reflects the increased fuel con- sumption of the electricity generation industry. Although this feature usually dates to the 1950s or 1960s it also varies with region and is present in virtu- ally every analysed core both within U.K. and throughout Europe (Wik & Nat- kanski, 1990; Cameron et al, 1993; Toro et al, 1993, Rose et al., in prep.).

The sub-surface maximum

Coal consumption throughout the U.K. continued to increase until the late 1970s. Since this time, consumption has remained reasonably steady in Eng- land and Wales (Electricity Council, 1988) although in Scotland it decreased from about 8 million tonnes in the late 1970s to about 3 million tonnes in 1985 mainly due to the reduction in heavy industry. Oil consumption throughout the U.K. increased from the 1950s until 1973 when the massive price increases imposed by OPEC (Organisation of Petroleum Exporting Countries) and the oil shortage caused by the Arab oil embargo of 1973-1974, caused consumption to decrease again.

Starting in the 1930s, the Central Electricity Board (C.E.B.), (later to become the Central Electricity Authority (C.E.A.) and then the Central Electricity Generating Board (C.E.G.B.)) continued a policy of creating fewer, larger and more efficient power stations. In addition, following the Clean Air Act of 1956 (and subsequent amendments and extensions) all new power stations were fitted with dust extraction equipment. The increase in particle removal effi- ciency and the implementation of more rigorous pollution control legislation meant that despite the continued increase in fuel consumption the increase in

31 STUK-A145

particulate emissions slowed down and from the mid 1970s started to de- crease.

The particle record in lake sediments for the post-War period shows a contin- ued increase in concentration to a peak, usuaLy dating to the late 1970s / early 1980s followed by a decline in SCP concentration to most recent times. This recent reduction in the number of SCPs reaching lake sediments throughout the British Isles is therefore probably due to a combination of two factors. Firstly, the small decrease in fossil-fuel consumption (probably more signifi- cant in Scotland than elsewhere) and secondly, improvements in the removal of participates from flue gases and to the more widespread use of these tech- niques, caused by the implementation of more stringent air pollution legisla- tion.

This last feature is the best defined of the three and in the U.K. can normally be ascribed with an accuracy of ± 3-4 years (Rose et al., 1995). As with the other two features the date for the SCP peak varies between the different re- gions. In this particular case, however, the range of dates is narrower and the difference between regions is only a few years. Most recent cores in Europe show this feature except those with very low sediment accumulation rates (Cameron et al., 1993, Rose et al., in prep.) Regional dating variations for all profile features are to be expected due to variations in emission across Europe. However, how localised these variations are has not been studied except in parts of the U.K. and Ireland (Rose et al., 1995).

Further, more detailed, dating using SCP data may be possible using cumula- tive SCP profiles. This is especially the case where the dates to be transferred are between multiple cores taken from a single lake. Although only limited data exist, cumulative SCP profiles appear to show much closer agreement than multiple SCP concentration profiles and it may be that more dates can be ascribed to sediment profiles between lakes using this approach in addition to using the ^traditional' three SCP features (Rose et al., in press).

IAS profiles and IAS:SCP ratios

As IAS are also produced from fossil-fuel combustion, the post-Industrial sediment record shows similar trends to those of SCP. However, there are two points of note when comparing the profiles. First, IAS shows a continuous

32 STUK-A145

background concentration at all pre-industrial levels, and second, there can be significant differences in concentration depending on where the sediment core is taken with respect to deposition sources (e.g. Fig. 4).

0-\ -1987 -1979 -1966 5- -1952 -1938 10- - 1925 -1912

15- -1897

t 20-

25-

30-

35-

40J 5000 10000 15000 20000 25000 30000

Particle concentration (qDM"')

Figure 4. IAS and SCP profiles for Loch Teanga in the Outer Hebrides, Scot- land. • - SCP, x = IAS. (Figure taken from Rose, 1996)

33 STUK-A145

The continuous background is due to naturally produced IAS either from vol- canic microspherules (Lefevre et al., 1986) or possibly micrometeorites (Wohletz & McQueen, 1984). These appear morphologically identical to the IAS under the light microscope and the particle types cannot be distinguished in this way (Rose, 1996). Meteoritic spheres could be isolated by chemical means but volcanic particles may still not be distinguishable. If this back- ground concentration is removed from all sediment levels then the IAS profile begins at the same date as the SCP profile. In a country where the dominant fuel is coal (such as the U.K.) then the other profile features are similar to those of SCP and dates can be attributed in the same way.

The second point of interest is the difference in sediment concentration of IAS and SCP. For many Scottish lochs, IAS concentrations are higher than SCP concentrations with a surface ratio of these particle concentrations (the IAS:SCP ratio) of greater than 5.0. For sites in the south-east of England how- ever, the concentrations of both particle types are approximately equal at the surface. For one site in London, on Hampstead Heath, there is a ratio mini- mum in the mid 1970s but prior to this there is a steady decrease in ratio since the 1920s.

Fossil-fuel combustion in Scotland at the time these cores were taken was al- most solely coal except one oil-fired power station situated at Peterhead near Aberdeen, and another at Inverkip in Glasgow which closed in 1987. It is known that pulverised coal fly-ash typically contains greater than 95% inor- ganic material and so the fly-ash particles emitted and deposited in Scotland should be predominantly IAS rather than SCP. The south-east of England however, has a much higher proportion of oil-fired stations and within a 40km radius of London there is 5,700 MW capacity of oil-fired electricity generation and 4,790 MW of coal. Oil fly-ash is predominantly composed of SCP and so it might be expected that the IAS:SCP ratio at a London site would be much lower than those from Scotland. The period between the mid 1950s and the oil crisis of the mid 1970s was a time of rapid expansion in the use of oil in the U.K. and it is interesting to note that this is the only period for the Hampstead Heath site where SCP concentrations are greater than those of IAS (Fig. 5).

An alternative explanation may be that as IAS generally have a smaller di- ameter than SCP they can travel further in air streams and it is distance from source that is the main criterion explaining IAS and SCP variability. How- ever, Loch Tinker, only 45km from Glasgow shows a high ratio throughout the

34 STUK-A145

Figure 5. IAS.SCP ratio and particle characterisation data for a sediment core taken from Hampstead Heath, north London, U.K. (Figure taken from Rose, 1996)

35 STUK-A145

profile (> 5.0) and therefore it seems unlikely that distance from source alone explains the difference between the IAS:SCP ratios. Regional fuel-use is there- fore the most likely factor explaining this difference and therefore it may be possible to use IAS:SCP ratios to gain an idea of the relative impacts and tem- poral changes in the combustion of coal and oil on atmospheric depositions at a lake site in the U.K. This information, in combination with a historical docu- mentary knowledge of combustion within that region, allows the possibility of ascribing additional dates to the sediment profile.

Dates from SCP characterisation

All the above dating approaches have used sediment particle concentration data only, but further information is available from SCP if they are character- ised to their fuel-type using their elemental chemistries (Rose et al., 1996). A technique has now been developed which allows sediment extracted SCP of coal, oil, peat, brown coal and oil shale origin to be allocated with high levels (>80%) of confidence. This approach is more accurate at determining fuels than the particle ratio technique and allows historical records of specific fuels to be used. It therefore provides more specific information from a sediment core. However, for regions such as the U.K., where there are only two main fuel-types, coal and oil, the two approaches provide similar information, albeit that the characterisation technique is more reliable. The advantage of the chemical approach becomes more obvious when making comparisons between countries. For example, particle ratios in Estonia will be similar to those in the U.K. despite the fuels being very different, and this is because oil shale, like coal, produces large numbers of IAS compared to SCP. In regions where such fuel-types overlap the particle ratio approach would be of limited use although the chemical characterisation approach may still be useful.

In areas where there has not been significant fuel-type changes through time, both particle ratio and chemical characterisation techniques will be of limited use. The 'traditional' profile dating approach will, however, still be useful.

Conclusions & the future

The use of fly-ash particle profiles allows a cheap and quick method for allocat- ing dates to sediment cores, but allocatable dates are few and accuracy is per- haps not as good as more established techniques such as 210Pb and varve

36 STUK-A145

counting. However, in sites without varves the future of SCP dating is perhaps more assured.

Over the last 10-15 years rb dating has proved very useful in many palaeo- limnological studies concerned with surface water acidification and pollutant deposition. This is because the half-life of 1 Pb enables a chronology to be con- structed for the past 100-150 years, spanning the major part of the period since the Industrial Revolution. In future years this will cease to be the case as the early years of this period progressively fall below the 210Pb dating horizon. Since there is no indication that SCPs are mobile within the sediment column (except possibly for some surficial mixing in the initial stages) or that they degrade with time, the start of the particle record may become one of the few ways of dating the early part of this period. The other features of the SCP profile (B and C in Figure 1) will be of increasing value in providing a rela- tively quick and cheap means of determining approximate core dates that can be used on their own, or as a means of increasing the reliability of more de- tailed Pb chronologies. Therefore unless new techniques are developed, fly- ash dating will become increasingly more important in future years and, in combination with 210Pb, will ensure that the entire post-Industrial period re- mains accessible to palaeolimnologists through the use of the lake sediment record.

References

Cameron, N.G, Rose, N.L., Appleby, P.G., Schnell, 0.A., Battarbee, R.W., Pat- rick, S.T. and Flower R.J., 1993. ALPE I: Palaeolimnology. Environmental Change Research Centre, University College London, Research Report No. 4.

Darley, J., 1985. Participate soot in Galloway lake sediments. Its application as an indicator of environmental change and as a technique for dating recent sediments. Dept. of Geog. University College London. B301 Project.

Electricity Council, 1988. Handbook of electrical supply statistics.

Griffin, J.J. and Goldberg, E.D., 1981. Sphericity as a characteristic of solids from fossil-fuel burning in a Lake Michigan sediment. Geochimica et Cosmo- chimica Acta, 45: 763-769.

37 STUK-A145

Lefevre, R., Gaudichet, A and Billon-Galland, M.A., 1986. Silicate micro- spherules intercepted in the plume of Etna . Nature, 322: 817-820.

Renberg, I. and Wik, M., 1984. Dating of recent lake sediments by soot particle counting. Verh. internat. Ver. Iimnol., 22: 712-718.

Renberg, I. and Wik, M., 1985a. Carbonaceous particles in lake sediments - Pollutants from fossil fuel combustion. Ambio, 14: 161-163.

Renberg, I. and Wik, M., 1985b. Soot particle counting in recent lake sedi- ments. An indirect counting method. Ecological Bulletin, 37: 53-57.

Rose, N.L., 1994. A note on further refinements to a procedure for the extrac- tion of carbonaceous fly-ash particles from sediments. Journal of Paleolimnol- ogy, 11: 201-204.

Rose, N.L., 1996. Inorganic fly-ash spheres as pollution tracers. Environmental Pollution, 91: 245-252.

Rose, N.L., Harlock, S., Appleby, P.G. and Battarbee, R.W., 1995. Dating of recent lake sediments in the United Kingdom and Ireland using spheroidal carbonaceous particle (SCP) concentration profiles. Holocene, 5: 328-335.

Rose, N.L., Juggins, S. and Watt, J., 1996. Fuel-type characterisation of carbonaceous fly-ash particles using EDS-derived surface chemistries and its application to particles extracted from lake sediments. Philosophical Transactions of the Royal Society (Series A), 452: 881-907

Rose, N.L., Harlock, S. and Appleby, P.G., (in preparation). The spatial and temporal distributions of spheroidal carbonaceous fly-ash particles (SCP) in the sediment records of European mountain lakes.

Rose, N.L., Harlock, S. and Appleby, P.G., (in press). Within-basin profile variability and cross-correlation of sediment cores using the spheroidal carbo- naceous particle record. Journal of .

Toro, M., Flower, R.J., Rose, N.L. and Stevenson, A.C., 1993. The sedimentary record of the recent history in a high mountain lake in central Spain. Verh. Internat. Verein. Limnol., 25: 1108-1112.

38 STUK-A145

Wik, M. and Natkanski, J., 1990. British and Scandinavian lake sediment rec- ords of carbonaceous particles from fossil-fuel combustion. Philosophical Transactions of the Royal Society of London, B327: 319-323.

Wohletz, K.H. and McQueen, R.G., 1984. Volcanic and stratospheric dust-like particles produced by experimental water-melt interactions. , 12: 591- 594.

39 STUK-A145

SEDIMENT MIXING MODELS IN THE INTERPRETATION OF 2i0Pb CONCENTRATION PROFILES IN CONTINENTAL SHELF SEDIMENTS OF EASTERN CANADA

Mulsow, S.1'\ B. Boudreau1 and J. N. Smith2

1Dalhousie University, Oceanography Department, Halifax, Canada 2Bedford Institute of Oceanography, Marine Chemistry Div. Dartmouth, Canada * Current address: IAEA-MEL Monaco

Abstract

We have investigated the effects of porosity on bioturbation rates and their influence on the organic carbon decay constants at 5 stations scattered on the Nova Scotia Shelf (NSS) and in the Gulf of Saint Lawrence (GSL) as part of the Canadian JGOFS Program.

Bioturbation rates (DB: biodiffusion coefficients) were determined from excess 210Pb concentration profiles in sediment cores. Three different mixing models were applied to determine mixing coefficients at each one of the study sites. The three models were distinguished by the way they treated porosity: con- stant porosity (MCP), porosity mixed (IEM) and porosity non-mixed (IAM). NSS profiles that were similar with the three models yielded DB values rang- ing from 0.37 to 2.58 cmfy-Vith no obvious correlation to season of sampling. GSL cores taken during the Summer 94 showed intense mixing in the upper 8 cm while mixing rates in the next 7 - 12 cm were at least an order of magni- tude smaller. Cores taken in Winter 93 did not show any well mixed layers. Organic carbon profiles, measured simultaneously at each station, exhibited exponential-like gradients with depth. Given the above DB values, a G-type model produced decay constants (k) of 0.003 y-1to 0.012 yJfor NSS sediments; k values for GSL sediments ranged from 0.003 y1 to 0.03 y1. Finally, bioturba- tion mixing rates were similar when MCP and IAM models were used. How- ever the IEM model revealed well mixed layers that were not shown with MCP and IAM and also enhanced well mixed layers present on some of the

40 STUK-A145

210Pb profiles. Organic Carbon decay constants were mostly independent of the mixing model used.

Introduction

Most of the particles reaching the seafloor are displaced several times by ani- mals before they become part of the sedimentary record. This is true not only for detrital particles but also for both labile and refractory organic particles. The rates at which recently deposited organic particles are decayed are strongly influenced by biological mixing of solids and solutes within the bio- turbation zone (Berner, 1980). Thus every determination of organic carbon decay constant (a-type sensu, Berner, 1980) needs to take into account biologi- cal mixing (biodiffusion coefficients, DB). Biological mixing has generally been estimated from the best fit curves of different models to naturally occurring radionuclides in sediments such as 210Pb concentration profiles [Goldberg and Koide (1962), Guinasso and Schink (1975), Nozaki et al.,(1977), Aller, 1977 and Aller, et al., (1980)]. Solid and fluid bioturbation processes take place at dis- tinctively different rates because of sediment irrigation, as it was shown by Aller (1977) for near shore sediments. Porosity, a measure of interstitial water volume, and tracer gradients are closely interrelated (Berner, 1980). The more complete theoretical models for tracer conservation in bioturbated sediments explicitly account for the existence and effects of porosity changes (gradients, or more precisely changes in the solid volume fraction gradients) [Schink, and Guinasso (1978), Berner, 1980; Aller, (1982) Officer and Lynch (1982); Chris- tensen (1982); Boudreau (1986) and Mulsow et al., (in press)].

As part of the Canadian JGOFS Program, the objective of this research was to 1) investigate the effects of porosity gradients in the modelling of bioturbation mixing rates and 2) to test if the differential model treatments of bioturbation rates would have an effect on the calculations of organic carbon decay constant within the bioturbation zone. To determine the effects of porosity three differ- ent mixing models were applied to 210Pb sediment concentration profiles. These models differed in the way each one of them treated porosity: constant porosity (MCP), mixed porosity (IEM) and porosity non-mixed (IAM). A G- model (Berner, 1980) was used for organic carbon decay constant calculations (k values).

41 STUK-A145

Methods

Stations and sampling

Sediment cores were collected from 5 stations located in the North East of Canada: 3 stations in the Gulf of Saint Lawrence and 2 on the Continental shelf off Nova Scotia (Fig. 1). Stations 1, 2 and 3 were sampled in Winter 93 and Summer 94, while station 4 was visited 3 times (Summer 93, Winter 93 and Summer 94) and station 5 in Summer 93 and Summer 94. Sediment sam- ples were collected with a multicore at all stations with the exception of sta- tions 4 and 5 during the Summer 93, where sediment cores were sub-sampled from box cores. In that occasion a Plexiglas corer was used with the exact di- mensions of the corers of the multicore. On board cores were sliced every half a centimetre from surface to 1 cm, and every 1 cm thereafter whenever possi- ble (i.e. no cracks present).

70° $5° 60°

50° 50°

45° -45°

70° 65° 60°

Figure 1. Map of eastern Canada showing the location and depth range of the 5 Canadian JGOFS Stations.

42 STUK-A145

Porosity

At each interval, 3 replicate sub-samples were collected to determine bulk density from the difference between wet and dry (105°C for 48 h) weights. A salt correction was applied based on the bottom salinity measured at each sta- tion (STD casts).

Organic Carbon

At each station, surface sediment samples were dried (105° C for 48 h) and total organic carbon was determined by high temperature oxidation (CHN Analyser). The organic carbon fraction was calculated by subtracting the inor- ganic carbon content from total organic carbon. The inorganic carbon fraction was measured on pore water of surface sediment samples collected simultane- ously at the same sampling sites. All the organic carbon values were corrected for salinity measured on the overlying water at each station.

210Pb activity

Sediments for radiogenic analyses were kept at 4° C during transportation to the laboratory and subsequently freeze dried and homogenised. 210Pb was de- termined by counting 210Po and using as tracer 208Po, deposited onto nickel disks, while 226Ra supported levels were determined on acid digested sediment samples using radon gas emanation technique (Mathieu, et al.,1988). Samples were counted until errors were less than 5% of the samples' activity.

Mixing Models

The first model used for biological mixing assumes constant porosity (MCP, mean constant porosity; top diagram, Fig. 2) and can be formulated as:

where x is the depth relative to the sediment water interface (cm), C is the activity or concentration of the tracer (dpm cm'3), DB is the biodiffusion coeffi- cient for constant porosity mixing (cm2 y1) and k is the decay constant for 210Pb« (Boudreau, 1986).

43 STUK-A145

mean constant porosity o • o °o°oo°o • o o • box a °o o °o mixing • OS boxb o v o« •:":*•*: *o o • o 9a=(Pb 93=% Intraphase mixing o o o mixing o • o O- O • O ~ o o o O ajB ^0 9 •*••*•• C* Cr^ O

(Pb

= (Da>(ph CPa CPh I Q I |J 1 Q T LJ

Figure 2. Schematic representation of sediment mixing in the three possible modes studied in this paper. The top diagram illustrates a conventional case of mixing where porosity is constant. The top layer, box a (white circles) and bot- tom layer, box b (black circles), which represent the particle characteristics within the sediment column, have the same initial porosity which remains un- changed after mixing of particles takes place. The middle diagram illustrates the case of intraphase mixing. The porosity and particles are initially different in both boxes; after mixing only the particle types are homogenised (porosity remains unchanged). The bottom diagram illustrates Interphase mixing. The initial conditions are identical to the ones of intraphase mixing, but this time, after mixing, both porosity and particle-type gradients are eliminated by com- plete mixing. Detailed explanation and analysis of the models can be found in Mulsow, et al. (in press).

44 STUK-A145

The second model states that bioturbation does not mix solids with fluid to eliminate porosity gradients (Thomson, et al., 1988). The flux of a solid species is restricted by the fractional area occupied by solids (i.e. l-

where j is the porosity and cps = (1-cp) is the solid volume fraction.

The third mixing model allows mixing of solids and fluid together to remove porosity gradients (Nozaki, 1977) and Anderson, et al., (1980) (bottom dia- gram, Fig. 2). It has been called Interphase mixing (IEM) by Boudreau (1986), and can be formulated as:

D

A detail description and analysis of the three models used here are given in Mulsow, et al. (1997).

Organic Carbon G-Model

Organic decomposition accompanying sulphate reduction takes place within the bioturbation zone. As oxygen profiles measured in sediment reach undetectable values at just a few millimetres within the sediment column (personal observation), bioturbation of labile organic matter follows a biodiffu- sion model similar to those mentioned above with a fixed value of DB. Steady state diagenesis can be assumed if mean annual concentrations of labile or- ganic matter and mean annual sulphate reduction rates are used (sulphate reduction rates were measured on the same study sites). With this assumption the appropriate diagenetic equation for reactive organic carbon is:

dx2 where G is the concentration (in mass per unit mass of total solids) of organic carbon which can readily be metabolised within the bioturbation zone, DB is

45 STUK-A145

the biodiffusion coefficient for solids (calculated with MCP, IEM and IAM), and k is the organic carbon decay constant (y1).

Regression Models

Regressions for the constant porosity (MCP) and Interphase (IEM) models and organic carbon were calculated with built-in routines of Kaleidagraph. Re- gressions for the intraphase model (IAM) were done with FORTRAN routines from ACM approved MINPACK package.

Results and Discussion

Porosity

Solid volume fraction gradients (q>s) were observed for all the sampled cores and followed a gradual increase with depth before to reach a constant, except Station 2 in Summer 94 and Station 4 in Summer 93 which may represent a small burrow or recent bioturbation event (Fig. 3). Values for porosity (cp) were no less than 2 for any of the stations, and as great as 3 for Station 1 in Sum- mer 94. These results suggest that porosity can not be well represented by a simple mean value.

2I0Pbacess activity and bioturbation coefficients

In 55% of the cases all three models agreed in explaining the observed 210Pbex data profiles and generated DB coefficients (Table I). However when using the Interphase mixing model (IEM) some 210Pbex profiles showed well-mixed top layer that where not shown with MCP or IAM (as shown in a representative case in Fig. 4) for the same data set. Similarly 45% of calculated DB values (Stations 1, 2, 3 and 4 in Winter 93 and Station 5 in Summer 94) were signifi- cantly higher (see Table I and Table II) when using the IEM model. Unfortu- nately, the data and analysis presented here can not in themselves tell us whether Interphase or Intraphase mixing or a combinations occurs at our study sites. Only in situ observations or microcosm experiments could shed some light on that question.

46 STUK-A145

Solid Volume Fraction (1- cp)

0.0 0.1 0.2 0.3 0.0 0.1 0.2 0.3 0 10- 20- 30- 40 -J 0 10 - 20 - 30- 40 -I 0 10 - f 20 - •o 30 - 40 -J Station 3

0.1 0.2 0.3 0.4 0.1 0.2 0.3 0.4 0.1 0.2 0.3 0.4 0 10 - 20 - 30 - 40 - 4 0 10- 20- 30 - 40 - Station 5 Summer 93 Winter 93 Summer 94

Figure 3. Plots of solid volume fraction (js) changes with depth in the upper few decimetres of sediment for all the stations and seasons. Points represent mean values of three replicates; error bars within drawn points.

47 STUK-A145

Station 5 (Summer 93)

210,Pb. ex 100 1 10 100 0.01 i

15 J mean constant porosity intraphase nixing interphase mixing

Figure 4. Selected plots of 210Pbex activity for Station 5 showing the best fit curves generated with each one of the models (MCP, IAM and IEM). The corre- sponding parameters are shown in Table I. The IEM model predicts a top well- mixed zone not shown with MCP and IAM models.

Organic Carbon decay constants, k

As most of the organic carbon profiles showed an exponential-like decrease with depth, we used a G-type diagenetic model (Berner, 1980) for the observed organic carbon data. Organic carbon decay constants (k) for NSS (Stations 4 and 5) ranged from 0.003 y1 to 0.04 y1 when using DB calculated with MCP, 0.003 y1 to 0.01 y1 when using DB -IEM and 0.004 y1 to 0.02 y1 when using DB -IAM (Table III); k-values for GSL (Stations 1, 2 and 3) ranged from 0.004 y1 to 0.03 y-1 when using DB - MCP, 0.004 y1 to 0.03 y1 iwhen using DB -IEM and 0.003 y1 to 0.006 y1 when using DB -IAM (Table III). There were no significant differences among the k-values calculated with DB coefficients generated using any of the three models. There was no seasonal pattern either in k-values for any of the stations. Organic carbon in the top 10-20 cm of sediment appeared to be reactive on a decadal scale. The organic carbon decay constants calcu-

48 STUK-A145

lated in this study are lower than expected for continental sediments; they are rather comparable to k-values for deep sea ocean sediment (Berner, 1980).

Table I. Comparison of calculated biodiffusion coefficients, (cm2 y1) for the different mixing models.

Sta- Sampling Mean-constant Intraphase Mixing Interphase Mixing tion time porosity (MCP) (1AM) OEM) num- range range range DB ber (cm) (cm) (cm) (cm y ) (cm y ) (cm y ) 1 winter 93 0 -10 0.13 ±0.04 0-8 0.15 ± 0.02 0-8 O22 ± 0 08^ 1 summer 94 0 -6 13.2 ±6.1 * 0-6 ' A^T\ * 6 •15 0.18 ±0.02 * 6-15 0.25 ± 0.03 2 winter 93 0 12 0.28 ±0.05 0-12 0.27 ± 0.03 0-12 0.61 ±0.12 2 summer 94 0 14 5.1 ± 2.0 * 0-14 s'Tw:-"- • 14 -21 0.08 ± 0.02 * 14-21 0.12 ±0.03 3 winter 93 0 12 0.08 ± 0.01 0-12 0.12 ±0.02 0- 12 0.11 ±0.03 3 summer 94 0- 5 12.1 ±6.2 * 0-5 5 16 0.68 ± 0.23 5-16 _2 29±H»1_ 4 summer 93 0- 18 0.47 ±0.11 0-11 0.23 ±0.13 0- 18 4 winter 93 0- 13 0.72 ±0.11 0-13 0.89 ± 0.09 „ ------0-5 5-11 a»£>.34 • 4 summer 94 0- 11 0.19 ±0.04 0-11 0.19 ±0.04 0-11 0.48 ±0.16 5 summer 93 0- 15 0.81 ±0.13 0-15 0.63 ± 0.09 •• ~ -- 0-6 2i22±^:0 ' ' - - - 6-15 5 summer 94 0- 12 0.37 ±0.12 0-12 0.35 ±0.05 0-12 0.95 ± 0.40 two layer model not developed tor intraphase mixing

49 STUK-A145

Table II. Statistical similarity in the mixing coefficients predicted by the Mean-Constant Porosity (MCP), Intraphase Mixing (JAM) and Interphase Mix- ing (IEM) models.

Station Sampling time Depth range Similarities* number (cm)

1 winter 93 0-10 MCP = LAM = IEM 1 summer 94 0-6 BMSJJM/ " $)_ 6-15 MCP = IEM 2 winter 93 0-12 MCP=^M = IEM 2 summer 94 0-14 14-21 3 winter 93 0-12 MCP^IAM = IEM 3 summer 94 0-5 wm&wm-

*Equality means 95% chance they are similar. 210 (t) IEM model has infinitely fast mixing on Pb time scale. (§) Agreement below 5 cm. (f) Agreement below 6 cm.

50 STUK-A145

Table III. Comparison of calculated organic carbon decay, k (y1), for the dif- ferent mixing models.

Station Sampling Mean Porosity Intraphase mixing Interphase mixing number time (MCP) (IAM) aEM)

1 winter 93 0.0036±0.0009 0.0032 ±0.0001 0.0047± 0.0008 1 summer 94 0.0023±0.0007 - - 2 winter 93 0.0026±0.0012 0.0048 ± 0.0012 0.0041± 0.0015 3 winter 93 0.0073±0.0022 0.0059 ± 0.0027 0.027± 0.0042 3 summer 94 0.0035±0.0013 0.38 ± 0.0017 - 4 summer 93 0.0072±0.0036 0.016 ±0.0082 0.0089 ±0.0028 4 winter 93 0.0033±0.0011 0.0035 ± 0.0026 0.0044 ±0.0010 4 summer 94 0.044±0.026 0.0600 ± 0.040 * 5 summer 93 0.018±0.0021 0.0110 ±0.0072 0.012 ± 0.0039 5 summer 94 0.0039±0.0012 0.0100 ± 0.0045 0.0028 ± 0.0006

* failure of numerical method, i.e., too stiff.

Table IV. Statistical similarity in the O.C. decay constants predicted by the Mean-Constant Porosity (MCP), Intraphase Mixing (IAM) and Interphase Mix- ing (IEM)models.

Station Sampling Similarities* number time

1 winter 93 MCP = IAM = IEM 1 summer 94 N/A 2 winter 93 MCP = IAM = IEM 3 winter 93 MCP = IAM = IEM 3 summer 94 MCP = IAM 4 summer 93 MCP = IAM = IEM 4 winter 93 MCP = IAM = IEM 4 summer 94 MCP = IAM = IEM 5 summer 93 MCP = IAM = IEM 5 summer 94 MCP = IAM = IEM

''Equality means 95% chance they are similar.

51 STUK-A145

Conclusions

Bioturbation mixing rates (mean values) based on 210Pbex profiles were similar when MCP and IAM mixing models were used. The Interphase mixing model (IEM) revealed well mixed layer that were not shown with MCP and IAM. It also enhanced well mixed layers on some of the 210Pbex profiles. This result shows that porosity gradients should be taken into account to fully understand the diagenetic processes taking place within the sediment column. To equate porosity with a single mean value (as with MCP model) might lead to over- simplification and misinterpretation of the mixing processes, as was evidenced in this study. In contrast, organic carbon decay constants were mostly inde- pendent of the mixing model.

Acknowledgements

This research was supported by the CJGOFS CSP Grant awarded by the Natural Sciences and Research Council of Canada. We wish to thank all the scientists and the technical group in the CJGOFS Benthic Processes Study who participated in the cruises; namely, B. Sundby, G. Desrosiers, A. Mucci, K. Juniper, and particularly D. Webb and T. Arakaki. Special thanks to Yutta Krieger for thorough editing of earlier versions.

References

Aller, R.C., 1977. The influence of macrobenthos on chemical diagenesis of marine sediments. PhD. dissertation. Yale University, New Haven.

Aller, R.C., Benninger, L.K., Cochran, J.K., 1980. Tracking particle-associated processes in nearshore environments by use of 234Th/238U disequilibrium. Earth and planetary letters, 47 (2): 161-175, Amsterdam.

Aller, R.C., 1982. The effects of macrobenthos on chemical properties of and overlying water. In: McCall, P.L. and Tevesz, M. (eds.), Animal- Sediment Relations, Plenum Press, New York, pp. 53-102.

Berner, R.A., 1980. Early Diagenesis. Princeton University Press. Princeton. 214 pp.

52 STUK-A145

Boudreau, B.P., 1986. Mathematics of tracer mixing in sediments. Spatially dependent, diffusive mixing. Am. J. Sci. 186: 161-198.

Christensen, E.R., 1982. A model for Radionuclides in Sediments Influenced by Mixing and Compaction. J. Geophys. Res., 87: 566-572.

Goldberg, E.D. and Koide, M., 1962. Geochronological studies of deep sea sediments by the ionium method. Geochim. et Cosmochim. Acta, 26: 417-450.

Guinasso, N.L. and Schink, D.R., 1975. Quantitative estimates of biological mixing rates in abyssal sediments. J. Geophys. Res., 80: 3032-3043.

Mathieu, G.G., Biscaye, P.E., Lupton, R.A., Hammond, D.E., 1988. System for measurement of 222Rn at low levels in natural waters. Health Physics, 55: 989- 992.

Mulsow S., B. Boudreau and J. N. Smith (in press). Porosity and gradients on sediment mixing models. Iimnol. Oceanogr.

Nozaki, Y., Cochran, J.K., Turekian, K.K., Keller, G., 1977. Radiocarbon and 210Pb distribution in submersible-taken deep-sea cores from project FAMOUS. Earth Planet. Sci. Lett., 34/2: 167-173, Amsterdam.

Officer, C.B. and Lynch, D.R., 1982. Interpretation procedures for the deter- mination of sediment parameters from time dependent flux inputs. Earth Planet. Sci. Lett., 61: 59, Amsterdam.

Schink, D.R. and Guinasso, N.L., 1978. Redistribution of dissolved and ad- sorbed materials in abyssal marine sediments undergoing biological stirring [bioturbation]. Am. J. Sci., 278/5: 687-702.

Thomson, J., Colley, S. and Weaver, P.P.E., 1988. Bioturbation into a recently emplaced deep-sea turbidite surface as revealed by 210Pb excess, 230Th excess and planktonic distributions. Earth Planet. Sci. Lett., 90: 157- 173.

53 STUK-A145

THE PAST AND THE FUTURE OF THE 210Pb METHOD 1 2 F. El-Daoushy and R. Garcia-Tenorio department of Physics, Uppsala University, Box 530, S-751 21 Uppsala, Sweden 2Dept. Fisica Aplicada, Univ. Sevilla, Reina Mercedes 2, E-41012 Sevilla, Spain

NOTE: This letter to the editor compiles and summarizes the main ideas pre- sented and discussed in the talk given in the 2nd NKS/EKO-1 Seminar on Dating of Sediments and Determination of Sedimentation Rate, Helsinki 2-3 April 1997. Relevant publications by the authors and others are given in the list of references.

Previous research on the dating of sediments and determination of sedimen- tation rates, through the 210Pb method, did not allow full and proper exploita- tion of the sedimentation/accumulation dynamics in various aquatic systems. We will give some examples on existing limitations, difficulties and problems in the development of the 210Pb "method". Poor attention to selec- tion/utilization of experimental facilities and narrow definition of sam- pling/sample specifications has induced serious consequences concerning the quality of radiometric dating. Sampling strategies and sample pretreatment are conditioned to limited number of existing facilities and do not consider the utilization of alternative tools on the one hand, and possible disturbances through various processes in aquatic systems on the other. The so-called "standard 210Pb-dating models" are indiscriminately used although they are based on general assumptions that are tested on few soft-water lakes in tem- perate regions. These facts considerably limit the developments of new radio- metric approaches, including dating ones, that consider possible variations of the actual chemical, physical and sedimentological conditions influencing the concentrations and the fluxes of key radionuclides especially 2l0Pb. Such unfor- tunate trends ought to be rectified to help proper developments and applica- tions of the 210Pb method, particularly in combination with other nuclides, on a wider scale.

54 STUK-A145

General introduction

The chronology of environmental and climatic changes in atmospheric and aquatic systems is essential for the understanding of dynamic processes forc- ing various exchanges in land-water systems. The 210Pb method can be used for the dating of recent deposits, and mass-balance studies of aquatic systems, of the past 150-175 years. This is very helpful for the evaluation of environ- mental and climatic impacts produced by human activities.

Previous research on the 210Pb dating of sediments and related determination of the sedimentation rates did not allow full and proper exploitation of the sedimentation dynamics in various aquatic systems. The 210Pb dating method has generally been used as a routine tool to solve problems for uncritical users. Although the complexity of aquatic systems this method has been applied to single cores with the smallest number of samples. Poor attention to selec- tion/utilization of experimental facilities and insufficient definition of sample specification has induced serious consequences concerning the quality of the 210Pb dating. We believe that appropriate 210Pb routines ought to be based on research. Special efforts need to be devoted to improve the 210Pb dating reli- ability and related quality aspects. Good accuracies and agreement between alternative/complementary tools as well as extension of the upper dating limit have to be considered to enhance the quality and reliability of the 210Pb method. It is essential to use proper experimental facilities for accurate de- termination of 210Pb, in addition to other relevant nuclides, in different sedi- ment layers. 210Pb could be determined by direct gamma-ray spectrometry through its weak gamma-emission at 46.5 keV (4.25%) and via alpha-particle spectrometry of its grand-daughter 210Po. p-counting of the daughter 210Bi is also a possibility in some cases where time constrains make 210Po less attrac- tive. These alternatives have various specifications, advantages and limita- tions. Their combination might, therefore, be imperative in many cases.

Gamma-ray Spectrometry

No chemical separations are needed. This tool allows simultaneous determi- nation of 210Pb, 226Ra and other relevant U/Th radionuclides for the correct application of the 210Pb method. Other artificial radionuclides like 137Cs and 241Am, that could be determined by gamma spectrometry, are extremely useful to validate/evaluate the 210Pb ages. 7Be is also an interesting cosmogenic ra-

55 STUK-A145

dioisotope that is suitable for -particle interactions in aquatic systems. Gamma-ray spectrometry has a relatively high low limit of detection (LLD) in comparison with alpha-particle spectrometry. Well-type detectors are suitable for samples with high specific activities whereas low-energy, planar and coax- ial Ge detectors could be used for low specific activities where compensation by mass allows measuring bigger samples. However utilization of big samples may produce less temporal resolutions. Large sample sizes can be considered appropriate at high sedimentation rates such as in marine sediments or the like. The low energy of the 210Pb gamma-ray photons imply, in addition to geometrical sample considerations, self-absorption corrections. Self-absorption corrections are not an easy task as they ought to be based on experimental studies and for every sample/detector configuration.

Alpha-particle Spectrometry

The very low LLD of alpha-particle spectrometry permits the measurement of weak 210Pb activities and could be applied to small samples (0.3 - 0.5 grams or less). The application of alpha spectrometry is specially suitable at very low and moderate sedimentation rates. The application of alpha spectrometry is essential either in cases where good accuracies are required or for the older sediments where the total 210Pb is slightly above the levels of the 226Ra. Alpha spectrometry has special importance in speciation studies, that are important in rehabilitation strategies, where gamma spectrometry can not be used unless detailed chemistry is involved. Radiochemical procedures are necessary for the application of isotope dilution. Utilization of 209Po (4.832 MeV) as internal yield tracer instead of 208Po (5.114 MeV) has induced drastic simplification of the radiochemical procedure because its alpha-peak can be very well resolved from the 210Po (5.305 MeV). However, information on other radionuclides are needed through additional techniques/methods.

Gamma and alpha spectrometries are complementary techniques and ought not to be considered as competitive ones. Combination of alpha/gamma spec- trometries and their alternative application to different layers is, therefore, the best approach. P-counting could be of value in some cases. In gamma spec- trometry optimization of sample sizes to detector specifications is quite impor- tant for the best utilization of this tool. Chemical procedures used in alpha spectrometry ought to be conditioned to sample composition. Proper evaluation

56 STUK-A145

of the supported 210Pb through 226Ra, as indicator of the supported 210Pb should, therefore, be critically considered.

Sampling and sample pretreatment

Quality and accuracy of the radiometric data are very much dependent on sampling strategies, quality and specification of the sampling sites. Traditional sampling in aquatic systems are generally reduced to the collec- tion/measurement of single cores. Some lakes, for example, were/are heavily utilized for various independent monitoring/research programmes and high risks exist for disturbed sediments. Coordination of existing activities and proper records on sampling programmes in lakes are imperative as the credi- bility of lakes as indicators is extremely important for future studies especially those related to follow-up and rehabilitation issues. Spatial distribution of key radionuclides in lake bottoms is quite necessary to understand the behaviour and sedimentation dynamics, through mass-balance studies, of direct and indi- rect atmospheric inputs. Additional information that allow the selec- tion/development of the most suitable 210Pb data/models could be obtained through studies over the whole lake basin. Limited samples and amounts of material should not cause constrains on the applicability of alternative models. Lack of limnological data such as water content, organic content and grain-size in different layers may limit the possibilities to carry out various site-related, sample-related and model-related corrections/normalizations. These are ex- amples on problems that ought to be considered via well-organized coordina- tion between samplers, analytical teams, modelers and managers of land- water resources.

Alternative models

Another critical step in the 210Pb method is the application of appropriate models as defined by site and sediment specifications rather than forcing few models to all sites and conditions of sedimentation. Two models: the C.I.C. (Constant Initial Concentration) and C.R.S. (Constant Rate of Supply) are indiscriminately used for the construction of chronologies. These models are based on general assumptions that have been tested in few soft-water lakes situated in temperate regions. These models ought not to be considered as "standard 210Pb models" as they are not verified on global scale and additional corrections/normalizations steps are not involved in these models. Unfortu-

57 STUK-A145

nately the application of these models without careful consideration of the specificity of sites/regions has considerably limited the development of techni- cal and dating approaches on wider scales. New corrections/normalizations, that consider actual sedimentological, biological, chemical and physical condi- tions influencing the concentrations and fluxes of 210Pb, are expected in the future. It must be emphazised that there are no standard or universal 210Pb dating models that could be valid for all possible cases since by definition models can only be good or bad mathematical representations of reality. Un- less strong evidences exist to show the constancy of temporal and spatial fluxes of the 210Pb the credibility of the C.R.S., in its general and simple form, could be questioned. It is, therefore, recommended that independent alterna- tives ought to be used to validate the reliability and accuracies of the 210Pb dates.

In cases where the 210Pb flux is a site-related parameter, on spatial and tempo- ral scales, rather than a constant quantity it would be very informative to know various processes influencing it as this would help the development of new applications. Coastal marine sites, such as fjords, estuaries and semi-arid lakes with variable input levels from land could be critical sites for dating.

Alternative and complemetary dating techniques

Alternative and independent dating tools used to validate the 210Pb dating can be classified to radiometric, geophysical and isotopic approaches. Non of these tools, if applied separately, can serve as a general tool for climatic and envi- ronmental studies as they only provide partial solutions to specific issues. 137Cs records have been served to provide chronological information through peak values of main atmospheric inputs. 241Am, if measured with good accuracies, provides suitable alternative in cases where 137Cs suffers from geochemical mobility as is usually the case in peatlands and lake-sites with high water content and/or low sedimentation. Other problems that complicate the utiliza- tion of 137Cs is 137Cs inputs from point sources. Exact information on input functions from point sources are generally lacking as in marine sites suffering from mixed inputs from different sources such as those occurring in the North Sea and the Baltic Sea. Long residence time caused by poor sedimentation of 137Cs in water bodies, such as in the Baltic Sea, may cause further limitations 137 241 in the utilization of Cs. 239+24opu is g^o as gOO(i alternative as Am, how- + ever the measurement of 239 24opu is limited to alpha spectrometry which is a

58 STUK-A145

rather difficult task because of its low concentrations (239+24opu/i37Cs activity ratio in fallout is about 102). Unlike these radionuclides 90Sr is a very conser- vative environmental isotope and its utilization, that necessitates low-level beta counting, is only limited to ice-deposits. The sedimentological behaviour of ^Sr under different geochemical and limnological conditions is not properly studied yet.

In addition to the construction of high-resolution chronologies through radio- metric tools there are potential applications to explore various settling and sedimentation dynamics of particulate matter in water-bodies. "Within- system" and "between-system" interactions in lakes, or the like, could be de- tailed through combined radiometric measurements using sediment traps and sediment cores. Differentiation between these interactions has very interesting applications in rehabilitation studies especially if management policies are to be based on the specific conditions of the land-water resources.

References

Abril, J. M., Garcia-Leon, M., Garcia-Tenorio, R., Sanchez, C. I. and El- Daoushy, F, 1992. Dating of a marine sediment by an incomplete mixing model. J. Environ. Radioact., 15: 135-151.

Anderson, R., Schiff, S. L. and Hesslein, R. H., 1987. Determining sediment accumulation and mixing rates using Pb-210, Cs-137 and other tracers: prob- lems due to post-deposition mobility or coring artifacts. Can. J. Fish. Aquat. Sci., 44: 231-250.

Appleby, P. G. and Oldfield, F., 1978. The calculation of lead-210 dates assum- ing a constant rate of supply of unsupported lead-210 to the sediment. Catena, 5: 1-8.

Appleby, P. G., Richardson, N. and Nolan, P. J., 1991. Am-241 dating of lake sediments. Hydrobiol., 214: 35-42.

Axelsson, V. and El-Daoushy, F., 1989. Sedimentation in the Edsviken Bay studied by the x-ray radiographic and the Pb-210 methods. Geograph. Ann., 71A: 87-93.

59 STUK-A145

Bloesch, J. and Evans, R. D., 1982. Lead-210 dating of sediments compared with accumulation rates estimated by natural markers and measured with sediment traps. Hydrobiol., 92: 579-596.

Brunskill, G. L., Ludlam, S. D. and Peng, T. H., 1984. Fayetteville Green Lake, New York, U.S.A. VIII. Mass balance for Cs-137 in water, varved and non- varved sediments. Chemical Geol., 44: 101-117.

Chanton, J. P., Martens, G. S. and Kipput G. W., 1983. Lead-210 sediment in a changing coastal environment. Geochim. Cosmochim. Acta, 47: 1791-1804.

Crusius, J. and Anderson, R. F., 1992. Inconsistencies in accumulation rates of Black Sea sediments inferred from records of laminae and Pb-210. Paleocean- ogr., 7: 215-227.

Crusius, J and Anderson, R. F., 1995. Evaluating the mobility of Cs-137, Pu- 239+240 and Pb-210 from their distribution in laminated lake sediments. J. Paleolim., 13: 119-141.

Eddington, D. N., Val Klump, J., Robbins, J. A., Kusner, Y. S., Pampura, V. D. and Sandimirov, I. V., 1991. Sedimentation rates, residence times and radio- inventories in lake Baikal from Cs-137 and Pb-210 in sediment cores. Nature, 350: 601-604.

El-Daoushy, F., 1986. Scandinavian limnochronolgy of sediments and heavy metals. Hydrobiol., 143: 267-276.

El-Daoushy, F., 1988. A summary on the lead-210 cycle in nature and related applications in Scandinavia. Environ. Inter., 14: 305-318.

El-Daoushy, F., (in press). Radionuclide time-scales for palaeolimnological and limnological studies. Verh. Internat. Verein. Lamnol..

El-Daoushy, F., Olsson, K. and Garcia-Tenorio, R., 1991. Accuracies in Po-210 determination for lead-210 studies. Hydrobiol., 214: 43-52.

60 STUK-A145

El-Daoushy, F. and Garcia-Tenorio, R., 1995. Well Ge and semiplanar Ge (HP) detectors for low-level gamma spectrometry. Nucl. Instrum. Meth., A 356: 376- 384.

El-Daoushy, F. and Garcia-Tenorio, R., 1995. Photon (20-60 keV) self- absorption in small aquatic deposits. Nucl. Instrum. Meth., A 359: 622-624.

El-Daoushy, F. and Eriksson, M., (in press). Radiometric dating of recent lake sediments from a highly eroded area in semiarid Tanzania. J. Palaeolim.

Jaakkola, T., Tolonen, K., Huttunen, P. and Leskinen, S., 1983. The use of fallout Cs-137 and Pu-239+240 for dating of lake sediments. Hydrobiol., 103: 15-19.

Joshi, S. R., 1989. Common analytical errors in the radiodating of recent sedi- ments. Environ. Geol. Water Sci., 14: 203-207.

Manjon, G., El-Daoushy, F. and Garcia-Tenorio, R., 1997. 90Sr in lake sedi- ments. J. Radioanal. Nucl. Chem. Lett., 219: 95-98.

Robbins, J. A., 1978. Geochemical and geophysical applications of radioactive lead. In: J.O. Nriagu (ed.), The Biogeochemistry of Lead in the Environment. Elsevier Scientific, Amsterdam, 285-393.

Robbins, J. A. and Herche, L. R., 1993. Models and uncertainty in Pb-210 dat- ing of sediments. Int. Ver. Limnol. Verh., 25: 217-222.

Schuler, C, Wieland, E., Santschi, P. H., Stumm, M., Luck, A., Bollhalder, S., Beer, J., Bonani, G., Hoffmann, H.J., Suter, M. & Wolfli, W., 1991. A multi- tracer study of radionuclides in Lake Zurich, Switzerland. 1. Comparison of atmospheric and sedimentary fluxes of 7Be, 10Be, 210Pb, 210Po and 137Cs. J. Geo- phys. Res., 96: 17051-17065.

Wieland, E., Santschi, P. H. & Beer, J., 1991. A multitracer study of radionu- clides in Lake Zurich, Switzerland. 2. Residence times, removal processes, and sediment focusing. - J. Geophys. Res., 96: 17067-17080.

61 STUK-A145

ARTIFICIAL RADIONUCLIDES IN AN INTERTIDAL SEDIMENT FROM NORTHWEST ENGLAND

K Morris1, M J Keith-Roach2, A S Hursthouse3, J C Butterworth2, F R Livens2, L K Fifield4, J P Day2 & R D Bardgett5 department of Chemistry, Florida State University, Tallahassee FL 32306, USA department of Chemistry, University of Manchester, Manchester Ml3 9PL, UK 3Department of Chemistry, University of Paisley, High Street, Paisley PA1 2BE, UK department of Nuclear Physics, Australian National University, Canberra ACT 2610, Australia 5School of Biological Sciences, University of Manchester, Man- chester M13 9PL, UK

Abstract

An intertidal sediment core has been analysed for the principal transuranium elements present in the BNFL Sellafield radioactive waste discharges (Np, Pu, Am) and the high yield fission products "Tc and 137Cs. In general, the profile distributions of Cs, Pu and Am closely resemble the historical pattern of dis- charges whereas the distribution of Np, for which the actual discharge history is unknown, cannot be matched easily either with the likely pattern of dis- charges, or with a recent reconstruction of them. The Np data can be inter- preted as suggesting remobilisation and upwards transport of this element within the sediment profile. The apparent mobility of Np is plausible, given the stability of the NpO2+ ion in the environment and its known, low solid- solution partition coefficient. The Tc distribution apparently matches that of 137Cs, which suggests that there has been no significant Tc redistribution. Al- though only Np has been remobilised on a large scale, detailed analysis of the isotopic and elemental ratios found in deeper sections of the profile suggest that most, if not all elements are also moved downwards to a limited extent,

62 STUK-A145

being affected in the order Tc > Cs > Np > Pu > Am. The radionuclide distri- butions are difficult to explain in detail, although measurements (Carr & Blackley, 1986a) show that the hydrology of the salt marsh is very complex, with both upwards and downwards water movement occurring.

Interstitial water samples have been collected using porous cup samplers and early results from these analyses show that there is a pronounced seasonally in the pattern of dissolved Pu, which apparently relates to changes in dis- solved Fe and Mn. More recent work has concentrated on the characterisation of changes in the sediment microbial community and on the development of analytical methods for the analysis of dissolved Np, apparently the most read- ily mobilised of the transuranic elements, which is present at concentrations of the order of 108 litre1.

Introduction

Since 1952, low-level, liquid radioactive wastes have been discharged from the BNFL Sellafield reprocessing plant into the north-eastern Irish Sea. These effluents have contained a wide variety of fission and activation products, of which the ones discussed in this paper are "Tc, 137Cs, 237Np, 238Pu, 239Pu, 240Pu and 241Am. The effluents have changed substantially over time, both in isotopic and elemental composition and in activity concentration. The discharge histo- 137 ries for some radioisotopes ( Cs, Pu-a (= 238pu+239pu+24opu); 24iAm) are well known (see e.g. Kershaw et al. 1990; 1995), whereas for others ("Tc, 237Np), information is much less comprehensive.

Many of the elements present in the discharges have complex environmental chemistries (see e.g. Allard et al., 1984; Choppin & Stout, 1989; Dozol & Hagemann, 1993). Note that in the following brief summary, Ra is used to de- note a solid-solution partition coefficient, but does not imply that the system is at equilibrium. Tc is likely to be present as TcCV except in strongly reducing conditions, when it can be reduced to cationic Tc(IV) species. The TcCV ion reacts only weakly with particles and surfaces (Rd < 10 ml g1) and therefore behaves conservatively in the water column. In contrast, the Tc(IV) species are much more reactive, particularly with organic matter, and are readily removed from solution.

63 STUK-A145

The transuranium have different but related chemistries. Np has access to oxidation states IV and V, Pu to III, IV, V and VI and Am to III at environmental Eh/pH values (Lieser & Muhlenweg 1988; Morse & Choppin, 1991). The low valency species form simple cations which are extensively hydrolysed and conse- quently behave non-conservatively (Rd « 105 - 106 ml g"1 for M3+ and M4+) whereas the higher valency species form the much more soluble dioxygenyl ions, MO2"1" and MO22"1" (Rd * 104 ml g-1).

Many of the radionuclides discharged into the Irish Sea from Sellafield are thus highly particle-reactive and become readily associated with suspended particu- late matter (Kershaw et al., 1990; 1995). Even those with relatively low Rd val- ues are significantly associated with particulates at the relatively high particle concentrations typical of inshore Irish Sea waters (McKay & Walker, 1990; McKay et al. 1993). The sediment-associated radionuclides are removed from the water column and accumulated in areas of fine-grained sediment on the sea bed. To a first approximation, the redistribution of these fine-grained sediments then controls radionuclide transport. It is well known that sediment from this source is transported by water movements into estuaries along the Irish Sea coast where it can be deposited and/or trapped by vegetation (Horrill, 1984). This has led to substantial radionuclide accumulation in a number of estuaries, most no- tably in those of the Rivers Esk, Mite and Irt, which he about 15 km south of the Sellafield site (Kelly & Emptage 1992).

Offshore, data from numerous sediment cores shows that the historical pattern of discharges from Sellafield is well preserved, although extensive bioturbation can occur (Kershaw et al., 1984) and it has been possible to match sediment profile activity distributions with known discharge patterns (MacKenzie et al., 1994). Similar patterns can be observed in harbour and estuarine sediments (e.g. Aston & Stanners, 1981; Kershaw et al. 1995) and tide-washed soils (e.g. Hurst- house et al., 1991) and have been used to reconstruct previously unknown dis- 2 240 charge histories [e.g. for 38Pu, 239pu and Pu individually (Kershaw et al., 1990; Kershaw et al., 1995) and for 237Np (Kuwabara et al., 1996)]. Comparison of the distributions of different radionuclides in the same sediment profiles and matching with known or reconstructed discharge histories can allow identifica- tion of post-depositional remobilisation, if it occurs. The purpose of this paper is primarily to interpret data from an intertidal sediment profile in terms of the known or reconstructed historical discharges and hence to identify possible post- depositional remobilisation. Since any remobilisation must involve the intersti-

64 STUK-A145

tial solution phase, a secondary objective is to attempt to define geochemical controls on element solubility.

Methods a) Sediment Cores. The site used in this study is a vegetated salt marsh in the valley of the River Esk, estimated frequency of inundation 20-30 times per year. An intact block of sediment was collected in late 1992 and sectioned at approxi- mately 1 cm intervals.

The samples were sieved through a 2mm mesh and air-dried. Samples for acti- nide analysis were ashed at 550°C and extracted by boiling in aqua regia. b) Interstitial Waters. Porous cup samplers were installed and allowed to bed in for 6 months. During this time, they were pumped out monthly and the solutions discarded. Subsequently, the samples were retained for analysis. These were initially evaporated to dryness, wet ashed with nitric acid and redissolved in 9 M HC1 for radiochemical separations. c) Radiochemical Separations. In brief, Pu was separated from the sediment extract solutions by anion exchange and electrodeposited on to a stainless steel planchet for cc-spectrometry using a 242Pu yield monitor (Livens & Quarmby, 1988). Np was also separated by anion exchange but determined by ICP-MS using a 239Np yield monitor which was separately analysed by y-spectrometry (Morris & Livens, 1996). "Tc was determined on separate aliquots of dried sediment, which were wet ashed with H2O2/HCI, then purified by coprecipita- tion, anion exchange and finally extraction chromatography. The "Tc recover- ies were monitored with stable Re and concentrations were determined by ICP-MS (Butterworth et al., 1995). Activity concentrations of 137Cs and 24iAm were determined by high resolution y-spectrometry using dried sediment samples (Morris & Livens, 1996).

The chemical separations used for the analysis of transuranium elements in the interstitial water samples were based on those used for the sediments, modified where necessary to cope with the high salt content and much lower activity concentrations. Accelerator (Fifield et al., 1997) rather than ICP-MS, was used to determine 237Np concentrations. The activity

65 STUK-A145

concentrations of 241Am were determined by anion exchange, extraction chro- matography and a-spectrometry with a 243Am yield monitor.

Results and Discussion

The results for all elements in the sediment profile are presented in Table I, expressed as Bq nr2, and illustrated in Fig. 1. Data tor the interstitial waters are presented in Table II and illustrated in Fig. 2. The discussion of the results is conveniently divided into three sections. a) Cs, Pu and Am. The distributions of these elements down the salt marsh profile qualitatively resemble the discharge histories of these radionuclides from Sellafield (Kershaw et al., 1990; Mackenzie et al., 1994). It is possible to match profile distributions and discharge histories to give an estimate of the sediment accumulation rate. In the case of Cs and Am, there is a single peak, 2 whilst there are two for 39,24opu The accumulation rates estimated from each of these peaks are consistent with each other at around 6 mm yr1 (Table III) and agree well with those measured directly by Carr & Blackley (1986a,b) and Emptage & Kelly (1992). The peak in the 137Cs distribution is broader and less sharply defined than that in the discharge history, which could result from extensive mixing of contaminated sediment in the offshore sediment reservoir, or a significant input of 137Cs in solution. There is certainly evidence for an occasional flushing of the sediment with sea water and complex hydrological behaviour in the marsh (Carr & Blackley, 1986b), which could introduce addi- tional 137Cs.

However, the ratio of the 137Cs to 238Pu + 239.24°Pu inventories found at this site is 2.38. Since the total discharges of 137Cs and Pu-a up to the date of sampling were 31550 and 671 TBq respectively and the measured Ra values in near- shore Irish Sea waters are of the order of 104 and 105 ml kg-1 respectively (McKay & Walker, 1990; McKay et al., 1993), the ratio which might be ex- pected if sediment is the dominant input is 4.7. Although the observed and expected inventory ratios are of the same order of magnitude, the fact that the expected ratio is a factor of two higher than the observed ratio could be ex- plained by one or more of: a lower 137Cs Rd, a higher Pu Ra or an additional Pu input equal to that associated with sediment particles. This last possibility seems very unlikely and since published Ra values show such variability, this seems the likely explanation.

66 STUK-A145

30

• Cs-137

O Np-237

A Pu-239,240

V Am-241

O Tc-99

o o reCO

15 20

Figure 1. Radionuclide distributions in a sediment profile. Each section is approximately 1.2 cm thick.

67 STUK-A145

Table I. Radionuclide activities in the saltmarsh sediment profile (Bq m-2).

Depth 99Tc 13'CS 237Np 238pu 2239.240pu Interval (kBq nr2) (Bq nv2)1 (kBq m-2) (Bq m-2) (kBq nr2) (kBq nr2) (cm)

0.0-1.2 16.5 ±0.1 107 ±4 2.2 ±0.2 10.4 ±0.5 19.1 ±0.4

1.2-2.4 460 ±30 38.7 ±0.3 153 ±5 3.9 ±0.6 16.7 ±1.0 38.7 ±0.8

2.4-3.6 65.8 ± 0.4 95.9 ± 0.9 4.8 ±0.5 23.6±1.1 46.7 ± 0.9

3.6-4.8 750 ± 50 195 ±1 102 ±3 11.0 ±1.8 53.7 ±3.3 90.3 ±1.8

4.8-6.0 310 ±1 113±4 20.0 ±3.5 99.3 ±7.0 162 ±3

6.0-7.2 1500 ±110 360 ±2 118 ±4 39.7 ±5.1 169 ±9 210 ±4

7.2-8.4 370 ±2 64.1 ±2.0 31.1 ±2.4 112 ±6 251 ±5

8.4-9.6 1300 ± 90 335 ±2 59.1 ±1.9 28.5 ±3.9 153 ±8 380 ±7

9.6-10.8 196 ±1 41.4 ±1.3 14.6 ±3.2 209 ± 10 151 ±3

10.8-12.0 250 ±30 87.0 ± 0.4 17.7 ±0.6 1.3 ±0.2 9.5 ±0.5 32.4 ± 0.7

12.0-13.2 91.0 ±0.5 18.1 ±0.6 0.90 ±0.12 14.1 ±0.9 15.2 ±0.4

13.2-14.4 290 ± 30 75.3 ±0.4 13.2 ±0.5 0.58 ±0.13 12.2 ±0.6 7.5 ±0.2

14.4-15.6 70.3 ± 0.4 11.2 ±0.4 0.56 ± 0.06 7.3 ±0.3 6.4 ±0.2

15.6-16.8 330 ± 30 70.3 ± 0.4 13.2 ±0.5 0.72 ±0.13 7.4 ±0.5 5.0 ±0.2

16.8-18.0 47.1 ±0.2 7.4 ± 0.2 0.39 ± 0.06 4.2 ± 0.2 2.7 ±0.2

18.0-19.2 290 ±30 72.9 ± 0.4 11.0 ±0.3 0.82 ±0.07 5.7 ±0.3 7.0 ±0.3

19.2-20.4 55.2 ±0.3 9.2 ±0.3 0.34 ±0.07 2.9 ±0.1 3.4 ±0.3

Total 5200 ± 160 2460 ±4 955 ± 10 120 ±9 910 ±18 1420 ± 10 1 Note that only alternate sections of the core were analysed for "Tc and the inventories for this isotope were determined by interpolation.

68 STUK-A145

Table II. Concentrations of plutonium (mBq I1) in interstitial waters Jan-Dec 1994.

Date 239.240pu (mBq 1-1)

11/1/94 2.3 ± 0.2 10/2/94 2.2 ±0.2 18/3/94 2.2 ± 0.2 3/5/94 2.3 ±0.2 3/6/94 3.5 ±0.3 12/7/94 1.9 ±0.2 11/8/94 1.2 ±0.1 6/9/94 1.1 ±0.1 11/10/94 1.9±0.1 15/11/94 Sample lost 20/12/94 3.0 ±0.2 15/1/95 2.1 ±0.2

Table III. Sedimentation rates at the saltmarsh calculated from peaks in 137Cs, 23S.24opu and 241Am profile distributions in the saltmarsh and compared with those reported elsewhere. Sedimentation Rate (mm yr1)

239,240pu 5.0 1978 peak

239,240pu 5.6 1973 peak

241Am 5.2

137CS 4.8 Direct measurement <10 Carr & Blackley, 1986a 137Cs 4 Kelly & Emptage, 1992

69 STUK-A145

Figure 2. Dissolved Pu concentrations in salt marsh interstitial waters.

Although there is a good qualitative similarity between discharge histories and sediment profiles, the proportion of the 137Cs inventory in the deeper sections of the core (Sections 13-17, corresponding approximately to deposition between 1968 and 1959 or, allowing a 2 year lag (Mackenzie et al., 1994), to discharges between 1966 and 1955) is 12.9% compared with 1.6% of the total discharges which occurred in that period. The presence of such a high proportion of the sediment inventory at depth suggests that there has been a limited redistri- bution downwards of 137Cs. In contrast, there appears to be a deficiency of

239,24opu at depth, with 2.9% of the inventory actually present, as against 5.3% expected. The 241Am data are inconclusive since they are greatly complicated by ingrowth from 241Pu decay. As yet, there is no convincing explanation for these observations.

70 STUK-A145

b) Np and Tc. It is difficult to interpret the data for these two elements in as much detail, since they were not measured in the discharges for most of the period of interest (Tc and Np data exist only from 1978). Nevertheless, there is no reason why the concentration patterns of these elements in the discharges should be substantially different from those of Cs, Pu, and Am, increasing from the start of discharges to maxima in the 1970s and falling after about 1980. Kuwabara et al. (1996) have recently reconstructed the 237Np discharge history from an intertidal sediment core and concluded that the total released was 6-8 TBq, with the broad pattern of Np discharges resembling those of the other elements. Little appears to be known of the total "Tc discharge and its marine chemistry is so different to that of the other elements that estimated releases are unrealistic.

137 The "Tc profile generally resembles those of Cs, 239,24opu and 24iAm( all of which can be interpreted as showing relatively little redistribution, so it seems reasonable to interpret the "Tc one in the same way. However, a higher than expected proportion (11.9%) of the "Tc inventory is in the deeper sections of the profile and, if the excess of 137Cs and 237Np (see below) at depth is evidence of limited downward migration, then the "Tc is affected to an even greater degree.

The 237Np profile distribution is quite unlike that of any other radionuclide. 237Np is the transuranium element with the greatest proportion of the inven- tory at depth (5.6% compared with the 0.15-0.2% respectively of the total dis- charges [the expected value is based on data in Kuwabara et al. (1996)]) but its distribution in the upper sections of the profile is very different from that of any of the other elements. In marked contrast to the data of Kuwabara et al. (1996), which suggest that there should be a substantial increase in 237Np re- leases in about 1978, there are no pronounced peaks in activity comparable to those observed for 239,24opu or 24iAm The preservation of Cs, Pu and Am dis- charge histories in the sediment profile provides evidence against substantial redistribution of sedimentary material, so we interpret the Np pattern as rep- resenting geochemical remobilisation and upward transort of this element. Upwards migration in association with low salinity water is possible, given the complex hydrology of the marsh (Carr & Blackley, 1986b) although, as noted above, there is also evidence for downwards transport of several radionuclides. c) Interstitial Waters. Although Np is clearly the element of greatest interest in the interstitial waters since it is apparently subject to the most extensive

71 STUK-A145

redistribution, the accelerator mass spectrometry technique for measurement of Np in these samples has only recently been developed (Fifield et al., 1997). Only Pu data are available for the earliest interstitial water samples and Np and Am data are only now being obtained using different samples to those in which Pu was measured.

The Pu data for the samples collected during 1994 are illustrated in Fig. 2 and show a pronounced seasonal pattern. Trace element (Fe and Mn) data (not shown) for these samples suggests that there is a complex relationship be- tween dissolved Pu and dissolved Fe and Mn. Dissolved Pu concentrations appear to rise with Mn concentrations, then fall as Fe increases. Later in the year, dissolved Fe falls and Pu rises again, before finally falling as dissolved Mn declines at the end of the year. Such a close relationship with Fe and Mn suggests strongly that Pu is incorporated in the microbially-driven redox cy- cling of these elements. Further work is in progress to characterise the micro- bial communities in more detail by analysis of microbial phospholipid fatty acids. The limited amount of data available for other transuranium elements suggest that there are significant temporal variations in these as well al- though there are, as yet, too few data to identify any patterns.

Conclusions

The distributions of "Tc, 137Cs, Pu and Am in the profile of an intertidal sedi- ment from west Cumbria closely resemble the known or expected temporal variations in the Sellafield discharges. In contrast, the distribution of Np is difficult to interpret in the same way, which suggests that this element has been remobilised within the profile. In addition, the profile distributions of all the elements suggest that they have been transported downwards to a limited extent.

Studies of interstitial waters have concentrated so far on Pu and show that there is a seasonality in its solubility, which is apparently related to changes in Fe and Mn solubility. More recently, dissolved Np concentrations have been measured for the first time and found to be of the order of 108 atoms I1. Fur- ther work is in progress to obtain more transuranic element data and to char- acterise the microbial community quantitatively.

72 STUK-A145

References

Allard, B., Olofsson, U. & Torstenfelt, B., 1984. Environmental actinide chemistry. Inorg. Chim. Acta, 94: 205-221.

Aston, S. R. & Stanners, D. A., 1981. Plutonium transport to and deposition and immobility in Irish sea intertidal sediments. Nature, 289: 581-582.

Butterworth, J. C, Makinson, P. R. & Livens, F. R., 1995. Development of a technique for the determination of "Tc in soils and sediments. Sci. Total Environ., 173/174: 293-300.

Carr, A. P. & Blackley, M. W. L., 1986a. Implications of sedimentological and hydrological processes on the distribution of radionuclides: the example of a saltmarsh near Ravenglass, Cumbria. Est. Coastal Shelf Sci., 22: 529-543.

Carr, A. P. & Blackley, M. W. L., 1986b. The effects and implications of tides and rainfall on the circulation of water within saltmarsh sediments. Limnol. Oceanogr., 31: 266-276.

Choppin, G. R. & Stout, B. E.( 1989. Actinide behaviour in natural waters. Sci. Tot. Environ., 83: 203-216.

Dozol, M. & Hagemann, R., 1993. Radioniiclide migration in groundwaters: review of the behaviour of actinides. Pure and Applied Chemistry, 65: 1081-1102.

Fifield, L. K, Cresswell, R. G., Day, J. P., Clacher, A. P., King, S. J., Morris, K, Livens, F. R. & Priest, N. D., 1997. AMS of the planetary elements. Nucl. Inst. Meth. Phys. Res. B, 123, 400-404.

Horrill, A. D., 1984. Concentrations and spatial distribution of radioactivity in an ungrazed saltmarsh. In: P. J. Coughtrey (ed.), Ecological aspects of radionuclide release. British Ecological Society Special Publication, 3: 199-215. Blackwell, Oxford.

Hursthouse, A. S., Baxter, M. S., Livens, F. R. & Duncan, H. J., 1991. Transfer of Sellafield-derived Np-237 to and within the terrestrial environment. J. Environ. Radioactivity, 14: 147-174.

73 STUK-A145

Kelly, M. & Emptage, M., 1992. Radiological assessment of the Esk estuary. UK Department of the Environment Report DOE/HMIP/RR/92/015. DoE, London.

Kershaw, P. J., Swift, D. J., Pentreath, R. J. & Lovett, M. B., 1984. The incorporation of plutonium, americium and curium into the Irish Sea bed by biological activity. Sci. Total Environ., 40: 61-81.

Kershaw, P. J., Woodhead, D. S., Malcolm, S. J., Allington, D. J. & Lovett, M. B., 1990. A sediment history of Sellafield discharges. Journal of Environmental Radioactivity, 12: 201-242.

Kershaw, P. J., Woodhead, D. S., Lovett, M. B. & Leonard, K. S., 1995. Plutonium from European reprocessing operations. J. Appl. Radiation , 46: 1121-1134.

Kuwabara, J., Yamamoto, M., Assinder, D. J., Komura, K. & Ueno, K., 1996. Sediment profile of Np-237 in the Irish Sea- estimation of the total amount of Np-237 discharged from Sellafield. Radiochim. Acta, 73: 73-81.

Lieser, K. H. & Muhlenweg, U., 1988. Neptunium in the hydrosphere and in the geosphere. I. Chemistry of neptunium in the hydrosphere and sorption of neptunium from groundwaters on sediments under aerobic and anaerobic conditions. Radiochimica Acta, 43: 27-35. livens, F. R. & Quarmby, C, 1988. Sources of variation in environmental chemical analysis. Environment Intl., 14: 271-275.

McKay, W. A. & Walker, M. I., 1990. Plutonium and americium behaviour in Cumbrian near-shore waters. J. Environ. Radioactivity, 12: 267-284.

McKay, W. A. & Pattenden, N. J., 1993. The behaviour of plutonium and americium in the shoreline waters of the Irish Sea- a review of Harwell studies in the 1980s. J. Environ. Radioactivity, 18: 99-132.

Mackenzie, A. B., Scott, R. D., Allan, R. L., Ben Shaban, Y. A., Cook, G. T. & Pulford, I. D., 1994. Sediment radionuclide profiles- implications for mechanisms of Sellafield waste dispersal in the Irish Sea. J. Environ. Radioactivity, 23: 39-69.

74 STUK-A145

Morris, K. & Livens, F. R., 1996. Interactions of actinides with sediments in intertidal areas in west Cumbria, UK. Radiochim. Acta, 74: 195-198.

Morse J. W. & Choppin G. R., 1991. The chemistry of transuranic elements in natural waters. Reviews in Aquatic Sciences, 4: 1-22.

75 STUK-A145

THE MOBILITY OF RADIOCAESIUM IN LAKE SEDIMENT AND IMPLICATIONS FOR DATING STUDIES

J.T. Smith1, D.G. Ireland2, R.N.J. Comans3, L. Nolan1.

institute of Freshwater Ecology, East Stoke, Wareham, BH20 6BB, United Kingdom 2Dept. of Physics, University of Glasgow, University Avenue, Glasgow, G12 8QQ, United Kingdom Netherland^Netherlands Energy Research FouFoundatior n (ECN), P.O. Box 1, 1755 ZG Petten, The Netherlands

Abstract

We have carried out experiments to study the uptake of radiocaesium by sediment cores from Esthwaite Water, Cumbria, UK in order to develop a model to simultaneously describe uptake of radiocaesium from the water column and diffusion of activity within the sediment. By measuring the rate of diffusion of activity through the sediment and sediment pore waters as a function of time we have assessed the validity of the equilibrium approach to modelling radiocaesium sorption and mobility in sediments. By comparison with in situ measurements it was found that rates of diffusion of activity introduced by diffusion across the sediment-water interface (as in our experiments) were approximately one order of magnitude greater than those determined from Chernobyl and weapons test activity-depth profiles. Modelling of the removal of activity from Esthwaite Water showed that the majority of the activity was transported to the sediments by attachment to and settling of suspended particles. It is concluded that activity deposited on particulates (forming the majority of activity observed in situ) was more strongly bound to sediments than that introduced by direct diffusion. This explanation is consistent with observed "two-phase" 137Cs profiles in sediments. Predictions of future of the weapons test peak are made on the basis of our studies. It is shown that the peak in the activity-depth profile will remain an important marker for dating studies for the forseeable future.

76 STUK-A145

Introduction

Previous studies into the sorption of radiocaesium to sediments have shown that the solids-aqueous distribution coefficient (Ka) is inversely proportional to the content of competing ions (specifically NH4+) in the sediment pore waters. It is also known that over time Cs progressively "fixes" to illitic clay in soils and sediments, the rate of uptake having a half time of around 100 d (Comans & Hockley 1992). The continuing mobility of radiocaesium in sediments over long time periods (years after fallout, Evans et al. 1983) further suggests a reverse reaction from "fixed" sites to chemically available phases. Using measurements of bomb- and Chernobyl-derived radiocaesium in the sediments of Hollands Diep and Ketelmeer in The Netherlands, a model for transport in lake sediments was developed which incorporated competitive and sorption kinetic effects (Smith & Comans 1996). This model showed that the mobility of radiocaesium in these sediments was broadly consistent with independent studies of its sorption properties. It was further shown that Cs may desorb from the "fixed" phase with a half-time of approximately 10 years.

Detailed measurements of 137Cs activity-depth profiles in laminated sediments (Crusius & Anderson 1989; Crooks 1991; Meili & Worman 1996), have, however, shown mobility which cannot be explained by current models. Such profiles are characterised by a sharp peak in activity corresponding to the 1986 Chernobyl fallout, with a "tail" of activity penetrating several cm below the peak. These profiles imply that 137Cs may be transported in two fractions: an "immobile" fraction corresponding the the peak and a more mobile fraction forming the tail. This behaviour cannot, to our knowledge, be explained by current models of Cs sorption chemistry, including that outlined in Smith & Comans (1996). In addition, modelling studies cannot reliably "deconstruct" such a profile to determine sorption processes since this would require modelling the evolution of either a series of profiles over time, or a knowledge of the distribution of 137Cs between sorption phases at the time of deposition.

We have carried out experiments to study the uptake of radiocaesium by sediment cores from Esthwaite Water, Cumbria, UK in order to develop a model to simultaneously describe uptake of radiocaesium from the water column and diffusion of activity within the sediment (Smith et al. 1998). By measuring the rate of diffusion of activity through the sediment and sediment pore waters as a function of time we have assessed the validity of different approaches to the

77 STUK-A145

modelling of radiocaesium sorption and mobility in sediments. Comparisons will be made of processes and models determined from these experiments with in situ measurements of activity-depth profiles of radiocaesium resulting from weapons testing and Chernobyl. The implications of diffusional transport of radiocaesium for dating studies will be discussed.

Methods

Experimental measurements.

During December 1994, 14 sediment cores were taken from Esthwaite Water at a water depth of 14 m using a Jenkin Corer. Each core was extruded into a plastic core tube which was constructed from a series of separable concentric rings (Fig. 1), the rings being sealed with waterproof tape covered by two coatings of wax.

Plastic Core Tube.

. Water Level

Concentric Kings

Wax Seal

Sediment

Figure J. Experiment design.

78 STUK-A145

The overlying water was reduced to a depth of 2 cm, then spiked with 10 kBq 134Cs. The cores were sealed and stored in a cold room at 4°C. After different incubation periods (15 min., 3 h, 24 h, 3 d, 10 d, 61 d, 1 y), two cores were sectioned by cutting through the tape and wax seal and removing each ring in turn (corresponding to sediment depths 0 - 0.5cm, 0.5 - 1.0 cm, 1.0 - 1.5 cm, 1.5 - 2.0 cm, 2 - 3 cm, 3 - 4 cm, 4 - 5 cm, 5 - 6 cm, 6 - 8 cm, 8 - 10 cm). At each time, the sediment from one core (labelled (a)) was centrifuged at 10,000 rpm for 10 min to remove the pore water. The pore water was filtered through a 0.45 urn cellulose acetate filter and counted on a Ge-Li gamma detector, a sub-sample being retained for determination of NH4+ concentration. The sediment was resuspended in 1 M ammonium acetate solution (solidrsolution ratio approximately 1:10) and agitated for 24 h before being re-centrifuged to remove the extractant. Three sequential extractions were carried out, the extractant being bulked and filtered through 0.45 fim cellulose acetate filters, then counted for 134Cs. The solid was dried, weighed, then counted for 134Cs. At each time period, the second core (core (b)) was sectioned in the above manner, the sediment from each slice being simply dried and counted for 134Cs. The overlying water from each core was also counted.

In-situ measurements.

During August 1991, three sediment cores were taken from Esthwaite Water at a water depth of 14m using a Mackereth type corer. The core samples were analysed according to the methods outlined in Comans et al. (1989) and Smith and Comans (1996). The cores were sliced in a glove box under nitrogen at- mosphere in order to preserve their anoxic status. The porewater from each sample was removed by centrifugation then filtration through a 0.2 um cellu- lose acetate filter. The porewater samples from each core were combined, each having been checked for their major ion composition to ensure that there were no lateral variations between the cores. The exchangeable 137Cs on the solid phase was determined by three 24h sequential extractions with 0.1 M NH4AC at a solid:liquid ratio of approximately 1:10. The solid and (after pre- concentration using ammonium molybdophosphate, AMP) aqueous and ex- change solutions were counted for 137Cs on a Ge-Li gamma-detector.

79 STUK-A145

Modeling

If it is assumed that the activity is input to the sediment as a sharp "spike" at the sediment surface, and that this spike diffuses through the sediment with constant diffusion coefficient, D (cm2.?1) then the transport of the activity with concentration C (Bq.g-1) can be described by the advection-diffusion equation:

(1)

dt ex2 ~ dx where o is the sedimentation rate (cm/y) and X is the radioactive decay constant. In our experiments co is zero and activity concentrations were decay-corrected to the date of the beginning of the experiment. Assuming that all of the sorbed activity is in exchangeable form, and that this exchange process is in equilibrium with respect to the diffusion timescale, then D is approximated by: (2)

£>*-

5 (e.g. Berner 1980) where Do is the molecular diffusion coefficient (1.45 x 10 cm2/s at 10 °C, Li & Gregory 1974) of Cs in water, § is the porosity, s is the drymass per unit wet volume and y/ is a tortuosity factor, approximated by y/ = 2 e (Lerman, 1979). K d is the equilibrium solids-aqueous distribution coeffi- cient (cm.g-1), the superscript "e" distinguishing this "exchangeable" distribu- tion coefficient from that calculated from the total sorbed activity (Kd) which may include non-exchangeable fractions. We will use the exchangeable distri- e bution coefficient, K d, to denote model determined values in which equilib- rium sorption is assumed. The "total" Ka denotes measured solid sorbed activ- ity divided by measured pore water activity. This latter quantity is not a true equilibrium distribution coefficient since slow sorption reactions mean that true chemical equilibrium may not have been reached. For an impermeable upper boundary (i.e. diffusion can only take place downwards from the sedi- ment surface), the solution to equation (1) for a decay-corrected tracer is ap- proximated by:

80 STUK-A145

(3) 2' " M \ (x-at/\ , f (x + o>t) , , . 4Dt \ r\ 4Dt 2(nDt)J x

(Kudelsky et al. 1996) where M is the mass of diffusing substance, t is the time since input to the sediment, and x is sediment depth.

Diffusion across the sediment/water interface & transport in the sediment.

For the case in which activity is transferred to the sediment across the sediment- water interface (as in our incubation experiments) we have used a numerical model based on that detailed in Smith & Comans (1996). The equations are somewhat simplified since, in our experiments, the sedimentation rate, co (cm.y1) is zero and there is no physical mixing of the sediment.

According to this model, activity sorbed to exchangeable sites on the solid is in e equilibrium with that in solution with distribution coefficient, K d . The trans- port of activity, Ce> in the aqueous/exchangeable phases is described by: (4)

Ot fa '-.^.u,,^.,, -kfC. + kbsCt-' where kf, kb (d1) are, respectively, rate constants for transport to and from "fixed" sites. The transport of activity in the "fixed" phase, G, is given by: (5)

s^+ = - kbsCi + kfCe - XsCi and the boundary conditions for transfers of activity across a boundary layer of thickness z (cm) are: (6)

- Ca) )

81 STUK-A145

(7) F,(O,t) = 0

where Ca is the changing concentration of activity in the overlying water:

(8)

c Ce(O,t))) 7C dt bz[a $+ sKed} a where 8 is the depth of overlying water.

Results

Experimental measurements.

The initial removal of activity from the overlying water (Fig. 2) was very rapid, around 90% of the activity being removed after 3 days, although removal at a slower rate continued over the whole experimental period. The removal of activity from the overlying water was accompanied by a progressive diffusional transport through the sediment and sediment pore waters (Fig. 3.). The similarity of the activity-depth profiles in the solid phase and pore waters show that the two phases are strongly coupled by sorption processes which are short with respect to the diffusion timescale.

None of the key chemical parameters (Kd, exchangeability and [NH4+]) showed any significant changes with depth at any of the incubation periods. Over the period of the experiment, measured Ka and exchangeability values remained remarkably constant at values of ca. 2060 ±740 l.kg-1 and 65 ±8.4% respectively. [NH4+] showed a tendency to increase over the period of the experiment, although with the exception of the last time period (after 1 year) it remained within the range observed from in situ measurements in Esthwaite Water (J. Hilton, unpubl. res.). The increase in ammonium concentration over time appeared to have no effect on 134Cs exchangeability or Kd. Potassium concentrations in the sediments were approximately a factor of 10 lower than [NH4+] and, also given the significantly higher selectivity of [NH4+] over [K+] in sediments (Wauters 1994), could therefore be considered negligible in its effect on radiocaesium Ka in these sediments.

82 STUK-A145

• Measured z = 0.1 cm z = 0.5cm z=1.0cm Kd = 500l/kg I- Kd= 10,000 I/kg o I <

10 Time (days)

Figure 2. Change in overlying water activity vs. time. Error bars show range in the two replicate samples (Smith et al. 1998). It can be seen that after a pe- riod of 10 days the thickness of the boundary layer, z, no longer has an influ- ence on removal: in the long term, removal is determined by the sediment Kd.

In situ measurements

Total 137Cs activities were partitioned between activity resulting from nuclear weapons testing (largely deposited in 1958-63) and the Chernobyl accident (1986) by measurement of 134Cs and using a 134Cs/137Cs ratio of 0.6 for activity resulting from the Chernobyl accident (Cambray et al. 1987). Due to its relatively short half life (ca. 2.2 years), no 134Cs activity deposited during nuclear weapons testing is now present in the environment. The activity-depth profiles resulting from the weapons test and Chernobyl fallout events are shown in Fig. 4.

83 STUK-A145

10 day Incubation 2 month Incubation 1 Year Incubation

300 -

250 -nl i - 0 200 - 1 •»- 1 «»- I so!

Depth (cm) Depth (cm) Depth (cm)

10 day Incubation 2 month Incubation 1 Year Incubation

018 -

016 - '

- »«i O 55 012 - 0. '

Depth (cm) Depth (cm) Depth (cm)

Figure 3. (a) Change in solid phase activity over time; (b) change in pore water activity over time (Smith et al. 1998). Model simulations are normalised to the mean of the total sediment activity at each time period. Solid lines show simu- e e lation for K'd - 2000 I/kg; dotted line for K d = 1000 I/kg; dashed line for K d = 3000 I/kg. (o) Measured values for core (a); (-) measured values for core (b).

The 137Cs activities measured in the sediment pore waters were close to the limit of detection and should be treated with some caution. Mean activity in the pore water was Ca. 4 mBq.H leading to an in situ Ka value of approximately 7 x 104 l.kg-1. There was no discernable change in Kd with depth. The fraction of solid sorbed activity which was exctractable with NHiAc solution (the "exchangeable fraction") was approximately 7% and again showed no dependence on depth.

84 STUK-A145

500 m

450

400 f\ Chernobyl Cs-137 350 1 / o NWTCs-137 300 ! i i - model fit line for 250 I O Chernobyl Cs-137 i - model fit line for NWTCs-137 200- i ' o \o Cs-13 7 (Bq/kg ) j w 150 -

100 6 \ o 50 J o \ • ° 0 -i c 5 10 15 20 Depth (cm)

Figure 4. Activity-depth profiles from in situ cores (Smith et al. 1998). Activity was partitioned between weapons test and Chernobyl using 134Cs measure- ments.

Modelling: experimental measurements.

If radiocaesium were gradually being fixed to sediment solids during our ex- periments we would observe a gradual slowing of the rate of diffusion of activ- ity through the sediment. Contrary to expectations (from Cs sorption times- cales in illite of ca. 100 days, Comans & Hockley 1992), this slowing effect was not observed. It appears that after a few days, diffusion of activity through these sediments occurs at a uniform rate, with no evidence of "fixation". This conclusion is in agreement with our measurements of Kd and exchangeability which also showed no significant changes over time.

85 STUK-A145

These observations greatly simplify the modelling of 134Cs transport in our experiments. We assume that transport of activity through the sediments can be modelled using an equilibrium distribution coefficient, Kd which is con- stant with time and depth. The removal of activity from the overlying water and transport within the sediments (solid and aqueous phases) is then mod- elled using a numerical solution of Eqs. (6-8) with kf, kb equal to zero. Assum- ing porosity, <(>, and solids concentration, s, equal to their mean values of 0.93 and 0.1 g.cnr3 respectively, we have modelled the transport of 134Cs in our ex- periments for different values of boundary layer thickness, z, and distribution coefficient, Kd (Figs. 2 & 3). As shown in the figures, values of z = 0.5 cm and 1 Kd = 2000 l.kg- (equal to the mean of measured Kd values) give a good corre- lation of model simulations with observations in the overlying water and in the solid and aqueous phases of the sediments.

Modelling: in situ measurements.

Eq. 3 was fitted to the activity-depth profiles resulting from the weapons test and Chernobyl fallout events (Fig. 4), as shown in Table I. If the mobility of 137Cs in the in situ measurements could be attributed wholly to diffusion, these effective diffusion coefficients, D, would translate, using Eq. 2, to Kd values of 1.5 x 104 and 1.3 x 104 l.kg-1 for Chernobyl and weapons test 137Cs, respec- tively. In fact, it is likely that these sediments have undergone significant bio- logical mixing: oligocheate worms have been found at a density of up to several thousand per square metre in Esthwaite sediments (Reynoldson 1983). The in situ Kd values we have determined from the activity-depth profiles are there-

Table I. Comparison of experimental and in situ measurements on diffusion of 137Cs in Esthwate Water bottom sediments (Smith et al. 1998).

Experimental In situ Chernobyl weapons test Ka (measured) 2 x 103 I/kg 7 x 104 I/kg - D (measured) 1.7 cm^y1 0.25 crn^y1 0.28 cm^y1 3 4 Kd (modelled) 2 x 10 17kg > 1.5 x 10" I/kg > 1.3 x 10 l/kg Exchangeability 65% 7% - Velocity, co - 0.58 cm/y 0.32 cm/y

86 STUK-A145

fore likely to be significant underestimates since Eq. 2 assumes that all of the transport is by pore water diffusion. This is further demonstrated by the very high Kd values obtained from the in situ pore water measurements (ca. 7 x 104 1-kg-i).

Role of direct diffusion in the removal of activity from Esthwaite Water after Chernobyl.

The removal of activity from the lakewater, Ca, can be described by the follow- ing equation (see, for example, Smith et al. 1997): (9) dC C f v Do(l-f ) a _ a P P n Pn C — c - dt T & a ° where Tw is the water residence time of the lake, fP is the fraction of activity sorbed to suspended participates: (10) s'K, d

and Vp is the participate settling velocity (m.d1) and s' denotes the solids con- centration of suspended matter in the lakewater (cf. "s", the solids concentra- tion of the bottom sediments).

Using the model for transport of radiocaesium within the bottom sediments outlined in Eqs. 4-7, we have modelled the removal of activity, Ca, from Esthwaite Water following a event such as that from the Cher- nobyl accident.

Assuming that the rate of transfer to the sediments by settling of suspended particles is zero, the removal of activity from the lakewater (defined by Eq. 9) is determined by the thickness of the benthic boundary layer, z, and the distri- bution coefficient, Kd, in the bottom sediments. For mean measured values of porosity, ((>, and solids concentration, s, of 0.93 and 0.1 g.cnr3' respectively, we have calculated the change in lakewater activity assuming only lake flushing and diffusion into the sediments. Independent measurements of the boundary layer thickness, z, (Santschi et al. 1983; Hesslein 1987; Devol 1987) gave val-

87 STUK-A145

ues in the range 0.05 - 0.2 cm. For our estimation of removal of 137Cs from the water column, we have used a value of 0.05 cm, though it was found that dif- ferent z values within the measured range had little effect on removal rates.

e 1 Fig. 5 shows that the K d value of around 2000 l.kg determined from our ex- periments is far too low to produce the observed rate of removal. It appears that diffusion of activity across the sediment-water interface is responsible only for a small fraction of the removal of activity. Fig. 5 also shows the re- 1 moval of activity under a particle settling model with fP = 0.2 and vj, = 0.7 m.d (means of measured values): clearly, rates of removal via particle settling and lake flushing are sufficient to produce the observed decline in lakewater activ- ity concentrations.

Measured Diffusion -Particle settling

o

so too iso Time since Chernobyl (days)

Figure 5. Removal of Chernobyl activity from Esthwaite Water by diffusion and particle settling models (Smith et al. 1998).

88 STUK-A145

Discussion and Conclusions

Mobility of radiocaesium in sediments

The transport of 134Cs into sediments in our controlled experiments can be very accurately modelled by assuming simple Fickian diffusion of activity in the sediment porewaters, retarded by sorption to solids with equilibrium dis- e 1 tribution coefficient K d of ca. 2000 Lkg- . By measuring the evolution of activ- ity-depth profiles over time in both solid and aqueous phases we have pre- sented direct evidence of this diffusional process. The quantitative agreement e between the model estimated K d value and the Kd measured in the experi- mental cores (2060 ± 740 l.kg"1) is excellent.

Comparison of the experimental measurements and model results with those made in situ on weapons test and Chernobyl radiocaesium show remarkable differences. As shown in Table 1, in situ Kd measurements are more than one order of magnitude lower than those measured in the experiments and, simi- larly, the in situ exchangeable fraction is one order of magnitude lower than that measured in the experiments. As a result (in spite of mixing effects), the mobility of activity in situ is significantly lower than observed in the experi- ments. This difference in mobility and speciation cannot be explained by the longer contact times of weapons test (ca.28 years) and Chernobyl (5 years) radiocaesium compared to our experiments since we have observed no pro- gressive "fixation" of 134Cs in the experimental cores.

We believe that the explanation for this discrepancy lies in the different mechanisms by which radiocaesium is deposited to the sediments. In our ex- periments, all of the 134Cs diffuses into the sediment as a free ion in solution. Within the sediment itself, it diffuses relatively rapidly and remains available for exchange with NH4+, possibly because "fixation" is hindered by high ammo- nium concentrations (as suggested by Comans & Hockley 1992, Comans et al. 1996).

In contrast, in the lake, the majority of the radiocaesium is transported to the sediment already attached to settling particles. The much lower mobility and availability of in situ caesium implies that the majority of the deposit must have been strongly bound to solids at the time of input to the sediments. This expla- nation is consistent with observed "two-phase" profiles in sediments, but the

89 STUK-A145

assumption of differing fixation rates between caesium sorbed in the catchment or water column and caesium sorbed in the bottom sediment remains to be tested.

Implications for dating studies

The Chernobyl and weapons test peaks in the 137Cs activity-depth profiles are commonly used as markers for dating lake sediments (see, for example, Ap- pleby 1997, this publication). Our studies have confirmed that dating sedi- ments by "first appearance" of 137Cs is not reliable since diffusion of 137Cs into older sediments than those which were at the surface at the time of fallout is likely. Our observations on diffusion rates have, however, shown that the peak in activity will not alter significantly over time since much of the deposited 137Cs is strongly bound to sediment solids. Assuming that diffusion of the weapons test peak will continue in the future at a rate determined by the in situ Kd measurements, we have calculated the predicted evolution of the peak over time (Fig. 6). Our assumption of future diffusion rates is justified as fol- lows:

(1) The weapons test peak is now below the zone of mixing, so mixing rates will be negligible;

(2) The majority of the activity is in the "fixed" phase and thus will diffuse at a (maximum) rate determined by the in situ Kd measurements;

(3) We can assume, for the purposes of modelling, that all of the activity is potentially mobile since on the timescales of decades we are here considering, the sorption of 137Cs is reversible (Smith & Comans 1996).

It can be concluded from Fig. 6 that weapons test 137Cs will remain an impor- tant marker for dating. It is likely that its future use will be limited by radio- active decay rather than diffusive mobility.

90 STUK-A145

300

250 o A / \ 200 0

1 \: o NWTCs-137 i r r "s. i i t = 50 years Bq/kg ) a \ ISO • .' i \ ! \ 1 = 100 years ' i \ i o i 100 j > \ o o • \ ! \

/ • \ 50 \ V \ c \ \ 0 • v -v c 10 20 30 40 50 Depth (cm)

Figure 6. Predicted change in weapons test profile over time, ignoring radioac- tive decay.

References

Berner, R. A., 1980. Early diagenesis. Princeton University Press.

Cambray, R.S., Cawse, PA, Garland, JA., Gibson, JAB., Johnson, P., Lewis, G.N.J., Newton, D., Salmon, L., Wade, B.O., 1987. Observations on radioactivity from the Chernobyl accident. Nucl. Energy, 26: 77-101.

Comans, R.N.J., Middelburg, J.J., Zonderhuis, J., Woittiez, J.R.W., De Lange, G.J., Das, H.A., Van Der Weijden, C.H., 1989. Mobilization of radiocaesium in pore water of lake sediments. Nature, 339: 367-369.

Comans, R.N.J. & Hockley, D.E., 1992. Kinetics of cesium sorption on illite. Geochim. Cosmochim. Acta, 56: 1157-1164.

91 STUK-A145

Comans, R.N.J., Hilton, J., Cremers, A., Geelhoed-Bonouvrie, P.A. & Smith, J.T., 1997. Interpreting and predicting in-situ distribution coefficients of radiocaesium in aquatic systems. In: Desmet, G., Blust, R., Comans, R.N.J., Fernandez, J., Hilton, J. & de Bettencourt, A. (eds.), Freshwater and estuarine radioecology, pp. 129-140. Elsevier, Amsterdam.

Crooks, P.R.J., 1991. The use of Chernobyl-derived radiocaesium for dating recent lake sediments. PhD. Thesis, University of Liverpool, UK.

Crusius, J. & Anderson, R.F., 1989. Resolving discordant ^Cs and 210Pb chronologies using laminated lake sediments and additional radionuclides. In: J. P. Vernet (ed.), Heavy Metals in the Environment, 1: 467-470.

Evans, D.W., Alberts, J.J., Clark, R.A., 1983. Reversible ion-exchange fixation of cesium-137 leading to mobilization from reservoir sediments. Geochim. Cosmochim. Acta, 47: 1041-1049.

Kudelsky, A.V., Smith, J.T., Ovsiannikova, S.V., Hilton, J., 1996. The mobility of Chernobyl-derived Cs-137 in a peatbog system within the catchment of the Pripyat River, Belarus. Sci. Tot. Environ., 188: 101-113.

Lerman, A., 1979. Geochemical Processes: Water and Sediment Environments. Wiley, New York.

Li, Y-H. & Gregory, S., 1974. Diffusion of ions in sea water and in deep-sea sediments. Geochim. Cosmochim. Acta, 38: 703-714.

Meili, M. & Worman, A., 1996. Desorption and diffusion of episodic pollutants in sediments: a three phase model applied to Chernobyl I37Cs. Applied Geochem., 11: 311-316.

Reynoldson, T.B., 1983. The population dynamics of the aquatic oligochaeta in the English Lake District. PhD. Thesis, University of Lancaster, UK.

Smith, J.T. & Comans, R.N.J., 1996. Modelling the diffusive transport and remobilisation of 137Cs in sediments: the effects of sorption kinetics and reversibility. Geochim. Cosmochim. Acta, 60, 6: 995-1004.

92 STUK-A145

Smith, J.T., Leonard, D.R.P, Hilton, J., Appleby, P.G., 1997. Towards a generalised model for the primary and secondary contamination of lakes by Chernobyl - derived radiocaesium. Health Physics, 72, 6: 880-892.

Smith, J. T., Comans, R. N.J., Ireland, P. G., Nolan, L., Hilton., J., 1998. The mobility of radiocaesium in lake sediment and transfer across the sediment/water interface: a model based on experimental and in situ measurements from Esthwaite Water, UK. Geochim. Cosmochim. Acta, submitted.

Wauters, J., 1994. Radiocesium in freshwater sediments: a kinetic view on sorption, remobilization and fixation processes. PhD. Thesis, Faculty of Agronomy, Katholieke Universiteit Leuven.

93 STUK-A145

KINETICS AND REVERSIBILITY OF RADIOCAESIUM SOLID/LIQUID PARTITIONING IN SEDIMENTS

Rob N.J. Comans

Netherlands Energy Research Foundation. P.O. Box 1, NL-1755 ZG Petten, The Netherlands

Abstract

The kinetics and reversibility of radiocaesium solid/liquid partitioning in sediments have been reviewed and interpreted in terms of a mechanistic framework. This framework is based on the premise that radiocaesium is al- most exclusively and highly-selectively bound to the frayed particle edges of illitic clay minerals in the sediments. Several processes with distinctly differ- ent rates can be distinguished in radiocaesium sorption to sediments. 2- and 3- box kinetic models can describe both the overall solid/liquid partitioning in sediments and the reversible (exchangeable) and irreversible (non- exchangeable or "fixed") fractions of radiocaesium in sediments over time scales relevant for natural aquatic systems. The obtained rate parameters indicate that reversible partitioning of radiocaesium dominates over the first few days following a contamination event, whereas irreversible kinetics be- comes important over time scales of weeks to months. The slow process, which reduces the exchangeability of sediment-bound radiocaesium over time, is be- lieved to result from a migration of radiocaesium from exchangeable sites on the frayed edges of illite towards less-exchangeable interlayer sites. Long-term extraction of radiocaesium from historically contaminated sediments has given evidence for a reverse (remobilization) process with a half-life of the order of tens of years. These findings suggest that the long-term exchangeability of radiocaesium in sediments may be higher than the few % which is generally assumed.

Introduction: radiocaesium binding in sediments

Freshwater sediments in Europe have generally recorded the Chernobyl acci- dent in April 1986 by a distinct radiocaesium peak. The accident has renewed

94 STUK-A145

the interest in the mobility of radiocaesium isotopes in the aquatic environ- ment, both from a radioecological and from a sediment-dating point of view. The traditional way to express mobility, e.g. in contaminant/radionuclide transport models, is by means of the solid/liquid distribution coefficient or KD (L/kg): (1)

(2) \137Cs+] or KD =_ I lads ~ \mCs*\ I laq

137 + 137 + where [ Cs ]aq and [ Cs ]ads represent the activity of radiocaesium in the aqueous (Bq/L) and participate phase (Bq/kg), respectively. The J&j-model can- not, however, account for differences in binding properties of different sedi- ments, which has resulted in a variability in published /Co-values for radio- caesium of more than 3 orders of magnitude (e.g. Smith, 1993). These large differences hamper the general use of iCo-based sediment transport models.

Studies since the 1960s have shown that Cs is sorbed very selectively by mi- caceous clay minerals, particularly illite, a process which is believed to be analogous to the fixation of potassium in the clay interlayer structures (e.g. Bolt et al., 1963; Sawhney, 1972). The more recent work by Cremers and co- workers (e.g. Cremers et al., 1988) has enabled a quantitative characterisation of the Cs-selective sorption sites in soils and sediments, which are believed to be located on the frayed edges of illite (the "frayed edge sites" or FES). These authors have shown that the interaction of caesium with sediment and soil particles can be described more mechanistically in terms of an ion exchange process: (3) FES-M + Cs*< £-

The relative affinity of the illite surface for Cs+, relative to the competing metal M+, can then be expressed in terms of a selectivity coefficient, KcCs/M:

95 STUK-A145

(4)

where [Cs^gjand [Mp"ES]are the concentrations of caesium and the compet- ing metal M+ adsorbed on the frayed edge sites (eq/kg), [Cs+] and [M+] repre- sent the aqueous concentrations (eq/L), and [FES] is the frayed edge site ca- pacity of the sediment (eq/kg). At the negligible stable caesium concentrations in the natural aquatic environment [Csp^ ]/[FES] -» 0 and [MpES]/[FES] -> 1, hence: (5)

substituting for KD (L/kg) which, for radiocaesium, can be calculated from its activity in the particulate (Bq/kg) and aqueous (Bq/L) phase: (6)

K

8 ' [FES]

(7) hence, KD is given by: K?/M -[FES] or (8) M log[KD] = logK?' [FES]-^\

The major metal ions (M+) competing with radiocaesium on the frayed edge sites on illitic clays in natural freshwater environments, are generally potas- sium (K+; in oxic environments, e.g. in unsaturated soils and surface waters) and ammonium (NH4+, under anoxic conditions in sediments).

96 STUK-A145

Fig. 1 shows a compilation of in-situ .Kb-values from a number of different Euro- pean sediments (Comans et al., 1998) and reveals a single relationship between the (total, see below) radiocaesium KD and the pore-water ammonium concen- tration. This observation is consistent with the above ion-exchange theory in that, among the major ions that compete with caesium for binding sites on illite clays, NH4+ outcompetes K+ in anoxic sediments as it reaches higher con- centrations and is about five times more selectively bound than K+ (De Preter, 1990). Fig. 1 clearly indicates that the data follow a straight line with a slope close to -1.

The intercept KcCs/NIN[FES] can be independently measured when the FES are "isolated" by blocking all other exchange sites in sediments with AgTU (Cremers et al., 1988; De Preter, 1990). Using this procedure, Wauters (1994) has measured i&Cs/Am-[FES] for the sediments in Fig. 1. Based on these val- ues, Fig. 1 includes the theoretical relationship between KD and [NH4+] for the lowest (Ketelmeer) and highest (Esthwaite) KcCs/NH4-\FES] value. The interpre- 137 tation of the radiocaesium ICDS in Fig. 1 clearly shows that Cs in the aquatic environment obeys ion-exchange theory. The ion-exchange model then allows the solid-liquid distribution coefficient (KD) of this radionuclide in sediment transport models to be predicted from environmental variables (i.e. the quantity of FES in the sediments and the pore-water NH4+ concentration), rather than to be erroneously treated as a constant.

Modelling kinetics and reversibility of radiocaesium partitioning

Another important feature of radiocaesium partitioning in sediments is its apparently irreversible behaviour. It has generally been observed that KD in- creases and the exchangeability of radiocaesium bound to aquatic particles decreases, with increasing contact time between the radionuclide and the par- ticles (e.g. Konoplev et al., 1996). This observation has been interpreted as a slow migration of radiocaesium from the frayed edge sites towards the deeper interlayer spaces between the illite layers, from where it cannot easily be re- leased (Evans et al., 1983; Comans et al., 1991). It is, therefore, important to obtain rate parameters for this process in order to enable prediction of the long-term availability of radiocaesium to the aqueous phase, and thus for up- take in the aquatic food chain and/or for further transport.

97 CO 00 C

>

• Ketelmeer, 1987 CJl A Hollands Dlep, 1987 B Efthwalte, 1991 • Esthwalte, 1992 • Devoke, 1992

• Devoke water column, 1992 o Devoke 1992, exch-KD o Ketelmeer 1987, exch-KD A Holl. Dlep 1987, exch-KD

D Esthwalte 1992, exch-KD Kc'[FESJ • 0.65 meq/g (Holland* Dlep) Kc*[FES] • 0.23 meq/g 1 - (Ketelmeer) Kc*[FES) = 0.64 meq/g A O (E(thwaite) Ke'IFES] • 0.24 meq/g (Devoke)

2 3 + log NH4 [ymoles/L]

Figure 1. Total (closed symbols), exchangeable (open symbols), and predicted (lines) iti-situ KD-values for different w-European lake sediments (Comans et al., 1998). STUK-A145

Laboratory sorption measurements of up to 1 month reaction time have shown that the overall kinetics of the sorption process of radiocaesium on illite cannot be described by a single rate, but involves at least a fast and a slower process (Comans et al., 1991). Fig. 2 illustrates that radiocaesium adsorp- tion/desorption on sediments from the lakes Ketelmeer and Hollands Diep in the Netherlands, at different total-Cs concentrations, shows the same features (Comans, 1998). Multiple-process kinetic models are, therefore, needed to ob- tain reliable rate parameters for the interaction of radiocaesium with natural sediments.

Jannasch et al. linearization

A very useful approach to identify the number of distinct processes that con- tribute to the observed overall kinetics and, hence, need to be considered in the development of a suitable kinetic model to evaluate the data, has been pre- sented by Jannasch et al. (1988). These authors derived a linearized equation for kinetic sorption controlled by a number of separate processes. Their analy- sis assumes that only one process controls overall sorption at any time and that each process can be described by a first order reversible reaction:

(9) IMd 1 - \Md ]«1 —In r.. j—)_. -i

where: [Md] = aqueous metal concentration [Md]e = aqueous metal concentration at equilibrium

[Mt] = total metal concentration r'j = forwards rate constant for ith first order process th r'b = backwards rate constant for i first order process

Plotting the left hand side of equation (9) versus time, using arbitrary ap- proximations of [Ma]e, allows the distinct first order processes to be identified. Of course, the slopes and intercepts of the linearizations change for different choices of [Md]e, but the number of distinct processes and the time scales over which they dominate remain the same.

99 STUK-A145

1.00

1 0.80 «

0.60 I * • i • CLO-10i>»M.

• CLO'lpgA.

X CLO • 1 vg/L

0.20 III '

0.00 t " 20 25 time[d] 1.00

0.80

0.60

0.40

I •

0.20 • ill i 0.00 10 t « 20 25 timo[d]

Figure 2. Kinetics of radiocaesium adsorption and desorption on sediments from (top) Ketelmeer and (bottom) Hollands Diep (Comans, 1998). do - con- centration of Cs in solution att = 0; CL - concentration of Cs in solution at time t (i.e. CL/CLO = fraction of total Cs in solution). The arrow indicates the start of desorption.

100 STUK-A145

Fig. 3 shows the linearization applied to the Cs sorption data for the Ketel- meer and Hollands Diep sediments (Comans, 1998). The data show an appar- ently instantaneous reaction and two other distinct processes. As noted by Jannasch et al. (1988), the sampling schedule affects the number of processes revealed by the linearization procedure. In particular, the initial reaction may, with the appropriate experimental technique and sampling schedule, be sub- divided into processes operating on minute or sub-minute time scales. The purpose here, however, is to model Cs sorption kinetics on the longer time scales which are more relevant for natural systems.

The Jannasch et al. linearization implicitly assumes a linear KD relationship between dissolved and sorbed concentrations. Earlier kinetic models, such as that of Nyffeler et al. (1984), also adopt that assumption. However, the data for Ketelmeer and Hollands Diep sediments in Fig. 2 cover a range of initial Cs concentrations for which sorption isotherms are non-linear. This non-linearity implies that /Co-models cannot be used to model sorption kinetics over the whole concentration range. Kinetic models that include non-linear isotherm equations have been used successfully to describe metal mobilities in soil sys- tems (e.g. Amacher et al., 1986; Amacher et al., 1988 and references therein). It should be noted that isotherm non-linearity does not invalidate the above conclusion that a number of distinct processes are active. Fig. 3 clearly shows that the individual processes operate within similar time frames over a wide concentration range.

In view of the similarities between the kinetic sorption behaviour of caesium on illite (Comans & Hockley, 1992) and natural sediments (Comans, 1998; Fig. 2-3), the 2- and 3-box models of Comans & Hockley (1992) have been used to model and obtain rate parameters for the sediment data in Fig. 2 (Comans, 1998).

2- and 3-box models for caesium sorption

The four and five parameter (2- and 3-box) kinetic models of Comans & Hock- ley (1992) have been fitted to the sediment data of Ketelmeer and Hollands Diep (Comans, 1998). No three parameter model could adequately describe these data. The differential equations defining the models were solved using a finite difference method and parameter fitting was done with a non-linear least squares regression package.

101 STUK-A145

3-

2.5

3. 1 I

< «

Urn* [days]

J.5-

— hn1.1 pg/L Iin2.1 lig/L • OX*2ug/l.

Iin2.2ps/L • CI01*10 p^L k -^Iin1. 1Opg/L O.S-

6 S 10 time (day*]

Figure 3. Kinetic data of Cs adsorption on (top) Ketelmeer and (bottom) Hol- lands Diep sediments linearized according to the method of Jannasch et al. (1988); (Comans, 1998).

102 STUK-A145

The first "two-box" model includes a reversible Freundlich reaction followed by an irreversible first order process. The equations defining this model are shown in Fig. 4. There are four independent parameters: the Freundlich pa- rameters, K and n, the reversible rate constant, r, and the irreversible rate constant, k. Initial estimates for the Freundlich parameters were obtained from independent isotherm fits of the data after approximately 2 weeks of adsorption. Fitting the kinetic data by eye provided the initial estimates for r and k.

The two-box model fits to the data are shown in Fig. 5. The parameter values are listed in Table I. The model provides an adequate description of the data, but tends to overpredict radiocaesium sorption during the first 24 to 48 hours, particularly at low total-Cs concentrations. As was needed for Cs sorption on illite (Comans & Hockley, 1992), an additional (fast) reaction may be consid- ered to allow for a better overall prediction of the data points. The apparently irreversible reaction in the model, which predicts readsorption during the desorption phase (period between 13-26 days in Fig. 5), is not reflected in the data points, as was the case for Cs sorption on illite (Comans & Hockley, 1992). Possibly, some ammonium was produced by the breakdown of organic matter in the sediments at the later times during the experiments (sediments were not spiked with bacteriocides). Competition by ammonium for frayed edge sites could have prevented the readsorption of radiocaesium that was noticed in the experiments with illite.

The second model, shown in Fig. 4, provides the additional reaction by split- ting the initial process of the two-box model into an equilibrium reaction and a parallel reversible reaction. Both these reactions tend towards the same Freundlich equilibrium. This model can be conceptualized as dividing a single set of Freundlich sorption sites into a portion which is rapidly accessible and a portion which is kinetically controlled. The Freundlich processes are followed in series by a third, irreversible process. In addition to the four independent parameters of the first model, the three-box model includes /, the fraction of the Freundlich sorption sites reaching equilibrium instantaneously.

Fig. 5 includes the three-box model fitted to the data. The rate parameters obtained on the basis of this model are included in Table I. The fit is slightly better, particularly for the first 48 hours, than that of the two-box model. Again, readsorption is predicted for the desorption period, albeit slightly less than the extent predicted by the two-box model. However, one might argue

103 STUK-A145

= r(K[X]"-M)-kM dt d[Z] dt

Mass Balance:

[Z])Cp = [X]l"JO [ZJ

dt d[Z] dt

Mass Balance:

%]+[r2]+[z])Cp = [x] [z]jcjcp

Figure 4. Schematic representation and equations defining the 2-box (top) and 3-box (bottom) kinetic models (Comans & Hockley, 1992). "X" represents metal in the dissolved state. The shaded boxes represent sorbed metal. The subscript "0" denotes concentrations at time zero.

104 STUK-A145

1.00

• fit 2-boi CLO • 20 IH J-box CLO > 20 0.80 CLO • 10 vail fit 2-Dos CLO * 10 fft 3-box CLO • 10 CLO "5 port. IK 2-box CLO • 5

1K2-OOXCL0.2 fitJ-boxCL0^2 0.40 CLO • 1 port, fit 2-60* CLO-1 lit J-OO1 CLO • 1

0.20

0.00

1.00

CLO

IK Wxm CLO • 20 0.80 CU'1O|ig/L (It 2-ixut CLO » 10 MMoxCLO-10 CLO-Spgn. - IK2-txuCL0-5 O.M (« W»* CLO "5 CL0-2VQA. 1K2-bo«CL0>2

0.40

OJO- -4,--

0.00

Figure 5. 2- and 3-box model predictions of Cs sorption on (top) Ketelmeer and (bottom) Hollands Diep sediments (Comans, 1998). do = concentration of Cs in solution at t = 0; Ct - concentration of Cs in solution at time t (i.e. CL/CLO = fraction of total Cs in solution). The arrow indicates the start of desorption.

105 STUK-A145

whether the addition of a fifth fitting parameter in the model is statistically justified in this case.

The obtained rate parameters shown in Table I are very similar, for both models, to those obtained for Ca-saturated illite by Comans & Hockley (1992). The sorption experiments with Ketelmeer and Hollands Diep sediments (Figs. 2, 3 and 5; Comans, 1998) were also carried out in a Ca-environment. If we use the three-box model for the best fit to the early data points, the fast reversible process (r) has a half life of a few (1-4) days. The slower, irreversible, process (k) has a half life of approximately 20-50 days.

The above models should not be used to predict Cs-sorption reversibility over time frames of more than a few months. Observations that radiocaesium, which has been in contact with contaminated sediments for many years, can still be mobilized by enhanced ammonium concentrations (Evans et al., 1983; Comans et al., 1989) give evidence for a backwards reaction. This reaction, which is interpreted as a remobilization of radiocaesium from edge-interlayer sites on illite, is apparently too slow to be observed in laboratory experiments of many weeks. Therefore, extraction of radiocaesium from historically con- taminated sediments have been used to obtain a first estimate of this slow remobilization rate.

Table I. Parameter values of the 2-box and 3-box kinetic models obtained after fitting the sorption data of Cs on sediments from Ketelmeer and Hollands Diep (Comans, 1998).

Ketelmeer Hollands Diep K* (2-box) 7.82± 0.404 15.34± 0.64 n* 0.5076± 0.0256 0.4160± 0.0185 r (d-i) 8.55300± 2.2470 5.8090± 0.9480 k (d-i) 0.04037± 0.00711 0.01736± 0.00448 RSS 8.92E-02 8.84E-02 K* (3-box) 10.25± 4.22 16.36± 1.00 n* 0.4976± 0.0256 0.4071± 0.0172 f 0.669± 0.2552 0.6757± 0.0695 r (d-0 0.1558± 0.2495 0.9964± 0.6226 k (d-i) 0.02543± 0.01969 0.01314± 0.00511 RSS 8.69E-02 6.93E-02 "Freundlich const, using ng/L and n-g/g for aqueous and sorbed cones., resp.

106 STUK-A145

Exchangeability and slow remobilization of radiocae- sium in sediments

Exchangeability of sediment-bound radiocaesium

The exchangeability of sediment-bound radiocaesium has been investigated by Comans et al. (1998) by extracting sediments with 0.1 M NH4-acetate. These extractions have been performed on all sediments shown in Fig. 1. Sediment samples from different depths were extracted three times sequentially, each step for 24 hours, with 0.1 M NH4-acetate at a liquid/solid ratio of 10 L/kg. Further details on the extraction procedure and radiocaesium measurements are given in Comans et al. (1998).

The exchangeable amounts of 137Cs are given in Table II. The amount of 137Cs that can be released by the three sequential NH4-extractions from the more mineral sediments of Hollands Diep, Ketelmeer and Esthwaite is very low; on average 2, 3, and 7%, respectively. The more organic-rich sediments of Devoke show a much higher exchangeability of 16%. The values for the three mineral sediments are low when compared, for instance, with those of Evans et al. (1983), who have measured 137Cs exchangeabilities (also in 0.1 M NKU+) of 10- 20% in the sediments of the Par Pond reservoir, 15-20 years after contamina- tion. These authors attribute their relatively high exchangeabilities to the high content of the Par Pond sediments. Western European sediments generally contain illite as the major clay mineral, which likely causes the strong fixation of radiocaesium that has been observed in the Hollands Diep, Ketelmeer and Esthwaite sediments. Devoke apparently behaves more like the Par Pond sediments in that similar amounts of 137Cs are exchangeable.

If we accept that radiocaesium in all four of the above sediments is bound solely to frayed edge sites on illitic clays in the sediments, the up to one order of magnitude differences between the exchangeable fractions of radiocaesium in these sediments are unexpected. The differences are clearly not related to the different contact times of radiocaesium with the sediments, as the lowest values are found for the sediments sampled only 1.5 years after the Chernobyl accident (Hollands Diep and Ketelmeer).

107 STUK-A145

Table II. Total and exchangeable radiocaesium in the different sediment cores (Comans et al., 1998).

Sediment Depth total exchange- exchange- »7Cs able 137Cs able 137Cs [fraction [cm] [Bq/kgl [Bq/kg] of the total] Hollands Diep 0.5 249.15 0.85 0.003 1.5 269.79 0.69 0.003 3.5 83.05 0.55 0.007 6.5 5.00 0.50 0.100 9.5 17.90 . 16.5 46.31 0.51 0.011 24.5 47.96 1.36 0.028 31.5 21.31 5.01 0.235 40.4 21.70 . average' ± s.d.: 0.022 ±0.038 Ketelmeer 1 115.46 10.06 0.087 3 124.73 4.33 0.035 5.5 224.23 5.33 0.024 8.5 59.07 0.27 0.005 11.5 37.466 0.76 0.020 18.5 64.41 0.61 0.009 25.5 51.79 14.89 0.288 31.5 9.44 2.94 0.312 average* ± s.d.: 0.030 ±0.030 Esthwaite 0.5 214.80 13.60 0.063 2.5 330.73 22.12 0.067 4.5 545.61 32.31 0.059 6.5 474.14 38.44 0.081 8.5 353.43 24.93 0.071 10.5 345.36 18.06 0.052 12.5 325.69 21.99 0.068 14.5 266.41 20.31 0.076 16.5 164.03 15.03 0.092 18.5 42.73 9.329 0.218 20.5 21.59 4.791 0.222 average" ± s.d.: 0.070 ±0.012

108 STUK-A145

Table II cont.

Sediment Depth total exchange- exchange- '"Cs able 137Cs able 137Cs [fraction [cm] IBq/kg] IBq/kg] of the total] Devoke 0.5 1813 343.78 0.190 1.5 1860 401.78 0.216 2.5 2142 345.69 0.161 3.5 2598 375.72 0.145 4.5 3096 363.83 0.118 5.5 2579 270.93 0.105 6.5 1524 208.66 0.137 7.5 1023 237.20 0.232 8.5 848 174.72 0.206 9.5 727 181.78 0.250 10.5 644 140.63 0.218 11.5 653 44.73 0.068 12.5 719 108.70 0.151 13.5 685 94.97 0.139 14.5 588 81.80 0.139 15.5 392 58.42 0.149 16.5 244 103.70 0.426 17.5 146 81.32 0.558 18.5 130 68.18 0.523 19.5 106 56.20 0.531 average* ± s.d.: 0.164 ±0.050 "average and s.d. values exclude the (unexplained) high exchangeabilities in the bottom sections of each core (indicated in italics)

"Exchangeable" iCo-values for radiocaesium in the sediments, calculated from the exchangeable rather than total amount of 137Cs in the sediments, are in- cluded in Fig. 1. We would expect "exchangeable" KDS to correspond better with values predicted on the basis of (short-term) laboratory measurements of KcCs/NH4-[FES] than the total KDS. Although this may be the case for the De- voke sediments, total and exchangeable iiCo-values for Esthwaite correspond about equally with the predictions, whereas the "exchangeable" KDS for Hol- lands Diep and Ketelmeer deviate much more from the predicted values than the total KDS. These observations strongly suggest that short-term exchange- ability measurements of radiocaesium, especially in the more mineral sedi- ments, underestimate the amount of the radionuclide that is actually taking

109 STUK-A145

part in ion-exchange with the pore waters and, hence, is available for remobi- lization by high concentrations of ammonium.

Slow (reverse) migration of radiocaesium from clay-mineral interlayers into solution

In the kinetic models discussed above, a reverse process of radiocaesium re- mobilization from interlayer sites was not considered. The equilibration times of up to 4-weeks were too short for the reverse process to become apparent. Nevertheless, the fact that radiocaesium in sediments is still exchangeable to a certain extent after more than 20 years of contact with sediments (Evans et al., 1983), indicates that such a reverse process must exist.

Because of its relevance for the long-term availability of sediment-bound ra- diocaesium, the long-term release of particle-bound radiocaesium has been studied in more detail (Comans et al., 1998). The primary objective of this part of the study was to estimate the existence and magnitude of a slow remobili- zation (reverse rate) of radiocaesium from clay mineral interlayers during contact with a high concentration of competing ions. After the three sequential 0.1 M NH-f-acetate extractions, the sediments were resuspended for a fourth time in a fresh 0.1 M NH-i-acetate solution and have been allowed to equili- brate for more than one year (400-842 days). After this period, the additional amount of extracted radiocaesium was measured.

Table III shows the average fraction of exchangeable-137Cs in sediments after 3 sequential 24-hr NH4-extractions and the additional fraction mobilized after the 4th, long-term (400-842 days) extraction. Assuming (1) that all (rapidly) "exchangeable" radiocaesium had been removed by the three prior extractions, and (2) a first order remobilization process, we can roughly calculate the re- verse rate constant that describes the slow remobilization of 137Cs from the sediments. Table III indicates that the half-life of this reaction, which is inter- preted as the slow release of radiocaesium from the interlayer sites, is of the order of 30-100 y1. These findings suggest that radiocaesium on interlayer sites, which is generally referred to as being "fixed", is not truly immobilized but can, at least partly, be very slowly remobilized.

110 STUK-A145

Table HI. Average fraction ofexchangeable-13'1TCs in sediments after 3 sequen- tial 24-hr NH4-extractions and the additional fraction mobilized after a 4th 400/560-d extraction. A reverse rate constant and half-life for the slow remobi- lisation of 137Cs from the sediments has been calculated on the basis of the 4th extraction, assuming a first order process (Comans et al., 1998).

Sediment exch. 137Cs after additional exch. reverse tl/2 3x 24-hr extraction IS7Cs after 400- rate 842-d extraction* constant [fraction of the [fraction of the total] total] fy-'l fyl Hollands 0.022 ± 0.038 0.0089 ± 0.0050 0.0082 85 Diep Ketelmeer 0.030 ± 0.030 0.0095 ± 0.0030 0.0087 79 Esthwaite 0.070 ± 0.012 0.0156 + 0.0053 0.0068 102 Devoke 0.164 ±0.050 0.0380 + 0.0355 0.0220 31 'equilibration time 4th extraction = 400 days for Hollands Diep & Ketelmeer; 560 days for Devoke; 842 days for Esthwaite.

Comparison of rate parameters with other estimates

Slow uptake rates

As was mentioned above, the slow uptake of radiocaesium by illite, and sedi- ments containing illite, is interpreted as a migration towards interlayer sites on this clay mineral. This process has been recognized for many years to cause the so-called "potassium fixation" in (agricultural) soils. De Haan et al. (1965) have studied the kinetics of that process by following the 40K-uptake from so- lution by illite clay over a period of 16 months. The removal of the radiotracer from solution was interpreted as isotopic exchange with interlayer-potassium, i.e. as a slow diffusion of 40K into the illite lattice. Re-evaluated in terms of a first order uptake process, their data indicate a rate constant for the interlayer migration of the order of 10-3 d-1. This value is very similar to the values found by Comans & Hockley (1992) for Cs-sorption on illite in a K-dominated envi- ronment, which lends support to the hypothesis of a similar binding mecha- nism for Cs and K. The slow Cs-uptake rate was found to be approximately an order of magnitude faster for a Ca-illite, which is supported by the values

111 STUK-A145

found for the illite-containing sediments of Ketelmeer and Hollands Diep in a Ca-solution (Table I). Comans & Hockley (1992) have shown that the slow interlayer migration of caesium proceeds faster when the edges of illite are expanded by reaction with calcium, relative to the collapsed structure of a K- saturated illite.

Evans et al. (1983) have studied the "fixation" of 134Cs by freshwater sediments over a period of 180 days. The increasingly incomplete extractability of radio- caesium, using 0.1 M NEU, with Cs-sediment contact time was interpreted as a slow movement of radiocaesium to clay interlayer sites in the sediment. If the Evans et al. (1983) data are interpreted in terms of a first order process, a rate of 10-2-103 d1 is obtained, which is again of the same order as the values found for illite (Comans & Hockley, 1992) and for Hollands Diep and Ketelmeer sediments (Table I).

Finally, Nyffeler et al. (1984) have modelled radiocaesium uptake by marine sediments, using a multi-reaction model with an irreversible process similar to that in the 2- and 3-box models of Fig. 4. These authors have also obtained a similar rate of 102 d*1.

Reverse (remobilization) rates

Smith & Comans (1996) have developed a transport model that includes radio- caesium sorption kinetics to simulate radiocaesium in each of three phases in the sediment profiles of Ketelmeer and Hollands Diep: aqueous, exchangeably- bound and slowly reversible (often termed "fixed"). The model gave evidence for a reverse reaction from less-exchangeable to exchangeable sites with a half- life of order 10 years, which is close to the independent estimates by the long- term extractions in Table III.

Using the exchangeability measurements of Evans et al. (1983), 85% fixed/15% exchangeable (after 180 days contact of 134Cs with the sediment), and a for- ward rate constant of half-life 100 days, a reverse rate of half-life 2 years is estimated. The same estimate can be made from the short-term exchangeabil- ity measurements of Comans et al., 1998) shown in Table II (84-98% fixed/2- 15% exchangeable). It is reasonable to assume that the long-term extractions discussed above have led to an upper estimate of the remobilization rate, as a new equilibrium between NH4 and remobilized Cs may have been reached dur- ing the single long-term extraction step. Given the limited available data, the

112 STUK-A145

remobilization half-life of sediment-bound radiocaesium may, therefore, be expected to be of the order of a few years to a few tens of years.

Conclusions

The kinetics and reversibility of radiocaesium solid/liquid partitioning in sediments have been reviewed and interpreted in terms of a mechanistic framework. This framework is based on the premise that radiocaesium is al- most exclusively and highly-selectively bound to the frayed particle edges of illitic clay minerals.

Several processes with distinctly different rates can be distinguished in radio- caesium sorption. 2- and 3-box models can describe both the overall partition- ing in sediments and the reversible (exchangeable) and irreversible (non- exchangeable/"fixed") fractions of radiocaesium over time scales relevant for natural aquatic systems. The models are consistent with knowledge of radio- caesium sorption mechanisms in sediments. The obtained rate parameters indicate that reversible partitioning of radiocaesium dominates over the first few days following a contamination event, whereas irreversible kinetics be- comes important over time scales of weeks to months. The slow process, which reduces the exchangeability of sediment-bound radiocaesium over time, is be- lieved to result from a migration of radiocaesium from exchangeable sites on the frayed edges of illite clays towards less-exchangeable interlayer sites. This hypothesis is supported by comparison of model results with previous findings.

Observations that radiocaesium, which has been in contact with contaminated sediments for many years, can still be mobilized have given evidence for a backwards reaction. This reaction is interpreted as a remobilization of radio- caesium from interlayer sites on illite, and is apparently too slow to be ob- served in laboratory experiments of many weeks. Long-term extraction of ra- diocaesium from historically contaminated sediments has confirmed the exis- tence of a reverse reaction and has provided a first estimate of the half-life of this slow remobilization rate of the order of tens of years. These findings sug- gest that the long-term exchangeability of radiocaesium in sediments may be higher than the few % which is generally assumed.

113 STUK-A145

References

Amacher, M.C., Kotuby-Amacher, J., Selim, H.M. & Iskandar, I.K., 1986. Re- tention and release of metals by soils - evaluation of several models. Geoderma, 38: 131-154. Amacher, M.C., Selim, H.M. & Iskandar, I.K., 1988. Kinetics of chromium(VI) and cadmium retention in soils; a nonlinear multireaction model. Soil Sci. Soc. Am. J., 52: 398-408. Bolt, G.H., Sumner, M.E. & Kamphorst, A., 1963. A study of the equilibria between three categories of potassium in an illitic soil. Soil Sci. Soc. Am. Proa, 27: 294-299. Comans, R.N.J., 1998. Kinetics and reversibility of radiocesium sorption on illite and illite-containing sediments. In: Sparks, D.L. & Grundl, T. (eds.) Ki- netics and Mechanisms of Reactions at the Mineral/Water Interface. ACS Symposium Series (in press). Comans, R.N.J., Middelburg, J.J., Zonderhuis, J., Woittiez, J.R.W., De Lange, G.J., Das, H.A. & Van Der Weijden, C.H., 1989. Mobilization of radiocaesium in pore water of lake sediments. Nature, 339: 367-369. Comans, R.N.J., Haller, M. & De Preter, P., 1991. Sorption of cesium on illite: non-equilibrium behaviour and reversibility. Geochim. Cosmochim. Acta, 55: 433-440. Comans, R.N.J. & Hockley, D.E., 1992. Kinetics of cesium sorption on illite. Geochim. Cosmochim. Acta, 56: 1157-1164. Comans, R.N.J., Hilton, J., Cremers, A., Bonouvrie, P.A. & Smith, J.T., 1998. Predicting radiocaesium ion-exchange behaviour in freshwater sediments, (submitted). Cremers, A., Elsen, A., De Preter, P. & Maes, A., 1988. Quantitative analysis of radiocaesium retention in soils. Nature, 335: 247-249. De Haan, F.A.M., Bolt, G.H. & Pieters, B.G.M., 1965. Diffusion of potassium- 40 into an illite during prolonged shaking. Soil Sci. Soc. Am. Proa, 29: 528-530. De Preter, P., 1990. Radiocesium retention in the aquatic, terrestrial and ur- ban environment: a quantitative and unifying analysis. Ph.D Thesis, Faculty of Agronomy, Katholieke Universiteit Leuven, Belgium.

114 STUK-A145

Evans, D.W., Alberts, J.J. & Clark, R.A., 1983. Reversible ion-exchange fixa- tion of cesium-137 leading to mobilization from reservoir sediments. Geochim. Cosmochim. Acta, 47: 1041-1049. Jannasch, H.W., Honeyman, B.D., Balistrieri, L.S. & Murray, J.W., 1988. Ki- netics of trace element uptake by marine particles. Geochim. Cosmochim. Acta, 52: 567-577. Konoplev, A.V., Bulgakov, A.A., Hilton, J., Comans, R.N.J. & Popov, V.E., 1996. Long-term investigation of 137Cs fixation by soils. Radiation Protection Dosimetry, 64: 15-18. Nyffeler, U.P., Li, Y.-H., & Santschi, P.H., 1984. A kinetic approach to describe trace-element distribution between particles and solution in natural aquatic systems. Geochim. Cosmochim. Acta, 48: 1513-1522. Sawhney, B.L., 1972. Selective sorption and fixation of cations by clay miner- als: a review. Clays Clay Miner., 20: 93-100. Smith, J.T., 1993. Mathematical modeling of 137Cs and 21«Pb transport in lakes, their sediments, and the surrounding catchment. Ph.D Thesis, University of Liverpool, UK. Smith, J.T. & Comans, R.N.J., 1996. Modelling the diffusive transport and remobilization of 137Cs in sediments: the effects of sorption kinetics and re- versibility. Geochim. Cosmochim. Acta, 60: 995-1004. Wauters, J., 1994. Radiocaesium in aquatic sediments: sorption, remobilization and fixation. Ph.D Thesis, Faculty of Agronomy, Katholieke Universiteit Leu- ven, Belgium.

115 STUK-A145

TRACING THE CHERNOBYL FALL-OUT PEAK IN FINNISH LAKE SEDIMENTS IN ORDER TO OBTAIN A GOOD TIME MARKER

Hogne Jungner

Dating Laboratory, University of Helsinki, POB 11, FIN-00014 Helsinki, Finland

Abstract

The fall-out 137Cs from the Chernobyl accident has been traced in sediments from a few lakes in Finland in order to obtain a time marker in the uppermost sediment layer. A beta multicounter system developed at Riso was used for detection of the radionuclides. The detector system is based on GM counters and has therefore no energy resolution. It is thus not possible to identify dif- ferent nuclides with the counting system. The low background and good effi- ciency of this detector make it possible to use small amount of sample and thus good depth resolution can be obtained for the sediment profile. A clear advan- tage is that practically no sample preparation is needed. Beside providing a time marker for year 1986 in a sediment, the activity depth profile obtained can provide information about sedimentation conditions and the removal of the fall-out cesium from the lake water. The influence from natural radionu- clides in the sediment material, however, puts a limit to the lowest detection level, and makes detailed mobility studies difficult. For direct time marking the beta counter is, however, an effective tool.

Introduction

From the Chernobyl accident on April 26, 1986 fall-out was observed over large areas of the northern hemisphere. In Finland active investigations were done in the period after the accident and the distribution of different nuclides over Finland is well known (Arvela et al. 1990). Today about eleven years after the accident, the radioactive elements still left in the nature are a few long- lived nuclides. Amongst these Cs-137 is the one, which is easy to detect due to its quite high-energetic gamma and beta ray emission. The activity left can today be used as a good time marker in different sediments in areas where it

116 STUK-A145

can be detected. The idea is well known from the atomic bombtest period in the 1950s and 1960s. Compared to the fall-out from the bomb tests, the Chernobyl fall-out is very sharply defined in time to a few days in the spring of 1986.

Methods

For detection of the Cs in the sediments we used a multicounter system devel- oped at Riso National Laboratory in Denmark (Better-Jensen and Nielsen 1989). This counter system based on GM counters is widely used in many labo- ratories for environmental studies. The counter system consists of five sepa- rate GM detectors surrounded by an anticoincidence detector. The counters have a low background (around 0.2 cpm) and reasonably high efficiency, and the system is therefore well suited for low-activity measurements of environ- mental samples.

A practical advantage is that almost no sample preparation is needed. Dried samples are directly packed in plastic holders. The amount of sample used range from 100 milligrams up to 1 gram. This small amount allows thin slices to be cut out from a sediment core and thus good depth resolution can be ob- tained. The normal counting time used was 1300 minutes giving a statistical uncertainty of about 1 %. Because the counter can take five samples in parallel an activity profile can be obtained in a few days.

Because the counter operates in the GM region no energy discrimination can be set except a low level treshold. This means that not only beta particles from Cs decay but also from different radionuclides as 40K and members in the uranium and thorium decay series in the sediment are detected. In all cases studied the activity from the Chernobyl fall-out can well be detected above the natural activity level, but the activity from the bomb test period is more diffi- cult to distinguish.

Because of the lack of energy resolution and the self absorption in the sample no absolute activity value for the samples were attempted to be given. The activity found in lake sediments from different regions of Finland reflect, how- ever, the general fall-out pattern obtained by direct measurements after the accident.

117 STUK-A145

I

5 10 Count rate (cpm)

Fig. 1. The beta count rate for the uppermost 45 sediment varves from lake Pohjajarvi, Central Finland. The activity of 11 cpm is clearly detected in the varve for the year of the accident. The instrumental background is 0.2 cpm.

Results and Discussion

In Fig. 1 the beta activity measured from a laminated lake sediment from a small lake Pohjajarvi in Central Finland is plotted for each annual varve. The mean varve thickness is about 3 mm and the 45 varves given cover a total thickness of ca 15 cm (Saarinen 1996). The activity of 11 cpm is clearly de- tected in the varve for the year of the accident. For the varves from number 45 up to this level the observed count slowly increases from about 1.5 to about 2.5 cpm. The varves number 40-30 compare to the bomb test period but the activ- ity observed is fairly close to the natural activity level. The activity in the varve for the year after the accident has already dropped to about 20 % of the maximum. In this lake the cesium was thus quite effectively removed from the water and fixed to the sediment. In other cases we have observed a much slower removal of cesium to the sediment.

The profile from lake Pohjajarvi represents a very clear case. For other lakes studied so far, we have cases where the activity peak is well defined but also

118 STUK-A145

cases where the activity maximum is very wide. The problem with sampling of very loose surface sediment can also be observed in some cases.

In sediment profiles, where a well defined activity maximum can be observed, this can be taken to represent the time of the accident as the influence of ce- sium mobility in the sediment during the ten years past the accident should not considerably influence the position of the maximum (Crusius and Ander- son 1995, and papers presented in these proceedings).

Beside providing a time marker for year 1986 in a sediment, the activity depth profile obtained can provide information about sedimentation conditions and the removal of the fall-out cesium from the lake water. The influence from natural radionuclides in the sediment material, however, puts a limit to the lowest detection level, and makes detailed mobility studies difficult. For direct time marking the beta counter is, however, an effective tool.

References

Arvela, H., Markkanen, M. and Lemmela, H., 1990. Mobile survey of environ- mental gamma radiation and fall-out levels in Finland after the Chernobyl accident. Radiation Protection Dosimetry, 32, 3: 177-184.

Botter-Jensen, L. and Nielsen, S.P., 1989. Beta multicounter systems for low- level measurements of environmental samples. In: W. Feldt (ed.). The radio- ecology of natural and artificial radionuclides. Proceedings of the XVth Re- gional Congress of IRPA, Visby, Gotland, 459-464.

137 Crusius, J. and Anderson, R.F. 1995. Evaluating the mobility of Cs, 239+24opu and 210Pb from their distributions in laminated lake sediments. Journal of Paleolimnology, 13: 119-141.

Saarinen, T., 1996. Paleomagnetic study of annually laminated sediments in Lake Pohjajarvi: Paleosecular variation in Finland during the last 3200 years. Geological Survey of Finland, Report Q29.1/96/2.

119 STUK-A145

RADIONUCLIDES IN SEDIMENT CORES FROM THULE, GREENLAND, 1991

Sven P. Nielsen and Henning Dahlgaard

Rise National Laboratory, Denmark

Introduction

A number of unique sediment cores taken in 1991 in the Thule accident area (North West Greenland) have been made available in collaboration with the Bedford Institute of Oceanography, Canada. The cores have been sectioned in 1-cm slices. The work on the sediment cores has focused on the following items: • to analyse the cores for 210Pb by gamma spectrometry

• to analyse the cores for 239+24opu by alpha spectrometry • to compare 210Pb and transuranics as tracers for sedimentation and biotur- bation processes • to determine sedimentation rates and bioturbation parameters of interest in connection with radiological assessments in the Arctic marine environ- ment.

Material and methods

Samples of marine sediments were collected in 1991 from Greenland by the Bedford Institute of Oceanography, Canada, and shared with Riso National Laboratory. The sampling covered the plutonium contamination of the Thule area where a US military aircraft carrying nuclear weapons crashed on the ice in January 1968. Sediment cores (0-50 cm) were sampled and sliced in 1-cm sections. Risa carried out radiometric analyses on 10 cores from the near-zone (within 10 km) of the point of impact (Fig. 1) and on 5 cores from the northern part of Baffin Bay. The analyses comprised gamma-spectrometric determina- tions of natural radionuclides including 210Pb and the fallout radionuclide 137Cs, and radiochemical isolation of the plutonium isotopes followed by the

detection of 239+24opu Dy alpha spectrometry.

120 STUK-A145

Thule 1991. Sample locations

TTN 77°N

76'N 76°N

CO

Figure 1. Locations (station #) for the Thule 1991 sediment samples. ~k: point of impact.

The vertical profiles of the radionuclide concentrations were interpreted in order to obtain information on sediment characteristics of importance for the evolution with time of the plutonium contamination in the Arctic marine envi- ronment. 210Pb in the sediment originates from the natural content of 226Ra and decay products (supported fraction), and from the flux of 210Pb on particles settling from the water column on surface sediments (unsupported fraction).

121 STUK-A145

The excess 210Pb in surface sediments thus contains a historical record of sedimentation within about two hundred years limited by the physical half life of 210Pb of 22 y. An undisturbed sediment profile will show an exponentially decreasing concentration of 210Pb where the rate of decrease is determined by the rate of sedimentation and the half life of 210Pb. Deviations from the expo- nential decrease of 210Pb in the surface sediments may be interpreted in terms of mechanical mixing due to biological activity and water flow generated by wind and tidal forces. It is important to understand and quantify the mixing processes because they speed-up and enhance the transfer of radionuclides from the sediment back to the water column from where they may more read- ily enter into human food chains.

A mathematical compartmental model for the activity of radionuclides in sediments was set up considering sedimentation and mixing. The model was based on the work by Christensen (1982). The model assumes that mixing may be expressed as a diffusive process with gaussian mixing in the upper layer. The diffusion coefficient is thus assumed to decrease with depth from a maxi- mum value at the water-sediment interface following a gaussian function.

where Do (cm2^1) is the diffusion coefficient (mixing rate) at the sediment sur- face, z (cm) the depth, and a (cm) the effective mixing depth. There are biologi- cal reasons for choosing a gaussian diffusion coefficient, since deposit-feeding animals, like worms, are mainly found in the top sediments and their distri- butions with depth are close to gaussian. Furthermore, physical mixing gener- ated by water movement also declines rapidly with depth.

For unsupported 210Pb, the model simulates continuous deposition for about two hundred years until 1991, and for plutonium isotopes the model starts with a single input of 239+24opu in IQQS followed by a 23-year simulation with- out input until 1991. The depth profiles calculated with the model are com- pared with the observed profiles and the model parameters are selected to minimise the difference.

The parameters used by the model for each sediment core include the sedi- mentation rate (cm y1), the mixing depth (cm), the mixing rate (cm2 y1), the

122 STUK-A145

210 2 1 2 continuous Pb input rate (Bq m- y ) and the single 239+24opu input (Bq m- ) in 1968.

Results and discussion

210 The concentration profiles of Pb and 239+24opu were analysed with the model for each core. The availability of the 239+24opu profiles proved to be important for the determination of mixing rates. If 210Pb profiles were analysed sepa- rately, best fits would often result in mixing rates that were very high (above 10 cm2 y1). Such high mixing rates effectively remove any structure in the

239+24opu sediment profile above the mixing depth, and the comparison between predicted and observed 239+24opu fata thus proved very valuable. It was thus not possible to select values for mixing depths and mixing rates directly from computer optimisation due to the variable and unreliable results produced by this approach. Instead, these values were selected based on best judgement and visual inspection of the agreement between observations and model pre- dictions.

210 Fig. 2 illustrates the result of the analysis of unsupported Pb and 239+24opu concentration profiles from one of the sediment cores, Station No. 20, collected in the near-zone of the Thule accident area. For this core the sedimentation rate is determined at a value of 0.26 cm y1 ± 9% (1 sd), the mixing depth at a value of 5 cm, the mixing rate at a value of 1 cm2 y1, the 210Pb input rate at a 2 1 value of 220 Bq nr y , and the single 239+24opu input in 1968 at a value of 71 kBq nr2. Furthermore, the observed 137Cs profile for the core is included for comparison.

The sediment parameters estimated from the 10 cores from the Thule accident area are shown in Table I listing the mixing depths and mixing rates, the sedimentation rates and the corresponding uncertainties (1 sd), the 210Pb input 137 rates, the 239+24opu inventories and the Cs inventories. The mixing parame- ters are generally consistent across all sediment cores with a mixing depth of 5 cm and a mixing rate of 1 cm2 y1. The mean sedimentation rate is 0.3 cm y1, varying from 0.2 to 0.4 cm y1. The 210Pb input rate varies from 74 to 460 Bq m-2 y1 with a mean value of 261 Bq nr2 y1. This value agrees with the average deposition flux of 210Pb of about 1 dpm cm-2 y1 (= 170 Bq nr2 y1) found in sedi- ments at temperate latitudes. The atmospheric flux of 210Pb is lower at north- ern latitudes, but at the Thule area sedimentation is high due to particle

123 STUK-A145

THULE SEDIMENT Core 98842 from 1991

UnauppoKM U»*-210

S 10 15 20 a 0 1 10 13 n 25

11 i" 3.06*00 J Mwwdmtfonf 1-0£*Q0 2J2E-O2 7.1 £-0* Pw-23» roul (84M21

Figure 2. Observed and calculated radionuclide concentration profiles for sediment core no. 98842 including estimated parameter values (Station No. 20, coordinates: 76°30.79'N, 69° 17.18 W). transport from land by run off. The 137Cs inventories are closely associated with the 210Pb input rates for the ten sediment cores: the two sets of data show a strong correlation of 0.9 (P<0.001). Actually both data sets are highly corre- lated with the rates of sedimentation (correlation coefficients of 0.8, P<0.01). These data are presented in Fig. 3 showing the 137Cs inventories and the 210Pb input rates versus the sedimentation rates. The 239+24opu inventories vary sig- nificantly from 2 to 71 kBq nr2 with a mean value of 15 kBq m2.

124 STUK-A145

Table I. Parameter values from sediment samples collected from the Thule area.

Station Mix. Mix. Sedim. Uncertainty Pb-210 Pu-239 Cs-137 No. depth rate rate rate Sample (cm) (cm2/y) (cm/y) (1SD, %) (Bq/m2/y) (Bq/m2) (Bq/m2) No. 15-98832 3 1 0.20 18 140 4400 650 16-98834 5 1 0.40 8 460 21000 1720 17-98836 5 1 0.18 20 74 2200 80 18-98838 5 1 0.18 29 110 2600 320 19-98840 5 1 0.22 10 280 5700 1030 20-98842 5 1 0.26 9 220 71000 1050 21-98844 5 1 0.28 12 350 16000 960 23-98848 5 1 0.21 14 330 7600 980 24-98850 5 1 0.38 15 370 7300 1130 25-98852 5 1 0.34 10 280 12000 1290 Mean 4.8 1 0.27 15 261 14980 921 SD (%) 13 0 31 44 48 137 51

Conclusions

Analyses of radionuclide profiles in sediment cores collected in 1991 at Thule in Greenland have provided information on processes that occur in the sedi- 210 ments. The radionuclides include the naturally occurring Pb, 239+24opu origi- nating from the aircraft accident in 1968, and 137Cs originating from nuclear weapons testing. The processes include mixing of the surface sediments mainly from biological activity and burial of sediments due to particle scavenging. These processes influence the time scale and the extent to which the pluto- nium contamination is in contact with seawater and thus available for further dispersion. The quantified description of these processes is necessary for nu- merical modelling of the impact of radioactive contamination of the marine environment. These processes are of particular importance for the tranuranic elements due to the relatively high radiotoxicity, the long physical half lives and the sediment-reactive properties of these elements.

125 STUK-A145

2000 500

400 1500

300 f xCs-137 — 1000 • R)-210 200 -

X • 500 a 100 H

0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.5O Sedimentation rate (cmJy)

Figure 3. I37Cs depositions and 210Pb input rates shown versus sedimentation rates for the ten sediment cores from the Thule accident area. The correlation coefficient for both of the two data sets equals 0.8.

The radionuclide profiles have been analysed with a numerical model to iden- tify values of parameters describing the sediment processes. The average pa- rameter values for the Thule area are: 0.3 cm y1 for the sedimentation rate, 5 cm for the mixing depth and 1 cm2 y1 for the mixing rate. It was found that the mixing parameters were not correctly identified from interpretation of the 210 unsupported Pb profile alone but that information from the 239+24opu profile was necessary. This stresses the need for caution when interpreting 210Pb profiles where no other information is available.

References

Christensen, E. A., 1982. Model for Radionuclides in Sediments Influenced by Mixing and Compaction. J. Geophys. Res., 87: 566-572.

126 STUK-A145

Cs DATING OF LAMINATED SEDIMENTS IN SWEDISH ARCHIPELAGO AREAS OF THE BALTIC SEA

M. Meili1, P. Jonsson1 & R. Carman2

1 Uppsala University, Inst. of Earth Sciences, , Vil- lavagen 16, S-752 36 Uppsala, Sweden 2Stockholm University, Dept. of Geology and Geochemistry, S-106 91 Stockholm, Sweden

In deep off-shore areas of the Baltic Sea, sediment accumulation rates are typically on the order of one or a few millimeters per year, and even less in consolidated sediments, based on lamina counts and radiometric dating (Ignatius 1958, Jonsson et al. 1990, Jonsson 1992). In lacustrine and marine basins, the highest sedimentation rates are usually found in the deepest part, since sediments and associated contaminants are known to be gradually "focused" from shallow to deep areas by resuspension. Accordingly, net sedi- mentation in coastal areas is usually low or absent due to strong erosion forces. On the other hand, coastal sediments are likely to be important in con- trolling the fate and turnover of contaminants that are released into coastal waters. Since little is known about the turnover of coastal sediments, in par- ticular for heterogeneous semi-enclosed areas such as the Baltic archipelagos, we recently have initiated a study of sediment accumulation rates, with a fo- cus on areas where erosion is likely to be minimal. The study is part of a pro- ject focusing on the relationship between eutrophication and contaminant cy- cling (EUCON).

88 sediment cores were collected during summer 1996 from accumulation bot- toms of 18 more or less protected bays in archipelago areas along the Swedish coast of the Baltic Sea. Hydrodynamic boundaries were determined from nau- tical charts, and the sampling sites within each area were selected after a comprehensive survey using a high-frequency echo-sounder and a side-scan sonar. Sedimentation rates were determined with two independent methods: from 137Cs measurements at intervals of usually 1 cm, and from lamina counts in vertically sectioned cores.

127 STUK-A145

Laminated sediments were found to be common and to cover a substantial fraction of most areas, even at water depths <20 m. In these sediments, bio- turbation can be neglected as a confounding factor for radiometric dating. Dating was further facilitated by comparatively high concentrations of 137Cs (0.3 to 2 Bq g'1 dw). Distinct vertical profiles of 137Cs originating from the Chernobyl fallout (1986) and to a minor extent from testing (1963) allowed dating with exceptional accuracy, in some cases even with sea- sonal resolution (Fig. 1). Radiometric dating usually coincided with lamina counts, suggesting annual varves. The lamination structure indicates fairly constant sedimentation conditions over adjacent years. High mean sedimenta- tion rates of 0.5 to 5 cm a"1 were observed, based on the location of the Cherno- byl horizon at depths of 5 to 50 cm below the sediment surface.

This implies that the sedimentation in semi-enclosed coastal areas can exceed the sedimentation in deep off-shore areas of the Baltic Proper by up to two orders of magnitude, if local conditions favour inflow and retention of settling particles. Similar patterns have recently been observed also in the Gulf of Finland (Vallius et al., 1998), but were partly attributed to river input, which is negligible in most of our areas. We attribute our findings to resuspension in more exposed coastal areas, in combination with a gradual land uplift exposing sediments to erosion due to ongoing crustal rebound after deglaciation (cf. Jonsson, et al., 1990). This is supported by the fact that the sedimentation rates along the Swedish coast are not only manyfold higher than those in off- shore areas, but also higher than in the Gulf of Finland (0.2-2 cm a1 including coastal sediments, Vallius et al., 1998), where land uplift is slower.

The steep concentration gradients of 137Cs extending sometimes <0.1 cm after 10 years suggest that, in contrast to many lacustrine sediments in the sur- rounding shield areas, the Cs migration by diffusion is minimal, despite anoxic conditions and brackish water in these laminated sediments. As a re- suit, Cs can be used in the Baltic Sea as a particle tracer to study vertical sediment mixing by bioturbation or horizontal sediment transport after resus- pension. Our three-dimensional survey of l Cs in coastal sediments may also provide valuable information on time scales of sediment resuspension and focusing, on the origin and fate of 137Cs in the Baltic Sea, and on the water exchange in Archipelago areas.

128 STUK-A145

Cs-137 (Bq/g dw) 0.0 0.5 1.0 1.5

10

50 60% 70% 80% 90% 100% Water content (%ww)

Cs-137 (Bq/gdw) 0.0 0.5 1.0 1.5

10

20

30

40

50 '

60 Slat 96 D 70 •

60% 70% 80% 90% 100% Water content (%ww)

Figure 1: Examples of vertical profiles of• 137,Cs (bold) and water content (thin) in coastal sediments from two Baltic Sea archipelago areas (Gdlnan and Sldt- baken). In the lower figure, seasonal changes in both variables are evident, which coincide with an annual lamination pattern.

129 STUK-A145

Acknowledgements

We acknowledge the excellent work by the crew or R/V Sunbeam as well as Pia Holmberg and numerous other field and laboratory assistants. This study was financially supported by the EUCON project of the Swedish Environmental Protection Agency.

References

Jonsson, P., Carman, R. & Wulff, F., 1990. Laminated sediments in the Baltic - a tool for evaluating nutrient mass balances. Ambio, 19: 152-158.

Jonsson, P., 1992. Large-scale changes of contaminants in Baltic Sea sedi- ments during the twentiath century. Acta Univ. Ups., 407, 52 p., ISBN 91-554- 2997-1.

Ignatius, H., 1958. On the rate of sedimentation in the Baltic Sea. Bull. Comm. Geol. Finland, 180: 130-145.

Vallius, H., Kankaanpaa, H., Niemisto, L. & Sandman, O., 1998. Sedimenta- tion and within-basin variations in the Gulf of Finland as determined by Cs tracer. In: Ilus, E. (ed.), Dating of sediments and determination of sedimenta- tion rate. Proceedings of a seminar held in Helsinki 2-3 April 1997. Report STUK-A145: 131-135. ISBN 951-712-226-8.

130 STUK-A145

SEDIMENTATION AND WITHIN-BASIN VARIATIONS IN THE GULF OF FINLAND AS DETERMINED BY 137Cs TRACER

Henry Vallius", Harri Kankaanpaa5, Lauri Niemistob and Olavi Sandman0

aGeological Survey of Finland, POB 96, FIN-02151, Espoo, Finland bFinnish Institute of Marine Research, POB 33, FIN-00931, Helsinki, Finland cMikkeli Water and Environment District, POB 77, FIN-50101 Mikkeli, Finland

Sedimentation rates at 98 stations in several coastal and open sea areas in the Gulf of Finland were determined using 137Cs distribution in sediment profiles. The 137Cs activities of the sediment cores were measured using direct gam- maspectrometric measurements (Kyzyurov et al., 1994) either on board re- search vessels or at the respective institutes. Relative rates were calculated. The results show that annual surficial sedimentation rate varies between 0.05 and 1.94 cm a"1 (Fig. 1), corresponding to a dry matter flux of 0.01 - 0.30 g cm- 2a1 (Fig. 2). The mean sedimentation rate for all studied cores (n = 98) was 0.6 cm a1 and the mean dry matter accumulation rate (n = 45) was 0.14 g cm-2a'. The accumulation rates were highest at bottoms of recent, muddy sediments. The 137Cs activities in cores from the studied area varied between 0.04 and 2.4 Bq g-1 wet weight (Fig. 3). The highest activities are found in layers corre- sponding to the year 1986 from areas mostly affected by the fallout cloud of the Chernobyl nuclear power plant accident. The areal distribution of 137Cs in sediments is correlated with that measured on land but the activities and dis- tribution pattern in sea sediments has been modified by drainage from land and secondary transport in the sea (Fig. 4).

Mixing of the sediment strata was found to be considerable in the Gulf of Fin- land. This is mostly caused by bioturbation, as the oxygen content in near bot- tom waters was good during the whole study period. This can be seen in the pattern of the 137Cs profiles.

131 STUK-A145

/ nmnm Finland / ^ P O O O OOO •- N N • «. o>t»~ra V (cm/yr) r^UHb • -t Vv, Russia • 6& • • • ; •* •<*% ^r^, Gulf of Finland • • • "*• ' 1 rO A.

Estonia

Figure 1. Measured average sedimentation rates in different parts of the Gulf of Finland

Russia

Figure 2. Measured dry matter accumulation rates in the Gulf of Finland.

132 STUK-A145

inland /

Russia

Figure 3. Maximum 13TCs activity in the sediment slices.

inland /

o o o o oo A (lcBq/m') Russia

Figure 4. Total l37Cs activity at the sampling stations.

133 STUK-A145

There seemed to be no correlation between maximum 137Cs concentration in cores and bottom depth, suggesting that the main factor affecting the occur- rence of 137Cs is the distribution pattern of Cs after the Chernobyl accident. Bottom depth and sedimentation rate did not correlate, but a slight tendency toward increasing rate with decreasing depth especially close to river outlets was observed.

Figure 5. Depth of the sediment containing the Chernobyl peak of 137Cs (white line) at two stations in the Gulf of Finland (GF2 = mid Gulf of Finland, hori- zontal scale 2.30 kms, F41 = eastern Gulf of Finland, horizontal scale 1.85 kms) and content of partially leachable copper.

134 STUK-A145

Within-basin variation of sedimentation was studied in two basins, one at Lavansaari, eastern Gulf of Finland and the other in mid Gulf of Finland, off Kasmu peninsula in Estonia. Transects of approximately 1 nautical miles length and with 7-9 sampling sites were taken. The sedimentation since 1986 could be counted from the Chernobyl peak. At Lavansaari sedimentation had been uniform in the whole basin through that time interval while at the sta- tion in mid Gulf of Finland sedimentation had varied a lot (range 1.5-8 cm in 9 years). There sedimentation had been fastest in the middle of the basin (6.5 - 7.5 mm a1) and slowest at the borders (1.5 - 2.5 mm a1), (Fig. 5). When com- paring the 137Cs dating curve of the basins to chemically analysed metal data a clear correlation can be seen. The basin study and the variations in within- basin sedimentation show the necessity of thorough investigations of the sedimentary environments in any studies comprising recent sediments in the Gulf of Finland.

The direct 137Cs measurement technique can be successfully used in the Gulf of Finland, southern Gulf of Bothnia and in parts of the main Baltic Sea because of high radiocaesium activities and relatively high accumulation rates. The method is also a good tool for evaluation of stations suitable for sediment studies and monitoring.

References

Kankaanpaa, H., Vallius, H., Sandman, O. and Niemisto, L., 1997. Determi- nation of recent sedimentation in the Gulf of Finland using Cs-137. Oceanol. Acta, 20; 6: 823-836.

Kyzyurov, V., Mikheev, J., Niemisto, L., Winterhalter, B., Hasanen, E. and Ilus, E., 1994. A simple method for the determination of deposition rates of recent sediments based on cesium-137 activity following the Chernobyl acci- dent. Baltica, 8: 64-67.

135 STUK-A145

EVALUATION OF SEDIMENTATION RATE AT TWO SAMPLING STATIONS IN THE GULF OF FINLAND BASED ON Pb-210, Cs-137 AND Pu 239,240 PROFILES IN SEDIMENTS

Erkki llus, Jukka Mattila, Seppo Klemola and Tarja K. Ikaheimonen

STUK, Radiation and Nuclear Safety Authority, Research and Environmental Surveillance, P.O.Box 14, FIN-00881 Helsinki, Finland

Introduction

Determination of sedimentation rate plays an important role in material bal- ance and model calculations of coastal seas and other water bodies. The Baltic Sea offers an exceptionally good opportunity to study sedimentation rate, be- cause the concentration peaks of Cs and Pu caused by nuclear weapons tests in the 1960s and the Chernobyl disaster in 1986 are easily detectable in corre- sponding layers of the bottom sediments. In addition, the rate of sedimentation is much higher in the Baltic Sea than in the oceans and many other coastal seas, and the lack of benthic animals on its large bottom areas, because of the poor oxygen conditions there, results in the formation of an undisturbed sedi- mentary package, enabling the sampling of sediment laminae in an undis- turbed historical order.

The Radiation and Nuclear Safety Authority (STUK) has long traditions in monitoring radionuclides in the Baltic Sea environment. Sediment samples have been taken in co-operation with the Finnish Institute of Marine Research (FIMR) since the late 1960s for the analysis of artificial radionuclides. Primar- ily, the samples were not taken for dating purposes, but the concentration profiles of Cs-137 and Pu-239,240 were also used to evaluate sedimentation rate (e.g. Tuomainen et al., 1986).

The accident at the Chernobyl NPP in 1986 left a clear marker in the sedi- ments in the Baltic Sea and roused new interest in the study of sedimentation rate. However, it was conspicuous that the datings made in the late 1980s and

136 STUK-A145

early 1990s using the Chernobyl-derived Cs-137 peaks seemed to result in significantly higher sedimentation rates than those made using the traditional Po-210/Pb-210 method. Therefore, when the NKS/EKO-1 Project began, the aims of our subproject were to determine sedimentation rates in different parts of the Baltic Sea, using Pb-210, Cs-137 Pu-238 and Pu-239,240 profiles in sediments, and to consider differences in the results obtained with different methods.

Material and methods

The project was carried out in co-operation with the Finnish Institute of Ma- rine Research, and the samples were mainly taken on board the Institute's Research Vessel ARANDA. In 1995, sediment samples were taken at eight stations in the Gulf of Finland and the Gulf of Bothnia. In 1996, samples were taken on three cruises of the RV ARANDA from 43 sampling stations situated in the Baltic Proper, the Bothnian Bay, the Bothnian Sea and the Gulf of Fin- land.

The samples were taken with a GEMINI Twin Corer designed by Mr. Lauri Niemisto (FIMR), consisting of two parallel coring tubes with a diameter of 80 mm. The cores were sliced into 1-cm slices down to 20, 16 or 10 cm with the precise slicing device of the GEMINI Corer (cf. Ilus 1996). The rest of the core was sliced at 2-cm intervals to a depth of about 30 cm. The parallel slices were combined and the samples were frozen and freeze-dried before analysis.

This paper is based on results from two sampling stations - one (F41) situated in the eastern part of the open Gulf of Finland, the other (LOV3A) in the ar- chipelago zone of the eastern Gulf of Finland (Table I).

Table I. Codes, sampling dates, positions and bottom depths of the sampling stations. Station Date Lat(N) Lon(E) Depth (m) F41 25.4.1995 60o07.04' 28°03.86' 51 LOV3A 8.6.1995 60°22.20' 26°23.10' 18

137 STUK-A145

The L0V3A station is situated in the discharge area of the Loviisa nuclear power plant. At this station, the sediment is very soft, black, sulphidic "gyttja clay" and the bottom is influenced by annual anoxic periods in the hy- polimnion. At station F41, the type of sediment is more solid sulphidic clay.

Pb-210 activities were analysed using gammaspectrometry (e.g. Appleby et al., 1986), taking into account the self-absorption of gamma rays in the samples of different density and height (Klemola et al., 1997). When the unsupported part of Pb-210 was calculated, Ra-226 activity was subtracted from the total Pb-210 activity measured. Ra-226 activities were obtained using Bi-214 (609.3 keV), Pb-214 (295.2 keV) and Pb-214 (351.9 keV) photon intensities measured by gammaspectrometry. Pu analyses were performed according to the method of Taipale and Tuomainen (1985), and Cs-137 was measured by gamma spec- trometry at the same time as the Pb-210 and Ra-226. Po-210 was analysed using the Hasanen method (1977), modified at our laboratory.

Dry matter accumulation rates were calculated using 1986 as the reference year for the Cs-137 peak and for the changed ratio of Pu-238 / Pu-239,240. In general, the Chernobyl disaster caused only a slight increase in the plutonium concentrations of Baltic Sea sediments, but it caused a distinct change in the Pu- 238 / Pu-239,240 ratio, which can be used as a time marker for sedimentation studies. Both 1963 and 1964 were used as references for the Pu-239,240 activ- ity peaks in sediment profiles. Accumulation rates were calculated assuming that the observed activity concentration peak is in the middle of the sediment slice. The limits of error are calculated to the upper and lower edge of the slice.

In the case of the Pb-210 method, we used two models to ensure reliable re- sults and to compare the results from these different types of model. The con- stant rate of supply model (CRS model) is described in detail by Appleby and Oldfield (1978) and by Robbins (1978). The constant flux: constant sedimenta- tion rate model (CF:CS model) is described by Robbins (1978), among others. The error bars for the CRS sedimentation rates are based on calculated mea- surement errors. With these errors, it is possible to calculate confidence inter- vals for the dates (Binford, 1990) and thus their effects on the sedimentation rates.

The sedimentation rates in [mm y1] were calculated from the dry matter ac- cumulation rates, using calculated porosities at depths of 0-1, 9-10 and 19-20 cm. The calculation of porosities was based on the water content measured in

138 STUK-A145

sediments. Because of the low salinity of the Baltic Sea water, the effect of salinity on the calculation of porosities is negligible. For dry matter density, we used the value 2.5 g cm-3 (e.g. Niemisto et al., 1978).

Results and discussion

The vertical distribution of unsupported Pb-210, Cs-137 and Pu-239,240 in the sediments at the sampling stations L0V3A and F41 are shown in Figs 1 and 2. The activity concentrations of unsupported Pb-210 were quite close to each other at both stations. At station L0V3A, Pb-210 activities were also analysed using the traditional Po-210/Pb-210 method and the results were in good agreement with those obtained in gammaspectrometric measurements (Fig. 3).

At both stations, a clear peak of Chernobyl-derived Cs-137 occurred at a depth of 7-8 cm, but at the coastal station L0V3A the peak concentration was much higher. The Pu-239,240 peak caused by the nuclear weapons tests was also clear at station L0V3A. At station F41 the peak was less distinct, but the activity level was more or less equal to that of the Loviisa station.

At station F41 the highest sedimentation rate [g cnr2 y1] was obtained using the Cs-137 method (Fig.4). The values obtained with the CRS method (calculated for both the last 10 and the last 30 years) and the CF:CS method agreed very well, but they were much lower than those based on the Cs-137 peak. The lowest sedimentation rate was obtained with the method based on the Pu-239,240 peak.

At the station LOV3A, the sedimentation rates calculated using different methods for the last 10 years give more or less equal results within the limits of error, but are somewhat higher than those calculated for the last 30 years. The sedimentation rate obtained with the CF:CS method agrees better with the latter ones (Fig.4). The highest sedimentation rate was obtained with the CRS method, calculated for the last 10 years. When CRS-method was used, it was also possible to calculate the sedimentation rate at station LOV3A from 1940 to 1993 (Fig 5).

139 V) c

[Bqkg'd.w.] [Bqkg'd.w.] 200 400 0 2000 4000 0 0 H 1 0

f •8

Figure 1. Vertical distribution of unsupported Pb-210, Cs-137 and Pu-239,240 activity concentrations in the bottom sediments at station L0V3A (standard deviations omitted). [Bqkg'd.w.] [Bq kg1 d.w.] 400 800 1600

unsupported Pb-210

Figure 2. Vertical distribution of unsupported Pb-210, Cs-137 and Pu-239,240 activity concentrations in the bottom sediments at station F41 (standard deviations omitted).

CO c

I > STUK-A145

[Bq kg'1] 0 200 400 600 n

5 — 10 T I 15 I g- 20 - •a Pb-210 25 - 30 Po-210 35-! -f

Figure 3. Vertical distribution of measured Po-210 and Pb-210 activities in the bottom sediments at station LOV 3A (standard deviations omitted).

F41 sedimentation rate [g cm'2 y'1] LOV3A sedimentation rate [g cm'2 y"1]

0.00 0.05 0.10 0.15 0.00 0.05 0.10 0.15

Cs-137 Cs-137 %,:•;•.•-:*; ••.-....;;-4-

Pu-ratio Pu-ratio

CRS 0-10 CRS 0-10 years yean

Pu-239,240 Pu-239,240

CRS 0-30 years • +

CF:CS CF:CS 1

Figure 4. Sedimentation rates [g cm^y1] and their error levels at stations F41 and LOV3A, calculated using the Cs-137, Pu-238/Pu-239,240, Pu-239,240, CRS and CF.CS methods.

142 STUK-A145

[gcn-TV]

0.160 T 0.140 - 0.120 t 0.100 0.080 0.060 " 0.040 + 0.020 i 0.000 i 1940 1950 1960 1970 1980 1990 year

Figure 5. Sedimentation rate [g cm2y1] at station LOV3A, calculated using the CRS method from 1940 to 1993. The straight line describes the roughly estimated increase in sedimentation rate.

Many facts support the assumption that the sedimentation rate has increased in the sea area off Loviisa during the last 10-15 years. For example, the annual phytoplankton primary production is now two to three times as high as it was thirty years ago (Ilus et al., 1997). This increase has been associated with the general eutrophication that has taken place in the whole Gulf of Finland. The higher primary production has increased the amounts of sinking participate matter, especially in coastal shallow-water areas, such as at station LOV3A. Correspondingly, when considering sedimentation rates calculated with the CRS method for the different time periods (0-10, 10-20 and 20-60 years) it seems that sedimentation has been clearly faster during the last 10 years than earlier (Table II). Comparison of the results obtained with methods based on the Pu- 239,240 and Cs-137 peaks, and the Pu-ratio, confirms that the sedimentation rate has increased at station LOV3A.

At station F41, the values calculated using the CRS method for different time periods and those obtained with the CF:CS method give more or less equal re- sults (Table II). The measured peak values of Cs-137 and Pu-ratio occurred in different sediment layers, and consequently resulted in different sedimentation rates, although both are based on same reference year of the Chernobyl accident. The sedimentation rates for the period 1963-1995 based on the Pu-239,240 method are clearly lower than values for the period 1986-1995 based on the Cs- 137 or the Pu-ratio-method. Such a big difference looks extremely unlikely in view of the fact that the period 1963-1995 includes the period 1986-1995.

143 STUK-A145

Table II. Sedimentation rates in [g cm-2 y1] and [mm y1] calculated using differ- ent methods for different periods of time at stations F41 and LOV3A. Porosities of sediments at depths 0-1, 9-10 and 19-20 cm were used in calculating the sedimen- tation rates in [mm y1].

Station F41, 26.4.1995 [gcm-V1! [mmy1] years aver. min max 0-lcm 9-10cm 19-20cm CRS method 0-10 0.082 0.075 0.092 9.4 2.9 2.6 10-20 0.080 0.068 0.098 9.1 2.8 2.5 20-60 0.075 0.069 0.082 8.5 2.6 2.4 0-20 0.081 0.077 0.087 9.2 2.8 2.6 0-30 0.083 0.079 0.088 9.4 2.9 2.6 0-40 0.081 0.078 0.086 9.2 2.8 2.6 0-50 0.081 0.077 0.085 9.2 2.8 2.6 0-60 0.077 0.073 0.082 8.7 2.7 2.4 CF:CS method 0.078 8.9 2.7 2.5 Cs-137 ref.year 1986 0.129 0.117 0.129 14.6 4.5 4.1 Pu-ratio ref.year 1986 0.105 0.100 0.105 11.9 3.7 3.3 Pu-239,240 refyear 1963 0.062 0.057 0.062 7.0 2.2 1.9 ref.year 1964 0.064 0.058 0.064 7.2 2.2 2.0 Station LOV3A 8.6.1995 [gcm-V1] [mmy-1] years aver. min max 0-lcm 9-10cm 19-20cm CRS method 0-10 0.101 0.091 0.113 18.1 6.1 5.0 10-20 0.078 0.066 0.093 13.9 4.7 3.9 20-60 0.070 0.064 0.078 12.6 4.3 3.5 0-20 0.089 0.084 0.095 16.0 5.4 4.4 0-30 0.084 0.080 0.089 15.1 5.1 4.2 0-40 0.079 0.075 0.084 14.3 4.8 4.0 0-50 0.078 0.074 0.083 14.1 4.8 3.9 0-60 0.077 0.072 0.082 13.8 4.7 3.8 CF-.CS method 0.074 13.3 4.5 3.7 Cs-137 ref.year 1986 0.088 0.080 0.088 15.8 5.3 4.4 Pu-ratio ref.year 1986 0.088 0.080 0.088 15.8 5.3 4.4 Pu-239,240 ref.year 1963 0.065 0.062 0.067 11.6 3.9 3.2 refyear 1964 0.067 0.064 0.070 12.0 4.1 3.3

144 STUK-A145

Furthermore, it is contradictory that the average sedimentation rate calculated with the CF:CS method over a very long period of time (about 110 years) is higher than the sedimentation rate obtained with the Pu-239,240 method. Considering all these facts it is likely that some kind of sediment redistribution has occurred during the period 1963-1986. Redistribution of sediments is possible, taking into account the pelagic location and the bottom depth (51m) at this station (cf. e.g. Ignatius et al., 1981).

The sedimentation rates in millimetres per year have been calculated from the dry matter accumulation values. The calculations are based on three different porosities, representing slice depths of 0-1, 9-10 and 19-20 cm. If the porosity of the uppermost layer is used in the calculations, the sedimentation rates are significantly higher than if the porosities of the older sediment layers are used. Thus, surface sedimentation rates higher than 10 mm are clearly overestimated because the porosity is valid only for the 0-1 cm sediment layer. The uppermost sediment layers are very loose at both stations and do not represent real compressed and solid sediment layers. Thus, the calculations based on the porosities of the deeper layers provide better estimates for longer time periods.

Conclusions

The comparisons presented above indicate that different methods may result in different sedimentation rates, but on the basis of the results obtained from these two sampling stations, at least, it is difficult to show that one method always gives higher or lower results than the others. Poor fulfilment of assumptions may often cause problems when the CRS method is used in the Baltic Sea. In addition, disturbed and poor-quality sediment samples can lead to false conclu- sions despite of the method used (Ilus, 1996). Therefore, the use of parallel inde- pendent methods is highly recommendable.

It is always important to remember the period for which sedimentation rate is estimated. In general, the differences decrease when the sedimentation rate is calculated for longer periods. Furthermore, the error limits of sedimentation rate are much broader in estimates made for recent or shorter periods than in those made for longer periods. The high sedimentation rates obtained in many studies on the basis of the Chernobyl-derived Cs-137 peak are probably caused by these facts.

145 STUK-A145

In this study, it was possible to prove a recent increase in sedimentation rate at station L0V3A using different independent methods. At station F41, the dis- crepancies between different methods could be explained by sediment redistri- bution during the period 1963-1986. For both stations it was possible to estimate the general level of sedimentation rate for different periods.

Acknowledgements

We received financial assistance from the NKS (Nordic Nuclear Safety Re- search). We are grateful to the Finnish Institute of Marine Research and the crew of R/V Aranda for excellent cooperation in sampling.

References

Appleby, P.G. and Oldfield, F., 1978. The calculation of lead-210 dates assum- ing a constant rate of supply of unsupported Pb-210 to the sediment. Catena, 5: 1-8.

Appleby, P.G., Nolan, P.J., Gifford, D.W., Godfrey, M.J., Oldfield, P., Ander- son, N.J. and Battarbee R.W., 1986. Pb-210 dating by low background gamma counting. Hydrobiologia, 143: 21-27.

Berner, R.A., 1978. Principles of Chemical Sedimentology. McGraw-Hill Book Company, 240 pp.

Binford, M. W., 1990. Calculation and uncertainty analysis of 210Pb dates for PIRLA project lake sediment cores. J. Paleolimn., 3: 253-267.

Hasanen, E., 1977. Dating of sediments, based on Po-210 measurements. Ra- diochem. Radioanal. Letters, 31(4-5): 207-214.

Ilus, E., 1996. Evaluation of sediment sampling devices and methods used in the NKS/EKO-1 Project. Report NKS/EKO-1(96)TR-1. Riso.

146 STUK-A145

Ilus, E., Ikaheimonen, T. K. and Klemola, S., 1997. Nuclear power and the environment in Finland - with special emphasis on the marine environment. Proceedings of the 11th Meeting of the Nordic Society for Radiation Protection and the 7th Nordic Radioecology Seminar, Reykjavik, Iceland, August 26-29, 1996, 431-436.

Klemola, S., Mattila, J. and Ikaheimonen, T. K., 1997. Determination of Pb- 210 in sediment samples by gamma ray spectrometry. Proceedings of the 11th Meeting of the Nordic Society of Radiation Protection and the 7th Nordic Ra- dioecology Seminar, Reykjavik, Iceland, August 26-29, 1996, 449-452.

Robbins, J.A., 1978. Geochemical and geophysical applications of radioactive lead. In: J. O. Nriagu (ed.).The biogeochemistry of lead in the environment. Elsevier, Amsterdam, 285-393.

Taipale, T.K. and Tuomainen, K., 1985. Radiochemical determination of plu- tonium and americium from seawater, sediment and biota samples. Report STUK-B-VALO 26, Finnish Centre for Radiation and Nuclear Safety.

Tuomainen, K., Ilus, E. and Taipale, T.K., 1986. Long-lived radionuclides in the sediments of the Gulf of Bothnia. Publications of the Water Research Insti- tute, National Board of Waters, Finland, 68: 91-95.

147 STUK-A145

DEVELOPMENT OF INSTRUMENTATION FOR 210Pb DATING METHOD

P. Reinikainen2, J. Merilainen1, I. Rekikoski2, A. Virtanen2, A. Witick1 and J. Aysto2 department of Physics, University of Jyvaskyla 2lnstitute for Environmental Research, University of Jyvaskyla

The radioactivity and dating laboratory project was started in 1993. Alpha- and gamma-ray spectroscopy systems and routines for a study of environ- mental samples have been developed. Particular interest has been focused to the 210Pb dating for lake sediments.

So far, about 40 sediment profiles from all around Finland have been analysed and dated using the 210Pb method. In gamma-ray spectroscopy low-background measurements have been performed in a lead shield, in which very radiopure materials have been chosen. During the year 1996 the accuracy of the 210Pb dating compared to other methods was evaluated (Reinikainen et al., 1997). The background reduction was achieved by flushing liquid nitrogen into a measurement chamber to get rid off inherent radon. The alpha measurements in turn were carried out with PIN and PIPS detectors.

Special attention was also paid to preliminary study of the applicability of alpha/beta separative liquid scintillation technique in 210Pb dating. The study was carried out in co-operation with two instrument manufacturers, Wallac and Hidex. The results achieved by liquid scintillator counting were very promising, although there is still a lot to do, especially in the field of sample preparation. Finnish sediment samples contain usually large amount of iron, which impairs the counting resolution significantly by clouding the samples. A typical spectrum of a sediment sample is shown in Fig. 1.

148 STUK-A145

350

Sample 0.26 g dried sediment 300

250 !-»-Beeta

200

2 150

100 -

50 -

0 - Alpha

200 400 600 800 1000 1200 Channel

Figure 1. Spectrum of a sediment sample measured with liquid scintillator (Quantulus) in 21 h.

Reference

Reinikainen, P., Merilainen, J. J., Virtanen, A., Veijola, H. and Aysto, J., 1997. Accuracy of 210Pb Dating in two Annually Laminated Lake Sediments with High Cs Background. Appl. Radiat. Isot., 48, 7: 1009.

149 STUK-A145

STUK-A reports

STUK-A145 Ilus E (ed.). Dating of STUK-A136 Pennanen M., Makelai- sediments and determination of sedi- nen I. & Voutilainen A. Huoneilman mentation rate. Proceedings of a semi- radonmittaukset Kymen laanissa: nar held in Helsinki 2-3 April 1997. Tilannekatsaus ja radonennuste. Hel- Helsinki 1998. sinki 1996. STUK-A144 Rannikko S, Karila KTK, STUK-A135 Hyvarinen J. On the fun- Toivonen M. Patient and population damentals of nuclear reactor safety doses of X-ray diagnostics in Finland. assessments. Inherent threads and Helsinki 1997. their implications. Helsinki 1996.

STUK-A143 Helariutta K, Rantavaara STUK-A134 Ylatalo S, Karvonen J, A, Lehtovaara J. Turvesoiden ja poltto- Ilander T, Honkamaa T, Toivonen H. turpeen radionuklidit. Helsinki 1997 Doserate mapping and search of radio- active sources in Estonia. Helsinki STUK-A142 Auvinen A. Cancer risk 1996. from low doses of . Helsinki 1997 STUK-A133 Rantavaara A. Puu- tavaran radioaktiivisuus. Helsinki STUK-A141 Jokela K, Leszczynski D, 1996. Paile W, Salomaa S, Puranen L, Hyysalo P. Matkapuhelimien ja tuki- STUK-A132 French S, Finck R, asemien sateilyturvallisuus. Helsinki Hamalainen RP, Naadland E, Roed J, 1997 Salo A, Sinkko K. Nordic Decision Conference: An exercise on clean-up STUK-A140 Moring M, Markkula M- actions in an urban environment after L. Cleanup techniques for Finnish a nuclear accident. Helsinki 1996. urban environments and external doses from 137Cs - modelling and cal- STUK-A131 Mustonen R, Koponen H. culations. Helsinki 1997 Sateilyturvakeskuksen tutkimushank- keet 1996 - 1997. Helsinki 1996. STUK-A139 Tapiovaara M, Lakkisto M, Servomaa A. PCXMC. A PC-based STUK-A130 Honkamaa T, Toivonen Monte Carlo program for calculating H, Nikkinen M. Monitoring of airborne patient doses in medical x-ray exami- contamination using mobile units. nations. Helsinki 1997. Helsinki 1996. STUK-A138 Lindell B Boice JD, Sin- STUK-A129 Saxen R, Koskelainen U. naeve J, Rytomaa T. Past and future Radioactivity of surface water and trends of radiation research. Proceed- fresh water fish in Finland in 1991- ings of the seminar at STUK in Hel- 1994. Helsinki 1995. sinki 28 February 1997. Helsinki 1997. STUK-A128 Savolainen S, Kairemo K, STUK-A137 Arvela H, Ravea T. Ra- Liewendahl K, Rannikko S. Radioim- donturvallinen rakentaminen Suomes- munoterapia. Hoidon radionuklidit ja sa. Helsinki 1997. annoslaskenta. Helsinki 1995. STUK-A145

STUK-A127 Arvela H. Asuntojen ra- STUK-A118 Reiman L. Expert judg- donkorjauksen menetelmät. Helsinki ment in analysis of human and organ- 1995. izational behaviour in nuclear power plants. Helsinki 1994. STUK-A126 Pöllänen R, Toivonen H, Lahtinen J, Ilander T. OTUS-reactor STUK-A117 Auvinen A, Castren O, inventory management system based Hyvönen H, Komppa T, Mustonen R, on ORIGEN 2. Helsinki 1995. Paile W, Rytömaa T, Salomaa S, Ser- vomaa A, Servomaa K, Suomela M. STUK-A125 Pöllänen R, Toivonen H, Säteilyn lähteet ja vaikutukset. Hel- Lahtinen J, Ilander T. Transport of sinki 1994. large particles released in a nuclear accident. Helsinki 1995. STUK-A116 Säteilyturvakeskuksen tutkimushankkeet 1994-1995. Mus- STUK-A124 Arvela H. Residential tonen R, Koponen H (toim.). Helsinki radon in Finland: Sources, variation, 1994. modelling and dose comparisons. Hel- sinki 1995. STUK-A115 Leszczynski K. Assess- ment and comparison of methods for STUK-A123 Aaltonen H, Laaksonen J, solar ultraviolet radiation measure- Lahtinen J, Mustonen R, Rantavaara ments. Helsinki 1995. A, Reponen H, Rytömaa T, Suomela M, Toivonen H, Varjoranta T. Ydinuhkat STUK-A114 Arvela H, Castren O. ja varautuminen. Helsinki 1995. Asuntojen radonkorjauksen kustan- nukset Suomessa. Helsinki 1994. STUK-A122 Rantavaara A, Saxen R, Puhakainen M, Hatva T, Ahosilta P, STUK-A113 Lahtinen J, Toivonen H, Tenhunen J. Radioaktiivisen laskeu- Pöllänen R, Nordlund G. A hypotheti- man vaikutukset vesihuoltoon. Hel- cal severe reactor accident in Sosnovyy sinki 1995. Bor, Russia: Short-term radiological consequences in southern Finland. STUK-A121 Ikäheimonen TK, Helsinki 1993. Klemola S, Uus E, Sjöblom K-L. Moni- toring of radionuclides in the vicinities of Finnish nuclear power plants in 1991-1992. Helsinki 1995. STUK-A120 Puranen L, Jokela K, The full list of publications is Hietanen M. Altistumismittaukset available from suurtaajuuskuumentimien hajasätei- lykentässä. Helsinki 1995. STUK - Radiation and Nuclear Safety Authority STUK-A119 Voutilainen A, Mäkeläi- P.O. BOX 14 nen I. Huoneilman radonmittaukset FIN-00881 HELSINKI Itä-Uudenmaan alueella: Tilannekat- Finland saus ja radonennuste. Askola, Lapin- Tel. +358 9 759 881 järvi, Liljendal, Loviisa, Myrskylä, Mäntsälä, Pernaja, Pornainen, Porvoo, Porvoon mlk, Pukkila, Ruotsinpyhtää ja Sipoo. Helsinki 1995, ISBN 951-712-226-8 ISSN 0781-1705

Oy Edita Ab, Helsinki 1998